diff --git "a/Polymer_Chemistry-CRC_Press/Polymer_Chemistry-CRC_Press.json" "b/Polymer_Chemistry-CRC_Press/Polymer_Chemistry-CRC_Press.json" new file mode 100644--- /dev/null +++ "b/Polymer_Chemistry-CRC_Press/Polymer_Chemistry-CRC_Press.json" @@ -0,0 +1 @@ +{"page_0": [["block_0", [{"image_0": "0_0.png", "coords": [0, 0, 472, 728], "fig_type": 0}]], ["block_1", [{"image_1": "0_1.png", "coords": [0, 4, 106, 274], "fig_type": "figure"}]], ["block_2", [{"image_2": "0_2.png", "coords": [1, 521, 100, 727], "fig_type": "figure"}]], ["block_3", ["E\n.\nPolymer\nC Chemistry\n"]], ["block_4", [{"image_3": "0_3.png", "coords": [3, 632, 94, 719], "fig_type": "figure"}]], ["block_5", [{"image_4": "0_4.png", "coords": [5, 16, 95, 103], "fig_type": "figure"}]], ["block_6", [{"image_5": "0_5.png", "coords": [7, 436, 90, 517], "fig_type": "figure"}]], ["block_7", [{"image_6": "0_6.png", "coords": [7, 235, 89, 317], "fig_type": "figure"}]], ["block_8", [{"image_7": "0_7.png", "coords": [113, 26, 396, 206], "fig_type": "figure"}]], ["block_9", [{"image_8": "0_8.png", "coords": [187, 564, 447, 648], "fig_type": "figure"}]], ["block_10", [". Poul C. Hiemenz\nTimothy P. Lodge\n"]], ["block_11", ["\u2019\n\ufb01\u2018\nQ\nC?\n"]], ["block_12", ["LB\n\u20180\nt\no-\n"]], ["block_13", ["an\na\u00bb\n"]], ["block_14", ["SecondEdition\n"]], ["block_15", ["_\n"]], ["block_16", ["E\n(,9...\n"]], ["block_17", ["E\n"]], ["block_18", ["E\n"]], ["block_19", ["E\n"]], ["block_20", ["E\n"]], ["block_21", ["E\n"]]], "page_1": [["block_0", [{"image_0": "1_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "1_1.png", "coords": [83, 56, 384, 202], "fig_type": "figure"}]], ["block_2", [{"image_2": "1_2.png", "coords": [88, 89, 365, 155], "fig_type": "figure"}]], ["block_3", [{"image_3": "1_3.png", "coords": [93, 17, 350, 160], "fig_type": "figure"}]], ["block_4", [{"image_4": "1_4.png", "coords": [100, 21, 356, 93], "fig_type": "figure"}]], ["block_5", ["Polymer\nChemistry\n"]], ["block_6", ["Second Edition\n"]]], "page_2": [["block_0", [{"image_0": "2_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "2_1.png", "coords": [83, 91, 373, 156], "fig_type": "figure"}]], ["block_2", [{"image_2": "2_2.png", "coords": [84, 58, 384, 204], "fig_type": "figure"}]], ["block_3", [{"image_3": "2_3.png", "coords": [93, 17, 351, 161], "fig_type": "figure"}]], ["block_4", ["Polymer\nChemistry\n"]], ["block_5", [{"image_4": "2_4.png", "coords": [104, 21, 356, 90], "fig_type": "figure"}]], ["block_6", [{"image_5": "2_5.png", "coords": [107, 362, 344, 447], "fig_type": "figure"}]], ["block_7", ["Paul C. Hiemenz\nTimothy R Lodge\n"]], ["block_8", ["Second Edition\n"]], ["block_9", ["Taylor & Francis Group, an informa business\n"]], ["block_10", ["CRC Press is an imprint of the\n"]], ["block_11", ["CRC Press\nTaylor & Francis Group\n"]], ["block_12", ["Boca Raton\nLondon\nNew York\n"]]], "page_3": [["block_0", [{"image_0": "3_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["CRC Press\nTaylor & Francis Group\n6000 Broken Sound Parkway NW, Suite 300\nBoca Raton, FL 33487-2742\n"]], ["block_2", ["This book contains information obtained from authentic and highly regarded sources. Reprinted material is quoted\nwith permission, and sources are indicated. A wide variety of references are listed. Reasonable efforts have been made to\npublish reliable data and information, but the author and the publisher cannot assume responsibility for the validity of\nall materials or for the consequences of their use.\n"]], ["block_3", ["For permission to photocopy or use material electronically from this work, please access www.copyright.com (http:/l\nwww.copyright.coml) or contact the C0pyright Clearance Center, Inc. (CCC) 222 Rosewood Drive, Danvers, MA 01923,\n978-750-8400. CCC is a not\u2014for\u2014profit organization that provides licenses and registration for a variety of users. For orga-\nnizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.\n"]], ["block_4", ["No claim to original U.S. Government works\nPrinted in the United States of America on acid-free paper\n10 9 8 7\n"]], ["block_5", ["No part ofthis book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or\nother means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any informa\u2014\ntion storage or retrieval system, without written permission from the publishers.\n"]], ["block_6", ["\u00a9 2007 by Taylor 8: Francis Group, LLC\nCRC Press is an imprint of Taylor 8: Francis Group, an Informa business\n"]], ["block_7", ["Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for\nidentification and explanation without intent to infringe.\n"]], ["block_8", ["and the CRC Press Web site at\nhttp://www.crcpress.com\n"]], ["block_9", ["Visit the Taylor 8; Francis Web site at\nhttp://www.taylorandfrancis.com\n"]], ["block_10", ["International Standard Book Number-10: 1\u201457444\u2014779\u20143 (Hardcover)\nInternational Standard Book Number\u201413: 978\u20141\u201457444-779\u20148 (Hardcover)\n"]], ["block_11", ["QD381.H52 2007\n547.7--dc22\n2006103309\n"]], ["block_12", ["Hiemenz, Paul C., 1936\u2014\nPolymer chemistry / Paul C. Hiemenz and Tim Lodge. 2nd ed.\np. cm.\nIncludes bibliographical references and index.\nISBN-13: 978-1-57444\u2014779v8 (alk. paper)\nISBN-10: 1\u201457444\u2014779\u20143 (alk. paper)\n1. Polymers. 2. Polymerization. I. Lodge, Tim. 11. Title.\n"]], ["block_13", ["Library of Congress Cataloging-in-Publication Data\n"]]], "page_4": [["block_0", [{"image_0": "4_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["essential, and are introduced as needed.\nThe philosophy underlying the approach in this book is the same as that in the first edition, as\nlaid out in the previous preface. Namely, we endeavor to develop the fundamental principles,\nrather than an encyclopedic knowledge of particular polymers and their applications; we seek to\nbuild a molecular understanding of polymer synthesis, characterization, and properties; we\nemphasize those phenomena (from the vast array of possibilities) that we judge to be the most\ninteresting. The text has been extensively reorganized and expanded, largely to re\ufb02ect the\nsubstantial advances that have occurred over the intervening years. For example, there is now\nan entire chapter (Chapter 4) dedicated to the topic of controlled polymerization, an area that has\nrecently undergone a revolution. Another chapter (Chapter 11) delves into the viscoelastic\nproperties of polymers, a topic where theoretical advances have brought deeper understanding.\nThe book also serves as a bridge into the research literature. After working through the appropriate\nchapters, the student should be able to make sense of a large fraction of the articles published today\nin polymer science journals.\nThere is more than enough material in this book for a full-year graduate level course, but as with\nthe first edition, the level is (almost) always accessible to senior level undergraduates. After an\nintroductory chapter of broad scope, the bulk of the text may be grouped into three blocks of four\nchapters each. Chapter 2 through Chapter 5 describe the many ways in which polymers can be\nsynthesized and how the synthetic route in\ufb02uences the resulting molecular structure. This material\ncould serve as the basis for a single quarter or semester chemistry course that focuses on polymer\nsynthesis. Chapter 6 through Chapter 9 emphasize the solution properties of polymers, including\ntheir conformations, thermodynamics, hydrodynamics, and light scattering properties. Much of this\nmaterial is often found in a quarter or semester course introducing the physical chemistry of\npolymers. Chapter 10 through Chapter 13 address the solid state and bulk properties of polymers:\nrubber elasticity, viscoelasticity, the glass transition, and crystallization. These topics, while\npresented here from a physical chemical point of view, could equally well serve as the cornerstone\nof an introductory course in materials science or chemical engineering.\nThe style of the presentation, as with the previous edition, is chosen with the student in mind. To\nthis end, we may point out the following features:\n"]], ["block_2", ["\u00b0\nThere are over 60 worked example problems sprinkled throughout the book.\n\u00b0\nThere are 15 or more problems at the end of every chapter, to reinforce and develop further\nunderstanding; many of these are based on data from the literature.\n"]], ["block_3", ["Polymer science is today a vibrant field. Its technological relevance is vast, yet fundamental\nscientific questions also abound. Polymeric materials exhibit a wealth of fascinating properties,\nmany of which are observable just by manipulating a piece in your hands. Yet, these phenomena\nare all directly traceable to molecular behavior, and especially to the long chain nature of polymer\nmolecules. The central goal of this book is to develop a molecular level understanding of the\nproperties of polymers, beginning with the underlying chemical structures, and assuming no prior\nknowledge beyond undergraduate organic and physical chemistry. Although such an understand-\ning should be firmly based in chemistry, polymer science is a highly interdisciplinary endeavor;\nconcepts from physics, biology, materials science, chemical engineering, and statistics are all\n"]], ["block_4", ["Preface to the Second Edition\n"]], ["block_5", ["V\n"]]], "page_5": [["block_0", [{"image_0": "5_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Jingshan Dong, Will Edmonds, Sandra Fritz, Carolyn Gamble, Piotr Grzywacz, Jeong-Myeong Ha,\nBenjamin Hamilton, Amanda Haws, Nazish Hoda, Hao Hou, Deanna Huehn, Shengxiang Ji, Karan\nJindal, Young Kang, Aaron Khieu, Byeong-Su Kim, BongSoo Kim, Hyunwoo Kim, Jin-Hong\nKim, Seung Ha Kim, Chunze Lai, Castro Laicer, Qiang Lan, Sangwoo Lee, Zhibo Li, Elizabeth\nLugert, Nate Lynd, Sudeep Maheshwari, Huiming Mao, Adam Meuler, Yoichiro Mori, Randy\nMrozek, Siddharthya Mujumdar, Jaewook Nam, Dan O\u2019Neal, Sahban Ozair, Matt Panzer, Alhad\nPhatak, William Philip, Jian Qin, Benjamin Richter, Scott Roberts, Josh Scheffel, Jessica Schom-\nmer, Kathleen Schreck, Peter Simone, Zach Thompson, Kristianto Tjiptowidjojo, Mehul Vora,\nJaye Warner, Tomy Widya, Maybelle Wu, Jianyan Xu, Dan Yu, Ilan Zeroni, Jianbin Zhang, Ling\nZhang, Yu Zhang, Ning Zhou, Zhengxi Zhu, and John Zupancich. Last but not least the\nlove, support, and tolerance of my family, Susanna, Hannah, and Sam, has been a constant source\nof strength.\n"]], ["block_2", ["An undertaking such as writing a textbook can never be completed without important contributions\nfrom many individuals. Large sections of manuscript were carefully typed by Becky Matsch and\nLynne Johnsrud; Lynne also helped greatly with issues of copyright permissions and figure prepar-\nation. My colleagues past and present in the Polymer Group at Minnesota have been consistently\nencouraging and have provided both useful feedback and insightful examples: Frank Bates, Shura\nGrosberg, Marc Hillmyer, Chris Macosko, Wilmer Miller, David Morse, Steve Prager, and Matt\nTirrell. In large measure the style adopted in this second edition has been inspired by the example set\nby my graduate instructors and mentors at the University of Wisconsin: R. Byron Bird, John Perry,\nArthur Lodge, John Schrag, and Hyuk Yu. In particular, it was in his graduate course Chemistry 664\nthat Hyuk Yu so ably demonstrated that no important equation need come out of thin air.\nI would like extend a special thank you to all of the students enrolled in Chemistry/Chemical\nEngineering/Materials Science 8211 over the period 2002\u20142005, who worked through various\ndrafts of Chapter 6 through Chapter 13, and provided many helpful suggestions: Sayeed Abbas,\nDavid Ackerman, Sachin Agarwal, Saurabh Agarwal, Julie Alkatout, Pedro Arrechea, Carlos\nLopez-Barron, Soumendra Basu, Jeff Becker, Joel Bell, A.S. Bhalla, Michael Bluemle, Paul\nBoswell, Bryan Boudouris, Adam Buckalew, Xiuyu Cai, Neha Chandra, Joon Chatterjee, Liang\nChen, Ying Chen, Juhee Cho, Seongho Choi, Jin-Hwa Chung, Kevin Davis, Michail Dolgovskij,\n"]], ["block_3", ["\u00b0\nThere are almost 200 figures, to illustrate concepts or to present experimental results from\nthe literature.\n0\nStudies chosen for the examples, problems, and figures range in vintage from very recent to\nover 50 years old; this feature serves to give the reader some sense of the historical progression\nof the field.\n\u00b0\nConcise reviews of many topics (such as thermodynamics, kinetics, probability, and various\nexperimental techniques) are given when the subject is first raised.\n\u00b0\nA conscious effort has been made to cross-reference extensively between chapters and sections\nwithin chapters, in order to help tie the various tOpics together.\n\u00b0\nImportant equations and mathematical relations are almost always developed step by step. We\nhave avoided, wherever possible, the temptation to pull equations out of a hat. Occasionally\nthis leads to rather long stretches of algebra, which the reader is welcome to skip. However, at\nsome point the curious student will want to know where the result comes from, and then this\nbook should be a particularly valuable resource. Surprisingly, perhaps, the level of mathemat\u2014\nical sophistication is only about the same as needed in undergraduate chemical thermodynam-\nics. As a further help in this regard, an Appendix reviews many of the important mathematical\ntools and tricks.\n"]], ["block_4", ["vi\nPreface to the 2nd Edition\n"]], ["block_5", ["Tim Lodge\n"]]], "page_6": [["block_0", [{"image_0": "6_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Physical chemistry has been defined as that branch of science that is fundamental, molecular,\nand interesting. I have tried to write a polymer textbook that could be described this way also. To\nthe extent that one subscribes to the former definition and that I have succeeded in the latter\nobjective, then the approach of this book is physical chemical. As a textbook, it is intended for\nstudents who have completed courses in physical and organic chemistry. These are the prerequi\u2014\nsites that de\ufb01ne the level of the book; no special background in physics or mathematics beyond\nwhat is required for physical chemistry is assumed. Since chemistry majors generally study\nphysical chemistry in the third year of the undergraduate curriculum, this book can serve as the\ntext for a senior\u2014level undergraduate or a beginning graduate\u2014level course. Although I use\nchemistry courses and chemistry curricula to describe the level of this book, students majoring\nin engineering, materials science, physics, and various specialties in the biological sciences will\nalso find numerous topics of interest contained herein.\nTerms like \u201cfundamental,\u201d \u201cmolecular,\u201d and \u201cinteresting\u201d have different meanings for\ndifferent people. Let me explain how they apply to the presentation of polymer chemistry in this\ntext.\nThe words \u201cbasic concepts\u201d in the title define what I mean by \u201cfundamental.\u201d This is the\nprimary emphasis in this presentation. Practical applications of polymers are cited frequently\u2014\nafter all, it is these applications that make polymers such an important class of chemicals\u2014but in\noverall content, the stress is on fundamental principles. \u201cFoundational\u201d might be another way to\ndescribe this. I have not attempted to cover all aspects of polymer science, but the topics that have\nbeen discussed lay the foundation\u2014built on the bedrock of organic and physical chemistry\u2014from\nwhich virtually all aspects of the subject are developed. There is an enormous literature in polymer\nscience; this book is intended to bridge the gap between the typical undergraduate background in\npolymers\u2014which frequently amounts to little more than occasional \u201crelevant\u201d examples in other\ncourses\u2014and the professional literature on the subject. Accordingly, the book assumes essentially no\nprior knowledge of polymers, and extends far enough to provide a usable level of understanding.\n\u201cMolecular\u201d describes the perspective of the chemist, and it is this aspect of polymeric\nmaterials that I try to keep in view throughout the book. An engineering text might emphasize\nprocessing behavior; a physics text, continuum mechanics; a biochemistry text, physiological\nfunction. All of these are perfectly valid points of view, but they are not the approach of this\nbook. It is polymer molecules\u2014their structure, energetics, dynamics, and reactions\u2014that are the\nprimary emphasis throughout most of the book. Statistics is the type of mathematics that is natural\nto a discussion of molecules. Students are familiar with the statistical nature of, say, the kinetic\nmolecular theory of gases. Similar methods are applied to other assemblies of molecules, or in the\ncase of polymers, to the assembly of repeat units that comprise a single polymer molecule.\nAlthough we frequently use statistical arguments, these are developed quite thoroughly and do\nnot assume any more background in this subject than is ordinarily found among students in a\nphysical chemistry course.\n"]], ["block_2", ["Preface to the First Edition\n"]], ["block_3", ["vii\n"]]], "page_7": [["block_0", [{"image_0": "7_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The book is divided into three parts of three chapters each, after an introductory chapter which\ncontains information that is used throughout the book.\nIn principle, the three parts can be taken up in any order without too much interruption in\ncontinuity. Within each of the parts there is more carryover from chapter to chapter, so rearranging\nthe sequence of topics within a given part is less convenient. The book contains more material than\ncan be covered in an ordinary course. Chapter 1 plus two of the three parts contain about the right\namount of material for one term. In classroom testing the material, I allowed the class to decide\u2014\nwhile we worked on Chapter l\u2014which two of the other parts they wished to cover; this worked\nvery well.\nMaterial from Chapter 1 is cited throughout the book, particularly the discussion of statistics. In\nthis connection, it might be noted that statistical arguments are developed in less detail further\nalong in the book as written. This is one of the drawbacks of rearranging the order in which the\ntopics are covered. Chapters 2 through 4 are concerned with the mechanical properties of bulk\npolymers, properties which are primarily responsible for the great practical importance of poly-\nmers. Engineering students are likely to have both a larger interest and a greater familiarity with\nthese topics. Chapers 5 through 7 are concerned with the preparation and properties of several\nbroad classes of polymers. These topics are closer to the interests of chemistry majors. Chapters 8\nthrough 10 deal with the solution properties of polymers. Since many of the techniques described\nhave been applied to biopolymers, these chapters will have more appeal to students of biochem\u2014\nistry and molecular biology.\nLet me conclude by acknowledging the contributions of those who helped me with the\npreparation of this book. I wish to thank Marilyn Steinle for expertly typing the manuscript. My\nappreciation also goes to Carol Truett who skillfully transformed my (very) rough sketches into\neffective illustrations. Lastly, my thanks to Ron Manwill for preparing the index and helping me\nwith the proofreading. Finally, let me acknowledge that some errors and/or obscurities will surely\nelude my efforts to eliminate them. I would appreciate reports about these from readers so that\nthese mistakes can eventually be eliminated.\n"]], ["block_2", ["1.\nOver 50 solved example problems are sprinkled throughout the book.\n2.\nExercises are included at the end of each chapter which are based on data from the original\nliterature.\n3.\nConcise reviews of pertinent aspects of thermodynamics, kinetics, spectrophotometry, etc. are\npresented prior to developing applications of these topics to polymers.\n4.\nTheoretical models and mathematical derivations are developed in enough detail to be\ncomprehensible to the student reader. Only rarely do I \u201cpull results out of a hat,\u201d and I\nscrupulously avoid saying \u201cit is obvious that .. .\u201d\n5.\nGenerous cross\u2014referencing and a judicious amount of repetition have been included to help\nunify a book which spans quite a wide range of topics.\n"]], ["block_3", ["The most subjective of the words which (I hope) describe this book is \n\u2018interesting.\u201d The\nfascinating behavior of polymers themselves, the clever experiments of laboratory researchers, and\nthe elegant work of the theoreticians add up to an interesting total. I have tried to tell about these\ntopics with clarity and enthusiasm, and in such a way as to make them intelligible to students. I can\nonly hope that the reader agrees with my assessment of what is interesting.\nThis book was written with the student in mind. Even though \u201cstudent\u201d encompasses persons\nwith a wide range of backgrounds, interests, and objectives; these are different than the corre\u2014\nsponding experiences and needs of researchers. The following features have been to assist\nthe student:\n"]], ["block_4", ["viii\nPreface to the lst Edition\n"]], ["block_5", ["6.\nSI units have been used fairly consistently throughout, and attention is paid to the matter of\nunits whenever these become more than routine in complexity.\n"]], ["block_6", ["Paul C. Hiemenz\n"]]], "page_8": [["block_0", [{"image_0": "8_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\n"]], ["block_2", ["1.7.1\nNumber-, Weight\u2014, and z\u2014Average Molecular Weights .................................. 25\n1.7.2\nPolydispersity Index and Standard Deviation .................................................26\n1.7.3\nExamples of Distributions ...............................................................................28\n1.8\nMeasurement of Molecular Weight............................................................................. 31\n1.8.1\nGeneral Considerations ....................................................................................31\n1.8.2\nEnd Group Analysis .........................................................................................32\n1.8.3\nMALDI Mass Spectrometry ............................................................................ 35\n1.9\nPreview of Things to Come......................................................................................... 37\n1.10\nChapter Summary ........................................................................................................ 38\nProblems .................................................................................................................................. 38\nReferences ................................................................................................................................ 41\nFurther Readings ...................................................................................................................... 41\n"]], ["block_3", ["Step-Growth Polymerization ...................................................................... 43\n"]], ["block_4", ["2.1\nIntroduction .................................................................................................................... 43\n2.2\nCondensation Polymers: One Step at a Time ............................................................... 43\n2.2.1\nClasses of Step\u2014Growth Polymers ..................................................................... 43\n2.2.2\nFirst Look at the Distribution of Products ........................................................44\n2.2.3\nA First Look at Reactivity and Reaction Rates ................................................46\n"]], ["block_5", ["Introduction to Chain Molecules ................................................................. 1\n"]], ["block_6", ["1.4\nAddition, Condensation, and Natural Polymers .......................................................... 11\n1.4.1\nAddition and Condensation Polymers ............................................................. 11\n1.4.2\nNatural Polymers ............................................................................................. 13\n1.5\nPolymer Nomenclature ................................................................................................ 18\n1.6\nStructural Isomerism .................................................................................................... 20\n1.6.1\nPositional Isomerism........................................................................................20\n1.6.2\nStereo Isomerism ............................................................................................. 21\n1.6.3\nGeometrical Isomerism .................................................................................... 22\n1.7\nMolecular Weights and Molecular Weight Averages ................................................. 24\n"]], ["block_7", ["1. 1\nIntroduction .................................................................................................................... 1\n1.2\nHow Big Is Big? ............................................................................................................ 3\n1.2.1\nMolecular Weight .............................................................................................. 3\n1.2.2\nSpatial Extent ..................................................................................................... 5\n1.3\nLinear and Branched Polymers, Homopolymers, and Copolymers ............................. 7\n1.3.1\nBranched Structures ...........................................................................................7\n"]], ["block_8", ["1 .3.2\nCopolymers ........................................................................................................9\n"]], ["block_9", ["Contents\n"]]], "page_9": [["block_0", [{"image_0": "9_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.1\nIntroduction .................................................................................................................... 77\n3.2\nChain-Growth and Step-Growth Polymerizations: Some Comparisons ....................... 77\n3.3\nInitiation ......................................................................................................................... 79\n3.3.1\nInitiation Reactions ............................................................................................ 80\n3.3.2\nFate of Free Radicals ......................................................................................... 81\n3.3.3\nKinetics of Initiation .......................................................................................... 82\n3.3.4\nPhotochemical Initiation....................... ............................................................. 84\n3.3.5\nTemperature Dependence of Initiation Rates.................................................... 85\n3.4\nTermination .................................................................................................................... 86\n3.4.1\nCombination and Disproportionation ................................................................ 86\n3.4.2\nEffect of Termination on Conversion to Polymer ............................................ 88\n3.4.3\nStationary-State Radical Concentration ............................................................ 89\n3.5\nPropagation .................................................................................................................... 90\n3.5.1\nRate Laws for Propagation ................................................................................91\n3.5.2\nTemperature Dependence of Propagation Rates ...............................................92\n3.5.3\nKinetic Chain Length.........................................................................................94\n3.6\nRadical Lifetime ............................................................................................................ 96\n3.7\nDistribution of Molecular Weights................................................................................ 99\n3.7.1\nDistribution of i\u2014mers: Termination by Disproportionation .............................99\n3.7.2\nDistribution of i-mers: Termination by Combination ..................................... 102\n3.8\nChain Transfer ............................................................................................................. 104\n3.8.1\nChain Transfer Reactions ................................................................................ 105\n3.8.2\nEvaluation of Chain Transfer Constants ......................................................... 106\n3.8.3\nChain Transfer to Polymer .............................................................................. 108\n3.8.4\nSuppressing Polymerization ............................................................................ 109\n3.9\nChapter Summary ........................................................................................................ 110\nProblems ................................................................................................................................ 110\nReferences.............................................................................................................................. 1 14\n.\nFurther Readings .................................................................................................................... I 15\n"]], ["block_2", ["2.3\nKinetics of Step\u2014Growth Polymerization ...................................................................... 49\n2.3.1\nCatalyzed Step\u2014Growth Reactions ..................................................................... 50\n2.3.2\nHow Should Experimental Data Be Compared with\nTheoretical Rate Laws? ..................................................................................... 52\n2.3.3\nUncatalyzed Step-Growth Reactions .................................................................53\n2.4\nDistribution of Molecular Sizes .................................................................................... 55\n2.4.1\nMole Fractions of Species ................................................................................. 56\n2.4.2\nWeight Fractions of Species .............................................................................. 58\n2.5\nPolyesters ....................................................................................................................... 60\n2.6\nPolyamides ..................................................................................................................... 64\n2.7\nStoichiometric Imbalance .............................................................................................. 67\n2.8\nChapter Summary .......................................................................................................... 71\nProblems .................................................................................................................................. 71\nReferences................................................................................................................................ 76\nFurther Readings ............................ 76\n"]], ["block_3", ["Chain-Growth Polymerization....................................................................77\n"]], ["block_4", ["Contents\n"]]], "page_10": [["block_0", [{"image_0": "10_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4\nControlled Polymerization ....................................................................... 117\n"]], ["block_2", ["Contents\nxi\n"]], ["block_3", ["5\nCopolymers, Microstructure, and Stereoregularity................................ 165\n"]], ["block_4", ["4.1\nIntroduction ................................................................................................................ 1 17\n4.2\nPoisson Distribution for an Ideal Living Polymerization ......................................... 118\n4.2.1\nKinetic Scheme .............................................................................................. 119\n4.2.2\nBreadth of the Poisson Distribution .............................................................. 122\n4.3\nAnionic Polymerization ............................................................................................. 126\n4.4\nBlock Copolymers, End\u2014Functional Polymers, and Branched\nPolymers by Anionic Polymerization........................................................................ 129\n4.4.1\nBlock Copolymers ......................................................................................... 129\n4.4.2\nEnd-Functional Polymers .............................................................................. 133\n4.4.3\nRegular Branched Architectures .................................................................... 135\n4.5\nCationic Polymerization ............................................................................................ 137\n4.5.1\nAspects of Cationic Polymerization .............................................................. 138\n4.5.2\nLiving Cationic Polymerization .................................................................... 140\n4.6\nControlled Radical Polymerization ........................................................................... 142\n4.6.1\nGeneral Principles of Controlled Radical Polymerization............................ 142\n4.6.2\nParticular Realizations of Controlled Radical Polymerization ..................... 144\n4.6.2.1\nAtom Transfer Radical Polymerization (ATRP)........................... 144\n4.6.2.2\nStable Free-Radical Polymerization (SFRP) ................................. 145\n4.6.2.3\nReversible Addition-Fragmentation Transfer (RAFT)\nPolymerization ............................................................................... 146\n4.7\nPolymerization Equilibrium....................................................................................... 147\n4.8\nRing-Opening Polymerization (ROP) ....................................................................... 150\n4.8.1\nGeneral Aspects ............................................................................................. 150\n4.8.2\nSpecific Examples of Living Ring\u2014Opening Polymerizations ...................... 152\n4.8.2.1\nPoly(ethy1ene oxide) ...................................................................... 152\n4.8.2.2\nPolylactide...................................................................................... 153\n4.8.2.3\nPoly(dimethylsiloxane) .................................................................. 154\n4.8.2.4\nRing-Opening Metathesis Polymerization (ROMP) ..................... 155\n4.9\nDendrimers ................................................................................................................. 156\n4.10\nChapter Summary ...................................................................................................... 160\nProblems ................................................................................................................................ 161\nReferences .............................................................................................................................. 163\nFurther Readings.................................................................................................................... 163\n"]], ["block_5", ["5.1\nIntroduction ................................................................................................................ 165\n5.2\nCopolymer Composition ............................................................................................ 166\n5.2.1\nRate Laws ...................................................................................................... 166\n5.2.2\nComposition versus Feedstock ...................................................................... 168\n5.3\nReactivity Ratios ........................................................................................................ 170\n5.3.1\nEffects of r Values ......................................................................................... 171\n5.3.2\nRelation of Reactivity Ratios to Chemical Structure.................................... 173\n5.4\nResonance and Reactivity.......................................................................................... 175\n5.5\nA Closer Look at Microstructure .............................................................................. 179\n5.5.1\nSequence Distributions .................................................................................. 180\n5.5.2\nTerminal and Penultimate Models ................................................................ 183\n"]]], "page_11": [["block_0", [{"image_0": "11_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["xii\nContents\n"]], ["block_2", ["5.6\nCOpolymer Composition and Microstructure: Experimental Aspects ...................... 185\n5.6.1\nEvaluating Reactivity Ratios from Composition Data ................................. 185\n5.6.2\nSpectroscopic Techniques.............................................................................. 188\n5.6.3\nSequence Distribution: Experimental Determination ................................... 190\n5.7\nCharacterizing Stereoregularity ................................................................................. 193\n5.8\nA Statistical Description of Stereoregularity ............................................................ 196\n5.9\nAssessing Stereoregularity by Nuclear Magnetic Resonance ................................... 200\n5.10\nZiegler\u2014Natta Catalysts ............................................................................................. 205\n5.11\nSingle-Site Catalysts .................................................................................................. 208\n5.12\nChapter Summary ...................................................................................................... 211\nProblems ................................................................................................................................ 212\nReferences .............................................................................................................................. 216\nFurther Readings .................................................................................................................... 216\n"]], ["block_3", ["6.1\nConformations, Bond Rotation, and Polymer Size ..................................................... 217\n6.2\nAverage End\u2014to-End Distance for Model Chains ....................................................... 219\nCase 6.2.1\nThe Freely Jointed Chain ....................................................................... 220\nCase 6.2.2\nThe Freely Rotating Chain ..................................................................... 221\nCase 6.2.3\nHindered Rotation Chain ........................................................................ 222\n6.3\nCharacteristic Ratio and Statistical Segment Length.................................................. 223\n6.4\nSemi\ufb02exible Chains and the Persistence Length ........................................................ 225\n6.4.1\nPersistence Length of Flexible Chains ............................................................ 227\n6.4.2\nWorm-Like Chains ..........................................................................................228\n6.5\nRadius of Gyration....................................................................................................... 230\n6.6\nSpheres, Rods, and Coils ............................................................................................. 234\n6.7\nDistributions for End\u2014to-End Distance and Segment Density .................................... 235\n6.7.1\nDistribution of the End-to\u2014End Vector ............................................................ 236\n6.7.2\nDistribution of the End-to\u2014End Distance.........................................................239\n6.7.3\nDistribution about the Center of Mass ............................................................ 240\n6.8\nSelf-Avoiding Chains: A First Look ........................................................................... 241\n6.9\nChapter Summary ........................................................................................................ 242\nProblems ................................................................................................................................ 242\nReferences .............................................................................................................................. 244\nFurther Readings .................................................................................................................... 245\n"]], ["block_4", ["7.1\nReview of Thermodynamic and Statistical Thermodynamic Concepts ..................... 247\n7.2\nRegular Solution Theory ............................................................................................. 249\n7.2.1\nRegular Solution Theory: Entropy of Mixing .................................................249\n7.2.2\nRegular Solution Theory: Enthalpy of Mixing ............................................... 251\n7.3\nFlory\u2014Huggins Theory ................................................................................................. 254\n7.3.1\nFlory\u2014Huggins Theory: Entropy of Mixing by a Quick Route ......................255\n7.3.2\nFlory\u2014Huggins Theory: Entropy of Mixing by a Longer Route .................... 255\n7.3.3\nFlory\u2014Huggins Theory: Enthalpy of Mixing .................................................. 257\n7.3.4\nFlory\u2014Huggins Theory: Summary of Assumptions ........................................ 258\n"]], ["block_5", ["Thermodynamics of Polymer Solutions ................................................. 247\n"]], ["block_6", ["Polymer Conformations ........................................................................... 217\n"]]], "page_12": [["block_0", [{"image_0": "12_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["8\nLight Scattering by Polymer Solutions ................................................... 289\n"]], ["block_2", ["9\nDynamics of Dilute Polymer Solutions ................................................... 327\n"]], ["block_3", ["Contents\nxiii\n"]], ["block_4", ["8.1\nIntroduction: Light Waves ........................................................................................... 289\n8.2\nBasic Concepts of Scattering ....................................................................................... 291\n8.2.1\nScattering from Randomly Placed Objects .....................................................292\n8.2.2\nScattering from a Perfect Crystal ....................................................................292\n8.2.3\nOrigins of Incoherent and Coherent Scattering ..............................................293\n8.2.4\nBragg\u2019s Law and the Scattering Vector ..........................................................294\n8.3\nScattering by an Isolated Small Molecule .................................................................. 296\n8.4\nScattering from a Dilute Polymer Solution................................................................. 298\n8.5\nThe Form Factor and the Zimm Equation .................................................................. 304\n8.5.1\nMathematical Expression for the Form Factor ...............................................305\n8.5.2\nForm Factor for Isotropic Solutions ................................................................ 306\n8.5.3\nForm Factor as n\u2014>0 .................................................................................... 307\n8.5.4\nZimm Equation ................................................................................................ 307\n8.5.5\nZimm Plot ........................................................................................................ 308\n8.6\nScattering Regimes and Particular Form Factors........................................................ 312\n8.7\nExperimental Aspects of Light Scattering .................................................................. 314\n8.7. 1\nInstrumentation ................................................................................................ 3 16\n8.7.2\nCalibration ........................................................................................................317\n8.7.3\nSamples and Solutions ..................................................................................... 319\n8.7.4\nRefractive Index Increment ............................................................................. 319\n8.8\nChapter Summary ........................................................................................................ 320\nProblems ................................................................................................................................ 321\nReferences .............................................................................................................................. 325\nFurther Readings .................................................................................................................... 325\n"]], ["block_5", ["9.1\nIntroduction: Friction and Viscosity............................................................................ 327\n9.2\nStokes\u2019 Law and Einstein\u2019s Law ................................................................................. 330\n"]], ["block_6", ["7.4\nOsmotic Pressure ......................................................................................................... 258\n7.4.1\nOsmotic Pressure: General Case .....................................................................259\n7.4.1.1\nNumber\u2014Average Molecular Weight ............................................... 261\n7.4.2\nOsmotic Pressure: Flory\u2014Huggins Theory ......................................................263\n7.5\nPhase Behavior of Polymer Solutions ......................................................................... 264\n7.5.1\nOverview of the Phase Diagram...................................................................... 265\n7.5.2\nFinding the Binodal .........................................................................................268\n7.5.3\nFinding the Spinodal ........................................................................................269\n7.5.4\nFinding the Critical Point ................................................................................270\n7.5.5\nPhase Diagram from Flory\u2014Huggins Theory ..................................................271\n7.6\nWhat\u2019s in x? ................................................................................................................. 275\n7.6.1\nX from Regular Solution Theory .....................................................................275\n7.6.2\nX from Experiment...........................................................................................276\n7.6.3\nFurther Approaches to X ..................................................................................278\n7.7\nExcluded Volume and Chains in a Good Solvent ...................................................... 280\n7.8\nChapter Summary ........................................................................................................ 283\nProblems ................................................................................................................................ 284\nReferences .............................................................................................................................. 287\nFurther Readings.................................................................................................................... 288\n"]]], "page_13": [["block_0", [{"image_0": "13_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["xiv\n"]], ["block_2", ["10\n"]], ["block_3", ["9.2.1\nViscous Forces on Rigid Spheres .............................................................331\n9.2.2\nSuspension of Spheres............................................................................... 332\n9.3\nIntrinsic Viscosity ................................................................................................... 334\n9.3.1\nGeneral Considerations ............................................................................. 334\n9.3.2\nMark\u2014Houwink Equation .......................................................................... 336\n9.4\nMeasurement of Viscosity ...................................................................................... 341\n9.4.1\nPoiseuille Equation and Capillary Viscometers .......................................341\n9.4.2\nConcentric Cylinder Viscometers ............................................................. 345\n9.5\nDiffusion Coefficient and Friction Factor .............................................................. 346\n9.5.1\nTracer Diffusion and Hydrodynamic Radius............................................ 347\n9.5.2\nMutual Diffusion and Fick\u2019s Laws ........................................................... 348\n9.6\nDynamic Light Scattering ....................................................................................... 354\n9.7\nHydrodynamic Interactions and Draining .............................................................. 357\n9.8\nSize Exclusion Chromatography (SEC) ................................................................. 360\n9.8.1\nBasic Separation Process .......................................................................... 361\n9.8.2\nSeparation Mechanism .............................................................................. 365\n9.8.3\nTwo Calibration Strategies ....................................................................... 367\n9.8.3.1\nLimitations of Calibration by Standards................................... 367\n9.8.3.2\nUniversal Calibration ................................................................ 368\n9.8.4\nSize Exclusion Chromatography Detectors .............................................. 369\n9.8.4.1\nRI Detector ................................................................................ 369\n9.8.4.2\nUV\u2014Vis Detector ....................................................................... 370\n9.8.4.3\nLight Scattering Detector.......................................................... 371\n9.8.4.4\nViscometer................................................................................. 372\n9.9\nChapter Summary ...................................................................................................... 372\nProblems .............................................................................................................................. 373\nReferences ........................................................................................................................... 378\nFurther Readings ................................................................................................................. 379\n"]], ["block_4", ["Networks, Gels, and Rubber Elasticity\n3\n"]], ["block_5", ["10.6\n"]], ["block_6", ["10.1\n"]], ["block_7", ["10.2\n"]], ["block_8", ["10.3\n10.4\n"]], ["block_9", ["10.5\n"]], ["block_10", ["Formation of Networks by Random Cross-Linking ............................................... 381\n10.1.1\nDefinitions ................................................................................................. 381\n10.1.2\nGel Point.................................................................................................... 383\nPolymerization with Multifunctional Monomers ................................................... 386\n10.2.1\nCalculation of the Branching Coefficient.................................................387\n10.2.2\nGel Point....................................................................................................388\n10.2.3\nMolecular-Weight Averages ..................................................................... 389\nElastic Deformation ................................................................................................ 392\nThermodynamics of Elasticity ................................................................................ 394\n10.4.1\nEquation of State .......................................................................................394\n10.4.2\nIdeal Elastomers ........................................................................................396\n10.4.3\nSome Experiments on Real Rubbers ........................................................397\nStatistical Mechanical Theory of Rubber Elasticity: Ideal Case ........................... 398\n10.5.1\nForce to Extend a Gaussian Chain ...........................................................400\n10.5.2\nNetwork of Gaussian Strands....................................................................402\n10.5.3\nModulus of the Gaussian Network ...........................................................403\nFurther Developments in Rubber Elasticity ........................................................... 406\n10.6.1\nNon-Gaussian Force Law..........................................................................406\n10.6.2\nFront Factor ...............................................................................................407\n"]], ["block_11", ["Contents\n"]]], "page_14": [["block_0", [{"image_0": "14_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Contents\nxv\n"]], ["block_2", ["11\n"]], ["block_3", ["12\n"]], ["block_4", ["1 1.7.1\nReptation Model: Longest Relaxation Time and Diffusivity...................451\n11.7.2\nReptation Model: Viscoelastic Properties ................................................453\n11.7.3\nReptation Model: Additional Relaxation Processes ................................. 456\n11.8\nAspects of Experimental Rheometry ...................................................................... 458\n11.8.1\nShear Sandwich and Cone and Plate Rheometers ....................................45 8\n11.8.2\nFurther Comments about Rheometry ........................................................459\n11.9\nChapter Summary .................................................................................................... 460\nProblems .............................................................................................................................. 461\nReferences ........................................................................................................................... 464\nFurther Readings ................................................................................................................. 464\n"]], ["block_5", ["Glass Transition ...................................................................................... 465\n"]], ["block_6", ["10.6.3\nNetwork Defects ........................................................................................408\n10.6.4\nMooney\u2014Rivlin Equation ..........................................................................409\n10.7\nSwelling of Gels ...................................................................................................... 410\n10.7.1\nModulus of a Swollen Rubber ..................................................................411\n10.7.2\nSwelling Equilibrium ................................................................................412\n10.8\nChapter Summary .................................................................................................... 414\nProblems .............................................................................................................................. 416\nReferences ........................................................................................................................... 4 18\nFurther Readings ................................................................................................................. 418\n"]], ["block_7", ["Linear Viscoelasticity ............................................................................. 419\n"]], ["block_8", ["11.1\nBasic Concepts ........................................................................................................ 419\n11.1.1\nStress and Strain ........................................................................................421\n11.1.2\nViscosity, Modulus, and Compliance .......................................................421\n11.1.3\nViscous and Elastic Responses .................................................................422\n11.2\nResponse of the Maxwell and Voigt Elements ...................................................... 423\n11.2.1\nTransient Response: Stress Relaxation .....................................................423\n11.2.2\nTransient Response: Creep........................................................................425\n11.2.3\nDynamic Response: Loss and Storage Moduli.........................................426\n11.2.4\nDynamic Response: Complex Modulus and Complex Viscosity ............429\n11.3\nBoltzmann Superposition Principle ........................................................................ 430\n11.4\nBead\u2014Spring Model................................................................................................. 432\n11.4.1\nIngredients of the Bead\u2014Spring Model .....................................................432\n11.4.2\nPredictions of the Bead\u2014Spring Model .....................................................434\n11.5\nZimm Model for Dilute Solutions, Rouse Model\nfor Unentangled Melts ............................................................................................ 439\n11.6\nPhenomenology of Entanglement ........................................................................... 444\n11.6.1\nRubbery Plateau ........................................................................................444\n11.6.2\nDependence of Me on Molecular Structure ..............................................447\n11.7\nReptation Model ...................................................................................................... 450\n"]], ["block_9", ["12.1\nIntroduction ............................................................................................................. 465\n12.1.1\nDefinition of a Glass ................................................................................. 465\n12.1.2\nGlass and Melting Transitions ..................................................................466\n12.2\nThermodynamic Aspects of the Glass Transition .................................................. 468\n"]], ["block_10", ["12.2.1\nFirst-Order and Second\u2014Order Phase Transitions .....................................469\n"]]], "page_15": [["block_0", [{"image_0": "15_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["xvi\n"]], ["block_2", ["13\n"]], ["block_3", ["Crystalline Polymers............................................................................... 511\n"]], ["block_4", ["12.2.2\nKauzmann Temperature ............................................................................471\n12.2.3\nTheory of Gibbs and DiMarzio ................................................................472\n12.3\nLocating the Glass Transition Temperature ........................................................... 474\n12.3. 1\nDilatometry ................................................................................................474\n12.3.2\nCalorimetry ................................................................................................ 476\n12.3.3\nDynamic Mechanical Analysis .................................................................478\n12.4\nFree Volume Description of the Glass Transition.................................................. 479\n12.4.1\nTemperature Dependence of the Free Volume ........................................480\n12.4.2\nFree Volume Changes Inferred from the Viscosity .................................481\n12.4.3\nWilliams\u2014Landel\u2014Ferry Equation .............................................................483\n12.5\nTime\u2014Temperature Superposition........................................................................... 486\n12.6\nFactors that Affect the Glass Transition Temperature ........................................... 491\n12.6.1\nDependence on Chemical Structure..........................................................491\n12.6.2\nDependence on Molecular Weight ...........................................................492\n12.6.3\nDependence on Composition ....................................................................492\n12.7\nMechanical Properties of Glassy Polymers ............................................................ 496\n12.7.1\nBasic Concepts ..........................................................................................496\n12.7.2\nCrazing, Yielding, and the Brittle-to-Ductile Transition ........................ .498\n12.7.3\nRole of Chain Stiffness and Entanglements .............................................501\n12.8\nChapter Summary.................................................................................................... 504\nProblems .............................................................................................................................. 505\nReferences ........................................................................................................................... 508\nFurther Readings ................................................................................................................. 508\n"]], ["block_5", ["13.1\nIntroduction and Overview ..................................................................................... 511\n13.2\nStructure and Characterization of Unit Cells ......................................................... 513\n13.2.1\nClasses of Crystals ....................................................................................513\n13.2.2\nX-Ray Diffraction .....................................................................................515\n13.2.3\nExamples of Unit Cells .............................................................................518\n13.3\nThermodynamics of Crystallization: Relation of Melting Temperature\nto Molecular Structure ............................................................................................ 521\n13.4\nStructure and Melting of Lamellae ......................................................................... 526\n13.4.1\nSurface Contributions to Phase Transitions..............................................526\n13.4.2\nDependence of Tm on Lamellar Thickness ...............................................527\n13.4.3\nDependence of Tm on Molecular Weight .................................................530\n13.4.4\nExperimental Characterization of Lamellar Structure ............................. 532\n13.5\nKinetics of Nucleation and Growth ........................................................................ 536\n13.5.1\nPrimary Nucleation ...................................................................................537\n13.5.2\nCrystal Growth ..........................................................................................539\n13.6\nMorphology of Semicrystalline Polymers .............................................................. 545\n13.6.1\nSpherulites .................................................................................................545\n13.6.2\nNonspherulitic Morphologies ....................................................................548\n13.7\nKinetics of Bulk Crystallization ............................................................................. 551\n13.7.1\nAvrami Equation .......................................................................................552\n"]], ["block_6", ["Contents\n"]]], "page_16": [["block_0", [{"image_0": "16_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Appendix ......................................................................................................... 567\n"]], ["block_2", ["A.l\nSeries Expansions ...............................................................................................................567\nA2\nSummation Formulae ..........................................................................................................568\nA3\nTransformation to Spherical Coordinates ...........................................................................569\nA.4\nSome Integrals of Gaussian Functions ...............................................................................570\nA5\nComplex Numbers ..............................................................................................................572\n"]], ["block_3", ["Contents\nxvii\n"]], ["block_4", ["Index ................................................................................................................ 575\n"]], ["block_5", ["13.7.2\nKinetics of Crystallization: EXperimental Aspects ..................................556\n13.8\nChapter Summary .................................................................................................... 561\nProblems .............................................................................................................................. 562\nReferences ........................................................................................................................... 565\nFurther Readings ................................................................................................................. 565\n"]]], "page_17": [["block_0", [{"image_0": "17_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["\u201cI am inclined to think that the development of polymerization is perhaps the biggest thing\nchemistry has done, where it has had the biggest impact on everyday life\u201d [1]. This assessment\nof the significance of polymer chemistry to modern society was offered 25 years ago by Lord Todd\n(President of the Royal Society and 1957 Nobel Laureate in Chemistry), and subsequent deve10p-\nments have only reinforced this sentiment. There is hardly an area of modern life in which polymer\nmaterials do not play an important role. Applications span the range from the mundane\n(e.g., packaging, toys, fabrics, diapers, nonstick cookware, pressure-sensitive adhesives, etc.) to\ndemanding specialty uses (e.g., bulletproof vests, stealth aircraft, artificial hip joints, resorbable\nsutures, etc.). In many instances polymers are the main ingredients, and the ingredients whose\ncharacteristic properties are essential to the success of a particular technology: rubber tires, foam\ncushions and insulation, high\u2014performance athletic shoes, clothing, and equipment are good\nexamples. In other cases, polymers are used as additives at the level of a few percent by volume,\nbut which nevertheless play a crucial role in the properties of the \ufb01nal material; illustrations of this\ncan be found in asphalt (to suppress brittle fracture at low temperature and flow at high tempera-\nture), shampoo and other cosmetics (to impart \u201cbody\u201d), automobile Windshields (to prevent\nshattering), and motor oil (to reduce the dependence of viscosity on temperature, and to suppress\ncrystallization).\nFor those polymer scientists \u201cof a certain age,\u201d the 1967 movie \u201cThe Graduate\u201d [2] provided\nan indelible moment that still resonates today. At his college graduation party, the hero Benjamin\nBraddock (played by Dustin Hoffman) is offered the following advice by Mr. McGuire (played by\nWalter Brooke):\n"]], ["block_2", ["In that period, the term \u201cplastic\u201d was often accompanied by negative connotations, including\n\u201cartificial,\u201d as opposed to \u201cnatural,\u201d and \u201ccheap,\u201d as opposed to \u201cvaluable.\u201d Today, in what we\nmight call the \u201cpost-graduate era,\u201d the situation has changed. To the extent that the advice offered\nto Benjamin was pointing him to a career in a particular segment of the chemical industry, it was\nprobably very sound advice. The volume of polymer materials produced annually has grown\nrapidly over the intervening years, to the point where today several hundred pounds of polymer\nmaterials are produced each year for each person in the United States. More interesting than sheer\nvolume, however, is the breadth of applications for polymers. Not only do they continue to\nencroach into the domains of \u201cclassical\u201d materials such as metal, wood, and glass (note the\ninexorable transformation of polymers from minor to major components in automobiles), but they\nalso play a central role in many emerging technologies. Examples include \u201cplastic electronics,\u201d\n"]], ["block_3", ["1.1\nIntroduction\n"]], ["block_4", ["BENJAMIN. Yes I am.\nMR. MCGUIRE. Plastics.\n"]], ["block_5", ["MR. MCGUIRE. I want to say one word to you. Just one word.\n"]], ["block_6", ["BENJAMIN. Yes, sir.\nMR. MCGUIRE. Are you listening?\n"]], ["block_7", ["Introduction to Chain Molecules\n"]], ["block_8", ["1\n"]]], "page_18": [["block_0", [{"image_0": "18_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a decade of controversy before this \u201cmacromolecular hypothesis\u201d began to experience widespread\nacceptance. Staudinger was awarded the Nobel Prize in 1953 for his work with polymers. By the\n1930s, Carothers\nbegan synthesizing polymers using\nwell-established reactions of organic\nchemistry such as esterification and amidation. His products were not limited to single ester or\namide linkages, however, but contained many such groups: they were polyesters and polyamides.\nPhysical chemists also got in on the act. Kuhn, Guth, Mark, and others were soon applying\nstatistics and crystallography to describe the multitude of forms a long-chain molecule could\nassume [3].\nOur purpose in this introduction is not to trace the history of polymer chemistry beyond the\nsketchy version above; interesting and extensive treatments are available [4,5]. Rather, the primary\nobjective is to introduce the concept of chain molecules, which stands as the cornerstone of all\npolymer chemistry. In the next few sections we shall explore some of the categories of polymers,\nsome of the reactions that produce them, and some aspects of isomerism which multiply the\nstructural possibilities. A common feature of all synthetic polymerization reactions is the statistical\nnature of the individual polymerization steps. This leads inevitably to a distribution of molecular\nweights, which we would like to describe. As a consequence of these considerations, another\nimportant part of this chapter is an introduction to some of the statistical concepts that also play a\ncentral role in polymer chemistry.\n"]], ["block_2", ["can cure such and such a dreaded disease, but they cannot do anything about the common cold\u201d or\n\u201cwe know more about the surface of the moon than the bottom of the sea.\u201d If such comparisons\nwere popular in the 19203, the saying might have been, \u201cwe know more about the structure of the\natom than about those messy, sticky substances called polymers.\u201d Indeed, Millikan\u2019s determin-\nation of the charge of an electron, Rutherford\u2019s idea of the nuclear atom, and Bohr\u2019s model of the\nhydrogen atom were all well-known concepts before the notion of truly covalent macromolecules\nwas accepted. This was the case in spite of the great importance of polymers to human life and\nactivities. Our bodies, like all forms of life, depend on polymer molecules: carbohydrates, proteins,\nnucleic acids, and so on. From the earliest times, polymeric materials have been employed to\nsatisfy human needs: wood and paper; hides; natural resins and gums; fibers such as cotton, wool,\nand silk.\nAttempts to characterize polymeric substances had been made, of course, and high molecular\nweights were indicated, even if they were not too accurate. Early workers tended to be more\nsuspicious of the interpretation of the colligative properties of polymeric solutions than to accept\nthe possibility of high molecular weight compounds. Faraday had already arrived at C5H3 as the\nempirical formula of \u201crubber\u201d in 1826, and isoprene was identi\ufb01ed as the product resulting from\nthe destructive distillation of rubber in 1860. The idea that a natural polymer such as rubber\nsomehow \u201ccontained\u201d isoprene emerged, but the nature of its involvement was more elusive.\nDuring the early years of the 20th century, organic chemists were enjoying success in deter-\nmining the structures of ordinary-sized organic molecules, and this probably contributed to their\nreluctance to look beyond structures of convenient size. Physical chemists were interested\nin intermolecular forces during this period, and the idea that polymers were the result of some\nsort of association between low molecular weight constituent molecules prevailed for a long while.\nStaudinger is generally credited as being the father of modern polymer chemistry, although a\nforeshadowing of his ideas can be traced through older literature. In 1920, Staudinger proposed\nthe chain formulas we accept today, maintaining that structures are held together by covalent\nbonds, which are equivalent in every way to those in low molecular weight compounds. There was\n"]], ["block_3", ["gene therapy, artificial prostheses, optical data storage, electric cars, and fuel cells. In short, a\nreasonable appreciation of the properties of chain molecules, and how these result in the many\ndesirable attributes of polymer-containing materials, is a necessity for a well-trained chemist,\nmaterials scientist, or chemical engineer today.\nScience tends to be plagued by cliches, which make invidious comparison of its efforts; \u201cthey\n"]], ["block_4", ["2\nIntroduction to Chain Molecules\n"]]], "page_19": [["block_0", [{"image_0": "19_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The term polymer is derived from the Greek words poly and meros, meaning many parts. We noted\nin the Section 1.1 that the existence of these parts was acknowledged before the nature of the\ninteraction which held them together was known. Today we realize that ordinary covalent bonds\n"]], ["block_2", ["One of the first things we must consider is what we mean when we talk about the \u201csize\u201d of a\npolymer molecule. There are two possibilities: one has to do with the number of repeat units and\nthe other to the spatial extent. In the former case, the standard term is molecular weight (although\nagain the reader must be aware that molar mass is often preferred). A closely related concept, the\ndegree of polymerization is also commonly used in this context. A variety of experimental\ntechniques are available for determining the molecular weight of a polymer. We shall discuss a\nfew such methods in Section 1.8 and postpone others until the appropriate chapters. The expression\nmolecular weight and molar mass should always be modified by the word average. This too is\nsomething we shall take up presently. For now, we assume that a polymer molecule has a\nmolecular weight M, which can be anywhere in the range 103\u2014107 or more. (We shall omit units\nwhen we write molecular weights in this book, but the student is advised to attach the units g/mol\nto these quantities when they appear in problem calculations.)\nSince polymer molecules are made up of chains of repeat units, after the chain itself comes the\nrepeat unit as a structural element of importance. Many polymer molecules are produced by\ncovalently bonding together only one or two types of repeat units. These units are the parts from\nwhich chains are generated; as a class of compounds they are called monomers. Throughout this\nbook, we shall designate the molecular weight of a repeat unit as M0.\nThe degree of polymerization of a polymer is simply the number of repeat units in a molecule.\nThe degree of polymerization N is given by the ratio of the molecular weight of the polymer to the\nmolecular weight of the repeat unit:\n"]], ["block_3", ["are the intramolecular forces that keep the polymer molecule intact. In addition, the usual types of\nintermolecular forces\u2014hydrogen bonds, dipole\u2014dipole interactions, London forces, etc.\u2014hold\nassemblies of these molecules together in the bulk state. The only thing that is remarkable about\nthese molecules is their size, but that feature is remarkable indeed. Another useful term is\nmacromolecule, which of course simply means \u201clarge (or long) molecule.\u201d Some practitioners\ndraw a distinction between the two: all polymers are macromolecules, but not all macromolecules\nare polymers. For example, a protein is not made by repeating one or two chemical units many\ntimes, but involves a precise selection from among 20 different amino acids; thus it is a macro\u2014\nmolecule, but not a polymer. In this text we will not be sticklers for formality, and will use the\nterms rather interchangeably, but the reader should be aware of the distinction.\n"]], ["block_4", ["One type of polymerization reaction is the addition reaction in which successive repeat units add\non to the chain. No other product molecules are formed, so the molecular weight of the monomer\nand that of the repeat unit are identical in this case. A second category of polymerization reaction is\nthe condensation reaction, in which one or two small molecules such as water or HCl are\neliminated for each chain linkage formed. In this case the molecular weight of the monomer and\nthe repeat unit are somewhat different. For example, suppose an acid (subscript A) reacts with an\nalcohol (subscript B) to produce an ester linkage and a water molecule. The molecular weight of\nthe ester\u2014the repeat unit if an entire chain is built up this way\u2014differs from the combined weight\nof the reactants by twice the molecular weight of the water; therefore,\n"]], ["block_5", ["How Big Is Big?\n3\n"]], ["block_6", ["1.2.1\nMolecular Weight\n"]], ["block_7", ["1.2\nHow Big ls Big?\n"]], ["block_8", ["N \nM0\n(1.2.1)\n"]], ["block_9", ["M\nM\nN:\u2014:\nM0\nMA +MB ZMHZO\n(1.2.2)\n"]]], "page_20": [["block_0", [{"image_0": "20_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "20_1.png", "coords": [27, 298, 218, 399], "fig_type": "figure"}]], ["block_2", ["as the number of repeat units in a structure increases. By the time polymeric molecular\nsizes are reached, the error associated with failure to distinguish between segments at the\nend and those within the chain is generally less than experimental error. In Section 1.8.2 we\nshall consider a method for polymer molecular weight determination based on chemical\nanalysis for the end groups in a polymer. A corollary of the present discussion is that the\nmethod of end group analysis is applicable only in the case of relatively low molecular weight\npolymers.\nAs suggested above, not all polymers are constructed by bonding together a single kind of\nrepeat unit. For example, although protein molecules are polyamides in which N amino acid repeat\nunits are bonded together, the degree of polymerization is a less useful concept, since an amino\nacid unit might be any one of the 20\u2014odd molecules that are found in proteins. in this case the\nmolecular weight itself, rather than the degree of polymerization, is generally used to describe the\nmolecule. When the actual content of individual amino acids is known, it is their sequence that is\nof special interest to biochemists and molecular biologists.\n"]], ["block_3", ["Although the difference is almost 5% for propane, it is closer to 0.1% for the case of N m 100,\nwhich is about the threshold for polymers. The precise values of these numbers will be\ndifferent, depending on the specific repeat units and end groups present. For example, if\nM0: 100 and Mend=80, the difference would be 0.39% in a calculation such as that above\nfor N g 100.\n"]], ["block_4", ["As a polymer prototype consider an n-alkane molecule consisting of N\u20142 methylenes and 2 methyl\ngroups. How serious an error is made in M for different Ns if the difference in molecular weight\nbetween methyl and methylene groups is ignored?\n"]], ["block_5", ["The effect of different end groups on M can be seen by comparing the true molecular weight with\nan approximate molecular weight, calculated on the basis of a formula (CH2)N. These Ms and the\npercentage difference between them are listed here for several values of N\n"]], ["block_6", ["a methyl group and the middle parts are methylene groups. Of course, the terminal group does\nnot have to be a hydrogen as in alkanes; indeed, it is often something else. Our interest in end\ngroups is concerned with the question of what effect they introduce into the evaluation of N\nthrough Equation\n(1.2.2).\nThe following\nexample examines\nthis\nthrough\nsome numerical\ncalculations.\n"]], ["block_7", ["4\nIntroduction to Chain Molecules\n"]], ["block_8", ["The end units in a polymer chain are inevitably different from the units that are attached on\nboth sides to other repeat units. We see this situation in the n-alkanes: each end of the chain is\n"]], ["block_9", ["Solution\n"]], ["block_10", ["Example 1.1\n"]], ["block_11", ["12\n170\n168\n1.2\n52\n730\n728\n0.3\n102\n1,430\n1,428\n0.14\n502\n7,030\n7,028\n0.028\n1002\n14,030\n14,028\n0.014\n"]], ["block_12", ["N\nM\nMapprox\n% Difference\n"]], ["block_13", ["The example shows that the contribution of the ends becomes progressively less important\n"]], ["block_14", ["7\n100\n98\n2.0\n"]], ["block_15", ["3\n44\n42\n4.5\n"]]], "page_21": [["block_0", [{"image_0": "21_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nBecause of the possibility of rotation around carbon\u2014carbon bonds, a molecule possessing\nmany such bonds will undergo many twists and turns along the chain.\n4.\nFully extended molecular length is not representative of the spatial extension that a molecule\nactually displays. The latter is sensitive to environmental factors, however, so the extended\nlength is convenient for our present purposes to provide an idea of the spatial size of polymer\nmolecules.\n"]], ["block_2", ["We began this section with an inquiry into how to de\ufb01ne the size of a polymer molecule. In\naddition to the molecular weight or the degree of polymerization, some linear dimension that\ncharacterizes the molecule could also be used for this purpose. As an example, consider a\nhydrocarbon molecule stretched out to its full length but without any bond distortion. There are\nseveral features to note about this situation:\n"]], ["block_3", ["these triangles, one of these atoms is common to two triangles;\ntherefore the number of\ntriangles is the same as the number of pairs of carbon atoms, except where this breaks down\nat the ends of the molecule. If the chain is sufficiently long, this end effect is inconsequential.\nThe law of cosines can be used to calculate the length of the base of each of these triangles:\n[2(0.154)2(1 cos 109.5\u00b0)] 0.252 nm. If the repeat unit of the molecule contributes two carbon\natoms to the backbone of the polymer\u2014as is the case for vinyl polymers\u2014the fully extended\nchain length is given by N(0.252) nm. For a polymer with N 104, this corresponds to 2.52 pm.\nObjects which actually display linear dimensions of this magnitude can be seen in an ordinary\nmicroscope, provided that they have suitable optical properties to contrast with their surroundings;\nan example will be given in Figure 1.1a. Note that the distance between every other carbon atom\nwe have used here is also the distance between the substituents on these carbons for the fully\nextended chains.\nWe shall see in Chapter 6 that, because of all the twists and turns a molecule undergoes,\nthe actual average end-to\u2014end distance of the jumbled molecules increases as N\u201d2. With the\n"]], ["block_4", ["same repeat distance calculated above, but the square root dependence on N, the actual end-to-\nend distance of the coiled chain with N 104 is closer to (104)\u201d2 X 0.252 nm as 25 nm. If we\npicture one end of this jumbled chain at the origin of a coordinate system,.the other end might be\nanywhere on the surface of a sphere whose radius is given by this end-to-end distance. This\nspherical geometry comes about because the random bends occurring along the chain length can\ntake the end of the chain anywhere in a spherical domain whose radius depends on N2.\nThe above discussion points out the difficulty associated with using the linear dimensions of a\nmolecule as a measure of its size: it is not the molecule alone that determines its dimensions, but\nalso the shape or conformation in which it exists. Fully extended, linear arrangements of the sort\ndescribed above exist in polymer crystals, at least for some distance, although usually not over the\nfull length of the chain. We shall take up the structure of polymer crystals in Chapter 13. In the\nsolution and bulk states, many polymers exist in the coiled form we have also described. Still other\nstructures are important, notably the rod or semi\ufb02exible chain, which we shall discuss in Chapter 6.\nThe overall shape assumed by a polymer molecule can be greatly affected by the environment. The\nshape of a molecule in solution plays a key role in determining many properties of polymer\nsolutions. From a study of these solutions, some conclusions can be drawn regarding the shape of\n"]], ["block_5", ["A fully extended hydrocarbon molecule will have the familiar all\u2014trans zigzag profile in\nwhich the hydrogens extend in front of and in back of the plane containing the carbons, with an\nangle of 109.5\u00b0 between successive carbon\u2014carbon bonds. The chain may be pictured as a row\nof triangles resting comer to comer. The length of the row equals the product of the number\nof triangles and the length of the base of each. Although it takes three carbons to define one of\n"]], ["block_6", ["1.2.2\nSpatial Extent\n"]], ["block_7", ["How Big ls Big?\n5\n"]], ["block_8", ["1.\nThe tetrahedral geometry at the carbon atoms gives bond angles of 109.5\u00b0.\nThe equilibrium bond length of a carbon\u2014carbon single bond is 0.154 nm or 1.54 A.\n"]]], "page_22": [["block_0", [{"image_0": "22_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "22_1.png", "coords": [12, 45, 311, 292], "fig_type": "figure"}]], ["block_2", ["6\nIntroduction to Chain Molecules\n"]], ["block_3", ["Figure 1.1\n(3) Individual molecules of DNA of various sizes, spread on a \ufb02uid positively charged surface,\nimaged by \ufb02uorescence. The scale bar is 10 um. (Reproduced from Maier, B. and Radler, J.O. Macromolecules\n33, 7185, 2000. With permission.) (b) Atomic force microscopy images of three-arm star polymers, where each\narm is a heavily branched Comb. The circles indicate linear molecules. (Reproduced from Matyjaszewski, K.,\nQin, S., Boyce, J.R., Shirvanyants, D., and Sheiko, S.S. Macromolecules 36, 1843, 2003. With permission.)\n"]], ["block_4", ["the molecule in the environment. Relevant aspects of polymer solutions are taken up in Chapter 6\nthrough Chapter 9.\nFigure 1.1a and Figure 1.1b are rather striking images of individual polymer molecules. Figure\n"]], ["block_5", ["1.1a shows single molecules of DNA that have been heavily labeled with \ufb02uorescent dyes; the dyes\n"]]], "page_23": [["block_0", [{"image_0": "23_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["as these.\nWe conclude this section by questioning whether there is a minimum molecular weight or linear\ndimension that must be met for a molecule to qualify as a polymer. Although a dimer is a molecule\nfor which N =2, no one would consider it a polymer. The term oligomer has been coined to\ndesignate molecules for which N < 10. If they require a special name, apparently the latter are not\nfull-\ufb02edged polymers either. At least as a first approximation, we shall take the attitude that there\nis ordinarily no discontinuity in behavior with respect to observed properties as we progress\nthrough a homologous series of compounds with different N values. At one end of the series, we\nmay be dealing with a simple low molecular weight compound, and at the other end with a material\nthat is unquestionably polymeric. The molecular weight and chain length increase monotonically\nthrough this series, and a variety of other properties vary smoothly also. This point of view\nemphasizes continuity with familiar facts concerning the properties of low molecular weight\ncompounds. There are some properties, on the other hand, which follow so closely from the\nchain structure of polymers that the property is simply not observed until a certain critical\nmolecular size has been reached. This critical size is often designated by a threshold molecular\nweight. The elastic behavior of rubber and several other mechanical properties fall into this latter\ncategory. In theoretical developments, large values of N are often assumed to justify neglecting end\neffects, using certain statistical methods and other mathematical approximations. With these ideas\nin mind, M 1000 seems to be a convenient round number for designating a compound to be a\npolymer, although it should be clear that this cutoff is arbitrary (and on the low side).\n"]], ["block_2", ["intercalate between the base pairs along the chain, without seriously altering the conformation of\nthe molecule. Under illumination the resulting \ufb02uorescence provides a good representation of the\nmolecules themselves. In this particular image, the DNA molecules are spread out in two\ndimensions, on a cationically charged imitation lipid membrane. DNA, as it turns out, is an\nexcellent example of a semi\ufb02exible chain, which can actually be inferred from these images; the\nmolecules are not straight rods, but neither are they heavily coiled around themselves. The\nscale bar corresponds to 10 um, indicating that these molecules are of very high molecular\nweight indeed. In Figure 1.1b, the image is of a star-shaped polymer, but one in which each arm\nof the star is a heavily branched comb or \u201cbottlebrush.\u201d The molecule is thus akin to a kind of\nstarfish, with very hairy arms. This picture was obtained by atomic force microscopy (AFM),\none of a series of surface-sensitive analysis techniques with exquisite Spatial resolution. The\nmolecules themselves were deposited from a Langmuir\u2014Blodgett trough onto a mica substrate.\nBoth situations depicted in Figure 1.1a and Figure 1.1b raise the question of the relationship\nbetween the conformation observed on the surface and that at equilibrium in solution. In\nChapter 6 through Chapter 9 we will encounter several ways in which the solution conformation\ncan be determined reliably, which can serve to confirm the impression derived from figures such\n"]], ["block_3", ["Most of the preceding section was based on the implicit assumption that polymer chains are\nlinear (with the striking exception of Figure 1.1b). In evaluating both the degree of polymeriza-\ntion and the extended chain length, we assumed that the chain had only two ends. While linear\npolymers are important, they are not the only type of molecules possible: branched and cross-\nlinked molecules are also common. When we speak of a branched polymer, we refer to the\npresence of additional polymeric chains issuing from the backbone of a linear molecule. (Small\nsubstituent groups such as methyl or phenyl groups on the repeat units are generally not\nconsidered branches, or, if they are, they should be specified as \u201cshort\u2014chain branches\u201d)\nBranching can arise through several routes. One is to introduce into the polymerization reaction\nsome monomer with the capability of serving as a branch. Consider the formation of a polyester.\n"]], ["block_4", ["Linear and Branched Polymers, Homopolymers, and Copolymers\n7\n"]], ["block_5", ["1.3.1\nBranched Structures\n"]], ["block_6", ["1.3\nLinear and Branched Polymers, Homopolymers, and Copolymers\n"]]], "page_24": [["block_0", [{"image_0": "24_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "24_1.png", "coords": [27, 455, 212, 670], "fig_type": "figure"}]], ["block_2", ["The presence of difunctional acids and difunctional alcohols allows the polymer chain to\ngrow. These difunctional molecules are incorporated into the chain with ester linkages at both\nends of each. Trifunctional acids or alcohols, on the other hand, produce a linear molecule by\nthe reaction of two of their functional groups. If the third reacts and the resulting chain continues\nto grow, a branch has been introduced into the original chain. A second route is through\nadventitious branching, for example, as a result of an atom being abstracted from the original\nlinear molecule, with chain growth occurring from the resulting active site. This is quite a\ncommon occurrence in the free-radical polymerization of ethylene, for example. A third route is\ngrafting, whereby pre-formed but still reactive polymer chains can be added to sites along an\nexisting backbone (so-called \u201cgrafting to\u201d), or where multiple initiation sites along a chain\ncan be exposed to monomer (so-called \u201cgrafting from\u201d).\nThe amount of branching introduced into a polymer is an additional variable that must be\nspecified for the molecule to be fully characterized. When only a slight degree of branching is\npresent, the concentration of junction points is sufficiently low that these may be simply related to\nthe number of chain ends. For example, two separate linear molecules have a total of four ends. If\nthe end of one of these linear molecules attaches itself to the middle of the other to form a T, the\nresulting molecule has three ends. It is easy to generalize this result. If a molecule has 12 branches, it\nhas 12 + 2 chain ends if the branching is relatively low. Two limiting cases to consider, illustrated in\nFigure 1.2, are combs and stars. In the former, a series of relatively uniform branches emanate\nfrom along the length of a common backbone; in the latter, all branches radiate from a central\njunction. Figure 1.1b gave an example of both of these features.\nIf the concentration of junction points is high enough, even branches will contain branches.\nEventually a point can be reached at which the amount of branching is so extensive that the\npolymer molecule becomes a giant three-dimensional network. When this condition is achieved,\nthe molecule is said to be cross-linked. In this case, an entire macroscopic object may be\nconsidered to consist of essentially one molecule. The forces that give cohesiveness to such a\nbody are covalent bonds, not intermolecular forces. Accordingly, the mechanical behavior of\ncross-linked bodies is much different from those without cross-linking. This will be discussed at\nlength in Chapter 10. However, it is also possible to suppress cross-linking such that the highly\nbranChed molecules remain as discrete entities, known as hyperbranched polymers (see Figure 1.2).\nAnother important class of highly branched polymers illustrated in Figure 1.2 are dendrimers, or\ntreelike molecules. These are completely regular structures, with well-defined molecular weights,\nthat are made by the successive condensation of branched monomers. For example, begin with a\n"]], ["block_3", [{"image_2": "24_2.png", "coords": [34, 606, 95, 645], "fig_type": "molecule"}]], ["block_4", ["Four-arm star\n"]], ["block_5", ["Figure 1.2\nIllustration of various polymer architectures.\n"]], ["block_6", ["8\nIntroduction to Chain Molecules\n"]], ["block_7", [{"image_3": "24_3.png", "coords": [37, 473, 173, 588], "fig_type": "figure"}]], ["block_8", ["Linear\n"]], ["block_9", ["Comb\nDendrimer\n"]], ["block_10", ["Cycle\n"]], ["block_11", [{"image_4": "24_4.png", "coords": [49, 454, 176, 567], "fig_type": "molecule"}]], ["block_12", [{"image_5": "24_5.png", "coords": [61, 477, 178, 531], "fig_type": "molecule"}]], ["block_13", [{"image_6": "24_6.png", "coords": [87, 537, 185, 648], "fig_type": "molecule"}]], ["block_14", ["Hyperbranched\n"]]], "page_25": [["block_0", [{"image_0": "25_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The empirical formula gives the relative amounts of A, B, and C in the terpolymer. The total\nmolecular weight of this empirical formula unit is given by adding the molecular weight contri-\nbutions of A, B, and C: 3.44(104) + 2.20(184)+ 1.00(128) =902 g/mol per empirical formula unit.\n"]], ["block_2", ["trifunctional monomer \u201cB3,\u201d or \u201cgeneration 0.\u201d This is reacted with an excess of AB2 monomers,\nleading to a generation\n1 dendrimer with 6 B groups. A second reaction with AB2 leads to\ngeneration 2 with 12 pendant B groups. Eventually, perhaps at generation 6 or 7, the surface of\nthe molecule becomes so congested that addition of further complete generations is impossible.\nNote that the \u201cB\u201d part of the ABg monomer needs to be protected in some way so that only one\ngeneration can be added at one time.\nA final class of nonlinear polymers to consider are cycles or rings, whereby the two ends of the\nmolecule react to close the loop. Such polymers are currently more of academic interest than\ncommercial importance, as they are tricky to prepare, but they can shed light on various aspects of\npolymer behavior. Interestingly, nature makes use of this architecture; the DNA of the Lambda\nbacteriophage reversibly cyclizes and uncyclizes during gene expression.\n"]], ["block_3", ["Just as it is not necessary for polymer chains to be linear, it is also not necessary for all repeat units\nto be the same. We have already mentioned proteins, where a wide variety of different repeat\nunits are present. Among synthetic polymers, those with a single kind of repeat unit are called\nhomopolymers, and those containing more than one kind of repeat unit are capolymers. Note that\nthese definitions are based on the repeat unit, not the monomer. An ordinary polyester is not really\na copolymer, even though two different monomers, acids and alcohols, are its monomers.\nBy contrast, copolymers result when different monomers bond together in the same way to produce\na chain in which each kind of monomer retains its respective substituents in the polymer molecule.\nThe unmodified term copolymer is generally used to designate the case where two different repeat\nunits are involved. Where three kinds of repeat units are present, the system is called a terpolymer;\nwhere there are more than three, the system is called a multicomponent copolymer. The copoly-\nmers we discuss in this book will be primarily two\u2014component molecules. We shall explore aspects\nof the synthesis and characterization of copolymers in both Chapter 4 and Chapter 5.\nThe moment we admit the possibility of having more than one kind of repeat unit, we require\nadditional variables to describe the polymer. First, we must know how many kinds of repeat units\nare present and what they are. To describe the copolymer quantitatively, the relative amounts of the\ndifferent kinds of repeat units must be specified. Thus the empirical formula of a copolymer may\nbe written A,C By, where A and B signify the individual repeat units and x and y indicate the relative\nnumber of each. From a knowledge of the molecular weight of the polymer, the molecular weights\nof A and B, and the values of x and y, it is possible to calculate the number of each kind of\nmonomer unit in the copolymer. The sum of these values gives the degree of polymerization of the\ncopolymer. The following example illustrates some of the ways of describing a copolymer.\n"]], ["block_4", ["A terpolymer is prepared from vinyl monomers A, B, and C; the molecular weights of the repeat\nunits are 104, 184, and 128, respectively. A particular polymerization procedure yields a product\nwith the empirical formula A355 32.20C1.00- The authors of this research state that the terpolymer\nhas \u201can average unit weight of 134\u201d and \u201cthe average molecular weight per angstrom of 53.5.\u201d\nVerify these values.T\n"]], ["block_5", ["TA. Ravve and IT Khamis, Addition and Condensation Polymerization Processes, Advances in Chemistry Series, Vol. 91,\nAmerican Chemical Society Publications, Washington, DC, 1969.\n"]], ["block_6", ["Linear and Branched Polymers, Homopolymers, and Copolymers\n9\n"]], ["block_7", ["Solution\n"]], ["block_8", ["1.3.2\nCopolymers\n"]], ["block_9", ["Example 1.2\n"]]], "page_26": [["block_0", [{"image_0": "26_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The total amount of chain repeat units possessing this total weight is 3.55 + 2.20 + 1.00 6.75\nrepeat units per empirical formula unit. The ratio of the total molecular weight to the total number\nof repeat units gives the average molecular weight per repeat unit:\n"]], ["block_2", ["With copolymers, it is far from sufficient merely to describe the empirical formula to charac-\nterize the molecule. Another question that must be asked concerns the location of the different\nkinds of repeat units within the molecule. Starting from monomers A and B, the following\ndistribution patterns can be obtained in linear polymers:\n"]], ["block_3", ["Since the monomers are specified to be vinyl monomers, each contributes two carbon atoms to the\npolymer backbone, with the associated extended length of 0.252 mm per repeat unit. Therefore, the\ntotal extended length of the empirical formula unit is\n"]], ["block_4", ["The ratio of the total weight to the total extended length of the empirical formula unit gives the\naverage molecular weight per length of chain:\n"]], ["block_5", ["4.\nGraft. This segregation is often accomplished by first homopolymerizing the backbone. This is\ndissolved in the second monomer, with sites along the original chain becoming the origin of\nthe comonomer side-chain growth:\n"]], ["block_6", ["Note that the average weight per repeat unit could be used to evaluate the overall degree of\npolymerization of this terpolymer. For example, if the molecular weight were 43,000, the corre-\nsponding degree of polymerization would be\n"]], ["block_7", ["The above structure has three blocks, and is called poly(A-block\u2014B\u2014block-A), or an ABA\ntriblock copolymer. If a copolymer is branched with different repeat units occurring in the\n"]], ["block_8", ["branches and the backbone, we can have the following:\n"]], ["block_9", ["Such a polymer could be called poly(A-stat-B) or poly(A-ran\u2014B).\n2.\nAlternating. A regular pattern of alternating repeat units in poly(A\u2014alt\u2014B):\n"]], ["block_10", ["3.\nBlock. Long, uninterrupted sequence of each monomer is the pattern:\n"]], ["block_11", ["10\nIntroduction to Chain Molecules\n"]], ["block_12", ["1.\nRandom (or statistical). The A\u2014B sequence is governed strictly by chance, subject only to the\nrelative abundances of repeat units. For equal proportions of A and B, we might have\nstructures like\n"]], ["block_13", ["5\u201475\n= 134 g/mol per repeat unit\n"]], ["block_14", ["902\n.\n\u2014 = 53 g/mol per A\n17\n"]], ["block_15", ["43 ,000\n"]], ["block_16", ["6.75(0.252 nm) 1.79 nm 17.0 A\n"]], ["block_17", ["902\n"]], ["block_18", ["l\nwAAAAAAAAAAAAAAAAA\u2014\n"]], ["block_19", ["l\nl\n\u2014BBBBBBBBBB\nBBBBBBB\u2014\n"]], ["block_20", ["134\n= 321 repeat units per molecule\n"]], ["block_21", ["\u2014\u2014ABABABABABAB\u2014\n"]], ["block_22", ["\u2014AAAAAAAAAAAAAABBBBBBBBBBBBBBBAAAAAAAAA\u2014\n"]], ["block_23", ["\u2014AAABABAABBABBB\u2014\n"]], ["block_24", ["BBBBBBBBBB\u2014\n"]]], "page_27": [["block_0", [{"image_0": "27_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In a cross\u2014linked polymer, the junction units are different kinds of monomers than the chain\nrepeat units, so these molecules might be considered to be still another comonomer. While the chemical\nreactions that yield such substances are technically copolymerizations, the productsare\ndescribed as cross-linked rather copolymers. In this instance, the behavior due \ntakes precedence over the presence of an additional type of monomer in the structure.\nIt is apparent from items\n1\u2014-3 above that linear copolymers\u2014even those with the same\nproportions of different kinds of repeat units\u2014can be very different in structure and properties.\nIn classifying a copolymer as random, alternating, or block, it should be realized that we are\ndescribing the average character of the molecule; accidental variations from the basic patterns may\nbe present. Furthermore, in some circumstances, nominally \u201crandom\u201d copolymers have\nsubstantial sequences of one or the other. In Chapter 5, we shall see how an experimental\ninvestigation of the sequence of repeat units in a copolymer is a valuable tool for understanding\ncopolymerization reactions.\n"]], ["block_2", ["In the last section, we examined some of the categories into which polymers can be classified.\nVarious aspects of molecular structure were used as the basis for classification in that section. Next\nwe shall consider the chemical reactions that produce the molecules as a basis for classification.\nThe objective of this discussion is simply to provide some orientation and to introduce some\ntypical polymers. For this purpose, many polymers may be classified as being either addition or\ncondensation polymers; both of these classes are discussed in detail in Chapter 2 and Chapter 3,\nrespectively. Even though these categories are based on the reactions which produce the polymers,\nit should not be inferred that only two types of polymerization reactions exist. We have to start\nsomewhere, and these two important categories are the usual places to begin.\n"]], ["block_3", ["Addition, Condensation, and Natural Polymers\n11\n"]], ["block_4", ["4.\nThe polymer repeat unit arises from reacting together two different functional groups, which\nusually originate on different monomers. In this case, the repeat unit is different from either of\n"]], ["block_5", ["These two categories of polymers can be developed along several lines. For example, in addition-\ntype polymers the following statements apply:\n"]], ["block_6", ["1.\nThe repeat unit in the polymer and the monomer has the same composition, although, of\ncourse, the bonding is different in each.\n2.\nThe mechanism of these reactions places addition polymerizations in the kinetic category of chain\nreactions, with either free radicals or ionic groups responsible for propagating the chain reaction.\n3.\nThe product molecules often have an all-carbon chain backbone, with pendant substituent groups.\n"]], ["block_7", ["the monomers. In addition, small molecules are often eliminated during the condensation\nreaction. Note the words usual and often in the previous statements; exceptions to both\nstatements are easily found.\n5.\nThe mechanistic aspect of these reactions can be summarized by saying that the reactions\noccur in steps. Thus, the formation of an ester linkage between two small molecules is not\nessentially different from that between a polyester and a monomer.\n6.\nThe product molecules have the functional groups formed by the condensation reactions\ninterspersed regularly along the backbone of the polymer molecule:\n"]], ["block_8", ["1.4\nAddition, Condensation, and Natural Polymers\n"]], ["block_9", ["1.4.1\nAddition and Condensation Polymers\n"]], ["block_10", ["In contrast, for condensation polymers:\n"]], ["block_11", ["\u2014C\u2014C\u2014Y\u2014C\u2014C\u2014Y\u2014\n"]]], "page_28": [["block_0", [{"image_0": "28_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["First, the reactive center must be initiated by\na suitable reaction to produce a free radical,\nanionic, or cationic reaction site. Next, this reactive entity adds consecutive monomer to\npropagate the polymer chain. Finally, the active site is capped off, terminating the polymer\nformation. If one assumes that the polymer produced is truly a high molecular weight sub-\nstance, the lack of uniformity at the two ends of the chain\u2014arising in one case from the\ninitiation and in the other from the termination\u2014can be neglected. Accordingly, the overall\nreaction can be written\n"]], ["block_2", ["Next let us consider a few Speci\ufb01c examples of these classes of polymers. The addition\npolymerization of a vinyl monomer CsCHX involves\nthree distinctly different\nsteps.\n"]], ["block_3", ["as the conditions of the reaction and other pertinent information will be discussed when these\nreactions are considered in Chapter 3 and Chapter 4. Table 1.1 lists several important addition\npolymers, showing each monomer and polymer structure in the manner of Reaction (1.A). Also\nincluded in Table 1.1 are the molecular weights of the repeat units and the common names of\nthe polymers. The former will prove helpful in many of the problems in this book; the latter\nwill be discussed in the next section. Poly(ethy1ene oxide) and poly(a-caprolactam) have been\nincluded in this list as examples of the hazards associated with classi\ufb01cation schemes. They\nresemble addition polymers because the molecular weight of the repeat unit and that of the\nmonomer are the same; they resemble condensation polymers because of the heteroatom chain\nbackbone. The reaction mechanism, which might serve as arbiter in this case, can be either of the\nchain or the step type, depending on the reaction conditions. These last reactions are examples of\nring-opening polymerizations, yet another possible category of classification.\nThe requirements for formation of condensation polymers are twofold: the monomers must\npossess functional groups capable of reacting to form the linkage, and they ordinarily require\nmore than one reactive group to generate a chain structure. The functional groups can be\ndistributed such that two difunctional monomers with different functional groups react or a single\nmonomer reacts, which is difunctional with one group of each kind. In the latter case especially,\nbut also with condensation polymerization in general, the tendency to form cyclic products from\nintramolecular reactions may compete with the formation of polymers. Condensation polymeriza-\ntions are especially sensitive to impurities. The presence of monofunctional reagents introduces\nthe possibility of a reaction product forming which would not be capable of further growth. If the\nfunctionality is greater than 2, on the other hand, branching becomes possible. Both of these\nmodifications dramatically alter the product compared to a high molecular weight linear product.\nWhen reagents of functionality less than or greater than 2 are added in carefully measured and\ncontrolled amounts, the size and geometry of product molecules can be manipulated. When such\nreactants enter as impurities, the undesired results can be disastrous. Marvel has remarked that\nmore money has been wasted in polymer research by the use of impure monomers than in any other\nmanner [6].\nTable 1.2 lists several examples of condensation reactions and products. Since the reacting\nmonomers can contain different numbers of carbon atoms between functional groups, there are quite\n"]], ["block_4", ["Again, we emphasize that end effects are ignored in writing Reaction (LA). These effects as well\n"]], ["block_5", ["a lot of variations possible among these basic reaction types. The inclusion of poly(dimethylsiloxane)\nin Table 1.2 serves as a reminder that polymers need not be organic compounds. The physical\nproperties of inorganic polymers follow from the chain structure of these molecules, and the concepts\ndeveloped in this volume apply to them and to organic polymers equally well. We shall not examine\nexplicitly the classes and preparations of the various types of inorganic polymers in this text.\n"]], ["block_6", ["12\nIntroduction to Chain Molecules\n"]], ["block_7", [{"image_1": "28_1.png", "coords": [36, 171, 169, 214], "fig_type": "molecule"}]], ["block_8", ["H\n\u2014 \n#4\"\nLA\nn\nH>\u2014exiglas\u00ae, Lucite\u00ae\ne\ne\nH\no\no\no\u201d\n"]], ["block_9", [{"image_3": "29_3.png", "coords": [35, 62, 315, 528], "fig_type": "figure"}]], ["block_10", [{"image_4": "29_4.png", "coords": [41, 389, 92, 442], "fig_type": "molecule"}]], ["block_11", ["H \\\nH\nn\n104\nPolystyrene\n"]], ["block_12", ["H\nMe\n>=<\n56.0\nn\nPolyisobutylene\nH\nMe\nMe Me\n"]], ["block_13", ["H)\n(H\nm\nPolyacrylonitrile\n"]], ["block_14", ["H\n"]], ["block_15", ["H\nH\n.\nPoly(vmyl\nH>'_ Ag)\u201d \nchloride), \u201cvinyl\"\n"]], ["block_16", ["H\nCl\n.\n.\n.\n>_<\n97.0\nW\u201d\nPoly(vrnylidene chloride)\n"]], ["block_17", ["H\nCI\nCl\nCl\n"]], ["block_18", ["H\nH\n>=<\n28.0\nN\u201d\nPolyethylene\nH\nH\n"]], ["block_19", ["E\nC\n100\nPoly(tetra\ufb02uoroethylene),\n\u2014\nn\nTe\ufb02on\u00ae\nF\nF\nF\nF\n"]], ["block_20", ["F\nF\nF\nF\n"]], ["block_21", ["Me\nMe\nPoly(methyl methacrylate),\n"]], ["block_22", ["H\n"]], ["block_23", ["\u2014\n"]], ["block_24", ["0\nPoly(ethy1ene oxide),\n\\g\n44'0\nW\n9\u2018\npoly(ethylene glycol)\n"]], ["block_25", ["ll\n0\nW\n4\nPoly(e\u2014caprolactam),\nN\n1 13\n50\n\"\nNylon-6\n\\H\n"]], ["block_26", ["CN\n"]], ["block_27", ["53.0\n"]], ["block_28", [{"image_5": "29_5.png", "coords": [209, 88, 401, 272], "fig_type": "figure"}]], ["block_29", [{"image_6": "29_6.png", "coords": [235, 524, 289, 575], "fig_type": "molecule"}]], ["block_30", [{"image_7": "29_7.png", "coords": [237, 339, 295, 392], "fig_type": "molecule"}]], ["block_31", [{"image_8": "29_8.png", "coords": [239, 494, 288, 523], "fig_type": "molecule"}]], ["block_32", ["ON\u201d\n\u2018Cacrylic\u201d\n"]]], "page_30": [["block_0", [{"image_0": "30_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "30_1.png", "coords": [33, 386, 298, 424], "fig_type": "molecule"}]], ["block_2", ["Table 1.2\nReactions by Which Several Important Condensation Polymers Can Be Produced\n"]], ["block_3", ["their molecules. For example, the denaturation of a protein involves an overall conformation\nchange from a \u201cnative\u201d state, often a compact globule, to a random coil. As another example,\nthe elasticity and integrity of a cell membrane is often the result of an underlying network of\nfibrillar proteins, with the origin of the elasticity residing in the same conformational entropy as in\na rubber band. Consequently, although we will not discuss the synthesis by, and contribution to the\nfunction of, living organisms by such biopolymers, many of the principles we will develop in detail\napply equally well to natural polymers.\nAs examples of natural polymers, we consider polysaccharides, proteins, and nucleic acids.\nAnother important natural polymer, polyisoprene, will be considered in Section 1.6. Polysaccharides\n"]], ["block_4", ["4. Polycarbonate\no\n\u201c.49\nre\n9\nn CIJLCI + n HOQcQ\u2014QH {o\u2014Qc @o\u2014c} \n"]], ["block_5", ["14\nIntroduction to Chain Molecules\n"]], ["block_6", ["2. Polyamide\n"]], ["block_7", ["5. Inorganic\n"]], ["block_8", ["3. Polyurethane\n"]], ["block_9", ["1. Polyester\n"]], ["block_10", [{"image_2": "30_2.png", "coords": [52, 91, 317, 130], "fig_type": "molecule"}]], ["block_11", ["HO\nO\nO\nO\nn HO/\\/OH\n+\nn m\n-\u2014I-\u2014\nof\n+\n2n H20\n0\nOH\nO\nn\nPoly(ethylene terephthalate), Terylene\u00ae,\nDacron\u00ae, Mylar\u00ae\n"]], ["block_12", ["nocN\\(\\A,NCO+nH0\u2019(\u201d\\i0H\u2014*\u2014t\"'\u00e9JLJOLo\u2019bio\n"]], ["block_13", [{"image_3": "30_3.png", "coords": [74, 225, 282, 264], "fig_type": "molecule"}]], ["block_14", ["n\nH2N\nNH2\nn M\n)-\nN\nN\nM;\n+\nCl\n4\nCl\n*4;\n4\n+\n2n HCl\n"]], ["block_15", [{"image_4": "30_4.png", "coords": [99, 458, 361, 519], "fig_type": "molecule"}]], ["block_16", [{"image_5": "30_5.png", "coords": [116, 471, 357, 502], "fig_type": "molecule"}]], ["block_17", ["Me\nI\n|\nn\nMe-SI'I\u2014CI + :1 H20\n\u2014\u20141-\u2014\nSli\u2014O}\n+\n2n HCI\n"]], ["block_18", [{"image_6": "30_6.png", "coords": [130, 91, 423, 142], "fig_type": "molecule"}]], ["block_19", [{"image_7": "30_7.png", "coords": [136, 157, 346, 194], "fig_type": "molecule"}]], ["block_20", [{"image_8": "30_8.png", "coords": [137, 225, 400, 275], "fig_type": "molecule"}]], ["block_21", ["n Ho)\\(\\/)/U\\10\n+\nn\nH20\n"]], ["block_22", ["Me\nMe\nn\nPoly(4,4-isopropylidenediphenylene carbonate)\nbisphenol A polycarbonate, Lexan\u00ae\n"]], ["block_23", ["Cl\nn\nPoly(dimethylsiloxane)\n"]], ["block_24", ["Cts\n"]], ["block_25", [{"image_9": "30_9.png", "coords": [184, 467, 345, 515], "fig_type": "molecule"}]], ["block_26", [{"image_10": "30_10.png", "coords": [194, 375, 426, 432], "fig_type": "molecule"}]], ["block_27", ["0\nCts\n"]], ["block_28", [{"image_11": "30_11.png", "coords": [201, 91, 393, 130], "fig_type": "molecule"}]], ["block_29", [{"image_12": "30_12.png", "coords": [218, 303, 369, 339], "fig_type": "molecule"}]], ["block_30", ["Poly(tetramethylenehexamethylene urethane),\nSpandex\u00ae, Perlon\u00ae\n"]], ["block_31", [{"image_13": "30_13.png", "coords": [223, 158, 331, 192], "fig_type": "molecule"}]], ["block_32", ["Poly(hexamethylene adipamide), Nylon-6,6\n"]], ["block_33", ["Poly(12\u2014hydroxystearic acid)\n"]], ["block_34", [{"image_14": "30_14.png", "coords": [225, 229, 389, 268], "fig_type": "molecule"}]], ["block_35", ["0\nO n\n"]]], "page_31": [["block_0", [{"image_0": "31_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "31_1.png", "coords": [27, 159, 375, 280], "fig_type": "figure"}]], ["block_2", [{"image_2": "31_2.png", "coords": [29, 193, 350, 399], "fig_type": "figure"}]], ["block_3", ["are macromolecules which make up a large part of the bulk of the vegetable kingdom. Cellulose\nand starch are, respectively, the first and second most abundant organic compounds in plants. The\nformer is present in leaves and grasses; the latter in fruits, stems, and roots. Because of their\nabundance in nature and because of contemporary interest in renewable resources, there is a great\ndeal of interest in these compounds. Both cellulose and starch are hydrolyzed by acids to D-glucose,\nthe repeat unit in both polymer chains. The con\ufb01guration of the glucoside linkage is different in the\ntwo, however. Structure (1.1) and Structure (1.11), respectively, illustrate that the linkage is a B-acetal\u2014\nhydrolyzable to an equatorial hydroxide\u2014in cellulose and an a-acetal\u2014hydrolyzable to an axial\nhydroxide\u2014in amylose, a starch:\n"]], ["block_4", ["Amylopectin and glycogen are saccharides similar to amylose, except with branched chains.\nThe cellulose repeat unit contains three hydroxyl groups, which can react and leave the chain\nbackbone intact. These alcohol groups can be esterified with acetic anhydride to form cellulose\nacetate; this polymer is spun into the fiber acetate rayon. Similarly, the alcohol groups in cellulose\nreact with CS2 in the presence of strong base to produce cellulose xanthates. When extruded into\nfibers, this material is called viscose rayon, and when extruded into sheets, cellophane. In both the\nacetate and xanthate formation, some chain degradation also occurs, so the resulting polymer\nchains are shorter than those in the starting cellulose. The hydroxyl groups are also commonly\nmethylated, ethylated, and hydroxypropylated for a variety of aqueous applications, including food\nproducts. A closely related polysaccharide is chitin, the second most abundant polysaccharide in\nnature, which is found for example in the shells of crabs and beetles. Here one of the hydroxyls on\neach repeat unit of cellulose is replaced with an \u2014NHCO\u2014CH3 amide group. This is converted to a\nprimary amine \u2014NH2 in chitosan, a derivative of chitin finding increasing applications in a variety of\nfields.\nAs noted above, proteins are polyamides in which a-amino acids make up the repeat units, as\nshown by Structure (1.III):\n"]], ["block_5", ["These molecules are also called polypeptides, especially when M 3 10,000. The various amino\nacids differ in their R groups. The nature of R, the name, and the abbreviation used to represent\n"]], ["block_6", ["Addition, Condensation, and Natural Polymers\n15\n"]], ["block_7", [{"image_3": "31_3.png", "coords": [43, 594, 108, 643], "fig_type": "molecule"}]], ["block_8", ["H\nO\n{/\u201cH\ufb01\n(1.111)\n"]], ["block_9", ["0\n0\n(1.1)\nHO\nHO\n0\nH\nOH\nH\nOH\n"]], ["block_10", ["R\n11\n"]], ["block_11", ["H\nOH\nH OH\n"]], ["block_12", ["H\nH\nH\nH\n"]], ["block_13", [{"image_4": "31_4.png", "coords": [136, 278, 342, 404], "fig_type": "figure"}]], ["block_14", [{"image_5": "31_5.png", "coords": [206, 159, 345, 278], "fig_type": "figure"}]], ["block_15", ["(1.11)\n"]]], "page_32": [["block_0", [{"image_0": "32_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 1.3\nName, Abbreviations, and R Group for Common Amino Acids\n"]], ["block_2", ["Lysine\nLys\nVNHz\n"]], ["block_3", ["HOOCYNHQ\nName\nAbbreviation\nR Group\nR\n"]], ["block_4", ["Me\nAlanine\nAla\nJ\u201d\n"]], ["block_5", ["Aspartic acid\nAsp\nKLOH\n"]], ["block_6", ["Glutamine\nGln\nm\n"]], ["block_7", ["H\nGlycine\nGly\n\u201c1\u201c,\n"]], ["block_8", ["NH\n/ x)\nHistidine\nHis\nN\n"]], ["block_9", ["H\nNH\nArginine\nArg\np Y\nNH2\n"]], ["block_10", ["Asparagine\nAsn\n\ufb01 NH2\n"]], ["block_11", ["Glutamic acid\nGlu\nm\n"]], ["block_12", ["Me\nMe\nIsoleucine\nIle\n"]], ["block_13", ["Leucine\nLeu\n\ufb01me\n"]], ["block_14", ["Cysteine\nCys\nNE\n"]], ["block_15", ["16\nIntroduction to Chain Molecules\n"]], ["block_16", [{"image_1": "32_1.png", "coords": [239, 140, 321, 187], "fig_type": "molecule"}]], ["block_17", [{"image_2": "32_2.png", "coords": [247, 620, 316, 644], "fig_type": "molecule"}]], ["block_18", [{"image_3": "32_3.png", "coords": [253, 512, 310, 560], "fig_type": "molecule"}]], ["block_19", [{"image_4": "32_4.png", "coords": [253, 374, 310, 421], "fig_type": "molecule"}]], ["block_20", [{"image_5": "32_5.png", "coords": [254, 185, 307, 239], "fig_type": "molecule"}]], ["block_21", [{"image_6": "32_6.png", "coords": [261, 457, 301, 502], "fig_type": "molecule"}]], ["block_22", [{"image_7": "32_7.png", "coords": [262, 243, 300, 282], "fig_type": "molecule"}]], ["block_23", [{"image_8": "32_8.png", "coords": [264, 330, 311, 368], "fig_type": "molecule"}]], ["block_24", ["0\n"]], ["block_25", ["Me\n"]], ["block_26", ["0\n"]], ["block_27", ["SH\n"]], ["block_28", ["NH2\n"]], ["block_29", ["OH\n"]]], "page_33": [["block_0", [{"image_0": "33_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "33_1.png", "coords": [9, 70, 314, 404], "fig_type": "figure"}]], ["block_2", ["Tyrosine\nTyr\nY\nAL/O/\n"]], ["block_3", ["/\nMethionine\nMet\nM\nf\nS\n"]], ["block_4", ["O\nM\nThreonine\nThr\nT\nI\n"]], ["block_5", ["Tryptophan\nTrp\nW\nI\nE\n"]], ["block_6", ["Valine\nVal\nV\n"]], ["block_7", ["some of the more common amino acids are listed in Table 1.3. In proline (Pro) the nitrogen and the\no-carbon are part of a five-atom pyrrolidine ring. Since some of the amino acids carry substituent\ncarboxyl or amino groups, protein molecules are charged in aqueous solutions, and hence can\nmigrate in electric \ufb01elds. This is the basis of electrOphoresis as a means of separating and\nidentifying proteins.\nIt is conventional to speak of three levels of structure in protein molecules:\n"]], ["block_8", ["Addition, \n17\n"]], ["block_9", ["Table \n"]], ["block_10", [".._\u2014\u2014\n"]], ["block_11", ["1.\nPrimary structure refers to the sequence of amino acids in the polyamide chain.\n2.\nSecondary structure refers to the regions of the molecule that have particular spatial arrange-\nments. Examples in proteins include the o-helix and the B-sheet.\n"]], ["block_12", ["Name\nAbbreviation\nCode\nR Group\n\\Fll/\n"]], ["block_13", ["Phenylalanine\nPhe\nF\nJ/Q\n"]], ["block_14", ["H\nProline\nPro\nP\n"]], ["block_15", ["Serine\nSer\nS\n"]], ["block_16", ["Letter\nHOOC\nN H2\n"]], ["block_17", [{"image_2": "33_2.png", "coords": [246, 423, 308, 476], "fig_type": "molecule"}]], ["block_18", [{"image_3": "33_3.png", "coords": [248, 105, 311, 136], "fig_type": "molecule"}]], ["block_19", [{"image_4": "33_4.png", "coords": [251, 204, 309, 253], "fig_type": "molecule"}]], ["block_20", [{"image_5": "33_5.png", "coords": [255, 488, 299, 515], "fig_type": "molecule"}]], ["block_21", ["H\n"]], ["block_22", ["Mel\n"]], ["block_23", ["'?'\nH\nO\n"]], ["block_24", ["to\u201d\n"]], ["block_25", ["\\\n"]], ["block_26", ["Me\n"]], ["block_27", ["O\n"]], ["block_28", ["N\n"]], ["block_29", ["e\n"]], ["block_30", ["OH\n"]]], "page_34": [["block_0", [{"image_0": "34_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nThe globular proteins albumin in eggs andfibrinogen in blood are converted to insoluble forms\nby modification of their higher-order structure. The process is called denaturation and occurs,\nin the systems mentioned, with the cooking of eggs and the clotting of blood.\n4.\nActin is a fascinating protein that exists in two forms: G-actin (globular) and F\u2014actin (fibrillar).\nThe globular form can polymerize (reversibly) into very long filaments under the in\ufb02uence of\nvarious triggers. These filaments play a crucial role in the cytoskeleton, i.e., in allowing cells\nto maintain their shape. In addition, the uniaxial sliding of actin filaments relative to filaments\nof a related protein, myosin, is responsible for the working of muscles.\n"]], ["block_2", ["3.\nTertiary structure refers to the overall shape of the molecule, for example, a globule perhaps\nstabilized by disul\ufb01de bridges formed by the oxidation of cysteine mercapto groups. By extension\nthe full tertiary structure implies knowledge of the relative spatial positions of all the residues.\n"]], ["block_3", ["Ribonucleic acid (RNA) and deoxyribonucleic acid (DNA) are polymers in which the repeat\nunits are substituted esters. The esters are formed between the hydrogens of phosphoric acid and\nthe hydroxyl groups of a sugar, D\u2014ribose in the case of RNA and D\u20142-deoxyribose in the case of\nDNA. The sugar rings in DNA carry four different kinds of substituents: adenine (A) and guanine\n(G), which are purines, and thymine (T) and cytosine (C), which are pyramidines. The familiar\ndouble-helix structure of the DNA molecule is stabilized by hydrogen bonding between pairs of\nsubstituent base groups: G\u2014C and A\u2014T. In RNA, thymine is usually replaced by uracil (U). The\nreplication of these molecules, the template model of their functioning, and their role in protein\nsynthesis and the genetic code make the study of these polymers among the most exciting and\nactively researched areas in science. As with the biological function of proteins, we will not discuss\nthese phenomena in this book. However, as indicated previously, DNA plays a very important role\n"]], ["block_4", ["Hydrogen bonding stabilizes some protein molecules in helical forms, and disulfide cross\u2014links\nstabilize some protein molecules in globular forms. Both secondary and tertiary levels of structure\nare also in\ufb02uenced by the distribution of polar and nonpolar amino acid molecules relative to the\naqueous environment of the protein molecules. In some cases, individual proteins associate in\nparticular aggregates, which are referred to as quaternary structures.\nExamples of the effects and modifications of the higher-order levels of structures in proteins are\nfound in the following systems:\n"]], ["block_5", ["1.\nCollagen is the protein of connective tissues and skin. In living organisms, the molecules are\nwound around one another to form a three\u2014stranded helix stabilized by hydrogen bonding.\nWhen boiled in water, the collagen dissolves and forms gelatin, thereby establishing a new\nhydrogen bond equilibrium with the solvent. This last solution sets up to form the familiar gel\nwhen cooled, a result of shifting the hydrogen bond equilibrium.\n2.\nKeratin is the protein of hair and wool. These proteins are insoluble because of the disulfide\ncross-linking between cysteine units. Permanent waving of hair involves the rupture of these\nbonds, reshaping of the hair \ufb01bers, and the reformation of cross-links, which hold the chains in\nthe new positions relative to each other. We shall see in Chapter 10 how such cross-linked\nnetworks are restored to their original shape when subjected to distorting forces.\n"]], ["block_6", ["Considering that a simple compound like CZHSOH is variously known as ethanol, ethyl alcohol,\ngrain alcohol, or simply alcohol, it is not too surprising that the vastly more complicated polymer\nmolecules are also often known by a variety of different names. The International Union of Pure\nand Applied Chemistry (IUPAC) has recommended a system of nomenclature based on the\n"]], ["block_7", ["18\nIntroduction to Chain Molecules\n"]], ["block_8", ["as a prototypical semi\ufb02exible polymer, as it is now readily obtainable in pure molecular fractions\nof varying lengths, and because it is readily dissolved in aqueous solution. It is also a charged\npolymer, or polyelectrolyte, and thus serves as a model system in this arena as well.\n"]], ["block_9", ["1.5\nPolymer Nomenclature\n"]]], "page_35": [["block_0", [{"image_0": "35_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["structure \nwidespread usage; these latter names seem even more resistant to replacement than is the case with\nlow molecular polymers of commercial importance are often \nknown by trade names that have more to do with marketing considerations than with scientific\ncommunication. Polymers of biological origin are often described in terms of some aspect of their\nfunction, preparation, or characterization.\nIf a polymer is formed from a single monomer, as in addition and ring-opening polymerizations,\nit is named by attaching the to the name of the monomer. In the IUPAC system, the\nmonomer is named according to the IUPAC recommendations for organic chemistry, and the name\nof the monomer is set off from the prefix by enclosing the former in parentheses. Variations of this\nbasic system often substitutefor the IUPAC name in designating the monomer.\nWhether or not parentheses are used in the latter case is in\ufb02uenced by the complexity of the\nmonomer name; they become more important\nas the number of words in the monomer\nname increases. Thus the polymer (CHg\u2014CHCI)n is called poly(l-chloroethylene) according to\nthe IUPAC system; it commonly called poly(vinyl chloride) or polyvinyl chloride.\nAcronyms are not particularly helpful but are an almost irresistible aSpect of polymer terminology,\n"]], ["block_2", ["as evidenced by the initials PVC, are widely used to describe the polymer just named. The\ntrio of names poly(l-hydroxyethylene), poly(vinyl alcohol), and polyvinyl alcohol emphasizes that\nthe polymer need not actually be formed from the reaction of the monomer named; this polymer is\nactually prepared by the hydrolysis of poly(l-acetoxyethylene), otherwise known as poly(vinyl\nacetate). These same alternatives are used in naming polymers formed by ring-opening reactions;\nfor example, poly(6-aminohexanoic acid), poly(6-aminocapr0ic acid), and poly(s-caprolactam) are\nall more or less acceptable names for the same polymer.\nThose polymers which are the condensation products of two different monomers are named by\napplying the preceding rules to the repeat unit. For example, the polyester formed by the condensation\nof ethylene glycol and terephthalic acid is called poly(oxyethylene oxyterphthaloyl) according to the\nIUPAC system, but is more commonly referred to as poly(ethylene terephthalate) or polyethylene\nterephthalate. The polyamides poly(hexamethylene sebacamide) and poly(hexamethylene adipamide)\nare also widely known as nylon-6,10 and nylon-6,6, respectively. The numbers following the word\nnylon indicate the number of carbon atoms in the diamine and dicarboxylic acids, in that order. On the\nbasis of this system, poly(s\u2014caprolactam) is also known as nylon-6.\nMany of the polymers in Table 1.1 and Table 1.2 are listed with more than one name. Also listed\nare some of the registered trade names by which these substances\u2014or materials which are mostly\nof the indicated structure~\u2014are sold commercially. Some commercially important cross-linked\npolymers go virtually without names. These are heavily and randomly cross-linked polymers\nwhich are insoluble and infusible and therefore widely used in the manufacture of such molded\nitems as automobile and household appliance parts. These materials are called resins and, at best,\nare named by specifying the monomers that go into their production. Often even this information is\nsketchy. Examples of this situation are provided by phenol-formaldehyde and urea\u2014formaldehyde\nresins, for which typical structures are given by Structure (1.1V) and Structure (1.V), respectively:\n"]], ["block_3", ["Polymer Nomenclature\n19\n"]], ["block_4", [{"image_1": "35_1.png", "coords": [41, 601, 127, 671], "fig_type": "figure"}]], ["block_5", [{"image_2": "35_2.png", "coords": [41, 536, 114, 609], "fig_type": "molecule"}]], ["block_6", ["0H\n"]], ["block_7", ["|\\\n(1.V)\n"]], ["block_8", ["(LIV)\n"]]], "page_36": [["block_0", [{"image_0": "36_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "36_1.png", "coords": [32, 225, 253, 324], "fig_type": "figure"}]], ["block_2", ["20\nIntroduction to Chain Molecules\n"]], ["block_3", ["In this section, we shall consider three types of isomerism that are encountered in polymers. These\n"]], ["block_4", ["Structure (1.VI) and Structure (1.VII), respectively, are said to arise from head-to-tat'l or head-to-\nhead orientations. In this terminology, the substituted carbon is defined to be the head and the\nmethylene is the tail. Tail-to-tail linking is also possible.\nFor most vinyl polymers, head-to-tail addition is the dominant mode of addition. Variations\nfrom this generalization become more common for polymerizations which are carried out at higher\ntemperatures. Head-to\u2014head addition is also somewhat more abundant in the case of halogenated\nmonomers such as vinyl chloride. The preponderance of head-to\u2014tail additions is understood to\narise from a combination of resonance and steric effects. In many cases, the ionic or free-radical\nreaction center occurs at the substituted carbon due to the possibility of resonance stabilization or\nelectron delocalization through the substituent group. Head-to-tail attachment is also sterically\nfavored, since the substituent groups on successive repeat units are separated by a methylene\ncarbon. At higher polymerization temperatures, larger amounts of available thermal energy make\nthe less-favored states more accessible. In vinyl \ufb02uoride, no resonance stabilization is possible and\nsteric effects are minimal. This monomer adds primarily in the head\u2014to-tail orientation at low\ntemperatures and tends toward a random combination of both at higher temperatures. The styrene\nradical, by contrast, enjoys a large amount of resonance stabilization in the bulky phenyl group and\npolymerizes almost exclusively in the head-to\u2014tail mode. The following example illustrates how\nchemical methods can be used to measure the relative amounts of the two positional isomers in a\npolymer sample.\n"]], ["block_5", ["are positional isomerism, stereo isomerism, and geometrical isomerism. We shall focus attention\non synthetic polymers and shall, for the most part, be concerned with these types of isomerism\noccurring singly, rather than in combinations. Some synthetic and analytical aspects of stereo\nisomerism will be considered in Chapter 5. Our present concern is merely to introduce the\npossibilities of these isomers and some of the associated vocabulary.\n"]], ["block_6", ["Positional isomerism is conveniently illustrated by considering the polymerization of a vinyl\nmonomer. In such a reaction, the adding monomer may become attached to the growing chain\nend (designated by =k) in either of two orientations:\n"]], ["block_7", ["1,2-Glycol bonds are cleaved by reaction with periodate; hence poly(vinyl alcohol) chains are\nbroken at the site of head\u2014to-head links in the polymer. The fraction of head\u2014to-head linkages\nin poly(vinyl alcohol) may be determined by measuring the molecular weight before (subscript b)\nand\nafter\n(subscript\na)\ncleavage\nwith\nperiodate\naccording\nto\nthe\nfollowing\nformula:\nFraction =44(1/Ma\u20141/Mb). Derive this expression and calculate the value for the fraction in the\ncase of Mb: 105 and M3,: 103.\n"]], ["block_8", ["1.6\nStructural Isomerism\n"]], ["block_9", ["1.6.1\nPositional lsomerism\n"]], ["block_10", ["Example 1.3\n"]], ["block_11", ["H\nH\nX\nL<\n+\n>=<\n.._._...\n(1.13)\n"]], ["block_12", [{"image_2": "36_2.png", "coords": [56, 249, 171, 303], "fig_type": "molecule"}]], ["block_13", ["X\nH\nH\nX\n"]], ["block_14", [{"image_3": "36_3.png", "coords": [152, 218, 438, 324], "fig_type": "figure"}]], ["block_15", [{"image_4": "36_4.png", "coords": [167, 273, 239, 326], "fig_type": "molecule"}]], ["block_16", ["\u2014\nmy\nX\nX'\n(1.VI)\n"]], ["block_17", ["_\nx\nH\u201d\n(1.VII)\n"]], ["block_18", [{"image_5": "36_5.png", "coords": [176, 229, 243, 268], "fig_type": "molecule"}]]], "page_37": [["block_0", [{"image_0": "37_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Begin by recognizing containing x of the head-to-head links will be cleaved into\nx + 1 molecules upon reaction. Hence if n is the number of polymer molecules in a sample of mass\nw, the following relations apply before and after cleavage: na=(x+1)nb or w/Ma=(x+ 1)\n(w/Mb). Solving for x and dividing the latter by the total number of linkages in the original\npolymer gives the desired ratio. The total number of links in the original polymer is Mb/Mo.\nTherefore M0(1/Ma\u20141/Mb). For poly(vinyl alcohol) M0 is 44, so the desired\nformula has been obtained. For the specific data given, x/nb =44(10_3\u2014-10_5) 0.044, or about\n4% 0f the additions are in the less favorable orientation. We shall see presently that the molecular\nweight of a polymer is an average, which is different depending on the method used for its\ndetermination. The present example used molecular weights as a means for counting the number\nof molecules present. Hence the sort of average molecular weight used should also be one which is\nbased on counting.\n"]], ["block_2", ["Structure (1.VIII) and Structure (LIX) are not equivalent; they would not superimpose if the\nextended chains were overlaid. The difference has to do with the stereochemical con\ufb01guration at\nthe asymmetric carbon atoms. Note that the asymmetry is more accurately described as pseudoa-\nsymmetry, since two sections of chain are bonded to these centers. Except near chain ends, which\nwe ignore for high polymers, these chains provide local symmetry in the neighborhood of the\ncarbon under consideration. The designations of D and L or R and S are used to distinguish these\nstructures, even though true asymmetry is absent.\nWe use the word con\ufb01guration to describe the way the two isomers produced by Reaction (1.C)\ndiffer. It is only by breaking bonds, moving substituents, and reforming new bonds that the two\nstructures can be interconverted. This state of affairs is most readily seen when the molecules are\ndrawn as fully extended chains in one plane, and then examining the side of the chain on\nwhich substituents lie. The con\ufb01gurations are not altered if rotation is allowed to occur around\nthe various bonds of the backbone to change the shape of the molecule to a jumbled coil. We shall\nuse the term conformation to describe the latter possibilities for different molecular shapes.\nThe configuration is not in\ufb02uenced by conformational changes, but the stability of different\nconformations may be affected by differences in configuration. We shall return to these effects\nin Chapter 6.\nIn the absence of any external in\ufb02uence, such as a catalyst that is biased in favor of one\ncon\ufb01guration over the other, we might expect Structure (1.VIII) and Structure (LIX) to occur at\nrandom with equal probability as if the configuration at each successive addition were determined\nby the toss of a coin. Such indeed is the ordinary case. However, in the early 19503, stereospeci\ufb01c\ncatalysts were discovered; Ziegler and Natta received the Nobel Prize for this discovery in 1963.\nFollowing the advent of these catalysts, polymers with a remarkable degree of stereoregularity\nhave been formed. These have such a striking impact on polymer science that a substantial part of\n"]], ["block_3", [{"image_1": "37_1.png", "coords": [34, 278, 294, 380], "fig_type": "figure"}]], ["block_4", ["The second type of isomerism we discuss in this section is stereo isomerism. Again we consider the\nnumber of ways a singly substituted vinyl monomer can add to a growing polymer chain:\n"]], ["block_5", ["Structural \n21\n"]], ["block_6", ["Solution\n"]], ["block_7", ["1.6.2\nStereo Isomerism\n"]], ["block_8", ["X\na,\nH\nX\nc\nc\n"]], ["block_9", [".j\u201c\n+\n>=<\n\u2014'*\n(l.C)\n"]], ["block_10", ["_\n(l.VIII)\nH\nH\n"]], ["block_11", [{"image_2": "37_2.png", "coords": [153, 275, 345, 382], "fig_type": "figure"}]], ["block_12", ["I\n\"\n"]], ["block_13", ["(LIX)\n"]]], "page_38": [["block_0", [{"image_0": "38_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "38_1.png", "coords": [31, 133, 313, 197], "fig_type": "molecule"}]], ["block_2", ["Figure 1.3 shows sections of polymer chains of these three types; the substituent X equals\nphenyl for polystyrene and methyl for polypropylene. The general term for this stereoregularity\nis tacticity, a term derived from the Greek word meaning \u201cto put in order.\u201d Polymers of\ndifferent tacticity have quite different properties, especially in the solid state. As we will\nsee in Chapter 13, one of the requirements for polymer crystallinity is a high degree of\nmicrostructural regularity to enable the chains to pack in an orderly manner. Thus atactic\npolypropylene is a soft, tacky substance, whereas both isotactic and syndiotactic polypropylene\nare highly crystalline.\n"]], ["block_3", ["The \ufb01nal type of isomerism we take up in this section is nicely illustrated by the various possible\nstructures that result from the polymerization of 1,3-dienes. Three important monomers of this type\nare 1,3-butadiene, 1,3-isoprene, and 1,3-chloroprene, Structure (1.X) through Structure (1.XII),\nrespectively:\n"]], ["block_4", ["22\nIntroduction to Chain Molecules\n"]], ["block_5", ["Chapter 5 is devoted to a discussion of their preparation and characterization. For now, only the\nterminology involved in their description concerns us. Three different situations can be distin-\nguished along a chain containing pseudoasymmetric carbons:\n"]], ["block_6", ["1.\nIsotactic. All substituents lie on the same side of the extended chain. Alternatively, the\nstereoconfiguration at the asymmetric centers is the same, say, \u2014~DDDDDDDDD\u2014\u2014.\n2.\nSyndiotactic. Substituents on the fully extended chain lie on alternating sides of the backbone.\nThis alternation of configuration can be represented as \u2014DLDLDLDLDLDL\u2014~.\n"]], ["block_7", ["3.\nAtactic.\nSubstituents\nare\ndistributed\nat\nrandom\nalong\nthe\nchain,\nfor\nexample,\nDDLDLLLDLDLL\u2014.\n"]], ["block_8", ["Figure 1.3\nSections of \u201cpolyvinyl X\u201d chains of differing tacticity: (a) isotactic, (b) syndiotactic, and\n"]], ["block_9", ["(c) atactic.\n"]], ["block_10", ["1.6.3\nGeometrical lsomerism\n"]], ["block_11", [{"image_2": "38_2.png", "coords": [41, 619, 111, 671], "fig_type": "molecule"}]], ["block_12", ["HNH\n(LX)\n"]], ["block_13", [{"image_3": "38_3.png", "coords": [54, 218, 301, 284], "fig_type": "molecule"}]], ["block_14", ["H\n"]], ["block_15", ["d9\nd'\n:3\n.4\n\u20185\n\u20ac-\n\u20185\n\u20185\n:3\nf.\n15\n:\n'5\n\u20185\n1\n9..\n5\n\u00a2\n0\nc\n1'.\nc\n'o,\n\u20185\nz\n\u2018\n1\nd\n0\n"]], ["block_16", ["\u201c\u20180\n\\tw\u201c\nI:\nI:\n.1-\n:\u2014\n.-\n[51:\n"]], ["block_17", [".\u2014\nI:\nI;\n"]], ["block_18", ["I,\n[f-\n"]], ["block_19", [".\no\n"]]], "page_39": [["block_0", [{"image_0": "39_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["To illustrate the possible modes of polymerization of these compounds, consider the following\nreactions of \n"]], ["block_2", ["Structural \n23\n"]], ["block_3", ["1.\n1,2\u2014 and 3,4-Polymerizations. As far as the polymer chain backbone is concerned, these\ncompounds could just as well be mono\u2014olefins, since the second double bond is relegated to\nthe status of a substituent group. Because of the reactivity of the latter, however, it might\nbecome involved in cross-linking reactions. For isoprene, 1,2- and 3,4-polymerizations yield\ndifferent products:\n"]], ["block_4", [{"image_1": "39_1.png", "coords": [37, 248, 319, 321], "fig_type": "figure"}]], ["block_5", [{"image_2": "39_2.png", "coords": [39, 47, 106, 104], "fig_type": "molecule"}]], ["block_6", [{"image_3": "39_3.png", "coords": [43, 507, 119, 601], "fig_type": "molecule"}]], ["block_7", [{"image_4": "39_4.png", "coords": [43, 95, 110, 142], "fig_type": "molecule"}]], ["block_8", [{"image_5": "39_5.png", "coords": [44, 410, 244, 468], "fig_type": "molecule"}]], ["block_9", [{"image_6": "39_6.png", "coords": [47, 255, 214, 307], "fig_type": "molecule"}]], ["block_10", ["These differences do not arise from 1,2\u2014 or 3,4\u2014polymerization of butadiene. Structure (1.X111)\nand Structure (1.XIV) can each exhibit the three different types of tacticity, so a total of six\nstructures can result from this monomer when only one of the olefin groups is involved in the\nbackbone formation.\n1,4-Polymerization. This mode of polymerization gives a molecule with double bonds along\nthe backbone of the chain. Again using isoprene as the example,\n"]], ["block_11", ["As in all double-bond situations, the adjacent chain sections can be either cis or trans\u2014\nStructure (1.XV) and Structure (1.XV1), respectivelymwith respect to the double bond,\nproducing the following geometrical isomers:\n"]], ["block_12", ["Figure 1.4 shows several repeat units of cis\u2014l,4-polyisoprene and trans\u20141,4\u2014polyisoprene.\nNatural rubber is the cis isomer of 1,4\u2014polyisoprene and gutta\u2014percha is the trans isomer.\nPolymers of Chloroprene (Structure (1.X11)) are called neoprene and copolymers of butadiene and\nstyrene are called SBR, an acronym for styrene\u2014butadiene rubber. Both are used for many of the\nsame applications as natural rubber. Chloroprene displays the same assortment ofpossible isomers\n"]], ["block_13", ["Me\nH\nm\n(1 .XV)\n"]], ["block_14", ["Me\nMe\nH\nn\nn\nn\nHNH\n\u2014+\n\\\nH\nor\nMe \nH\n(LD)\n"]], ["block_15", ["n\nM\n"]], ["block_16", ["H\nH\nH\n"]], ["block_17", ["(LE)\nE3i\n"]], ["block_18", ["H\nn\n"]], ["block_19", ["H\n(1 .X1)\n"]], ["block_20", ["H\n(1 .XII)\n"]], ["block_21", [{"image_7": "39_7.png", "coords": [102, 423, 235, 459], "fig_type": "molecule"}]], ["block_22", [{"image_8": "39_8.png", "coords": [162, 246, 309, 335], "fig_type": "molecule"}]], ["block_23", ["(1 .X111)\n(1.X1V)\n"]], ["block_24", ["(1.XVI)\n"]]], "page_40": [["block_0", [{"image_0": "40_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "40_1.png", "coords": [22, 54, 341, 100], "fig_type": "molecule"}]], ["block_2", [{"image_2": "40_2.png", "coords": [25, 55, 153, 107], "fig_type": "molecule"}]], ["block_3", ["This example, as simplistic as it is, actually underscores two important points. First, polymer\nchemists have to get used to the idea that while all carbon atoms are identical, and all 1,3-butadiene\nmolecules are identical, polybutadiene actually refers to an effectively in\ufb01nite number of distinct\nchemical structures. Second, almost all synthetic polymers are heterogeneous in more than one\nvariable: molecular weight, certainly; isomer and tacticity distribution, probably; composition and\nsequence distribution, for copolymers; and branching structure, when applicable.\n"]], ["block_4", ["Almost every synthetic polymer sample contains molecules of various degrees of polymerization.\nWe describe this state of affairs by saying that the polymer shows polydispersity with respect to\nmolecular weight or degree of polymerization. To see how this comes about, we only need to think\nof the reactions between monomers that lead to the formation of polymers in the first place.\nRandom encounters between reactive species are responsible for chain growth, so statistical\n"]], ["block_5", ["as isoprene; the extra combinations afforded by copolymer composition and structure in SBR\noffset the fact that Structure (1 .XHI) and Structure (LXIV) are identical for butadiene.\n4.\nAlthough the conditions of the polymerization reactions may be chosen to optimize the\nformation of one specific isomer, it is typical in these systems to have at least some\ncontribution of all possible isomers in the polymeric product, except in the case of polymers\nof biological origin, like natural rubber and gutta-percha.\n"]], ["block_6", [". . - x 3 = 3N, where N is the degree of polymerization. (Recall that the combined probability of a\nsequence of events is equal to the products of the individual probabilities.) In this case N 54,000/\n54: 1000, and thus there are about 31000 m 10500 possible structures. Now we need to count\nhow many molecules we have. Assuming for simplicity that the tank car is 3.3 m\nX\n3.3 m x\n10 m 100 m3 108 cm3, and the density of the polymer is 1 g/cm3 (it is actually closer to 0.89\ng/cm3), we have 108 g of polymer. As M 54,000 g/mol, we have about 2000 moles, or 2000 X 6\nx\n1023 a 1027 molecules. Clearly, therefore, there is essentially no chance that any two molecules\nhave the identical structure, even without taking the molecular weight distribution into account.\n"]], ["block_7", ["We will not attempt to provide a precise answer to such an artificial question; what we really want\nto know is whether the probability is high (approximately 1), vanishing (approximately 0), or finite.\nFrom the discussion above, we recognize three geometrical isomers: trans-1,4, cis-1,4, and 1,2.\nWe will ignore the stereochemical possibilities associated with the 1,2 linkages. Assuming all three\nisomers occur with equal probability, the total number of possible structures is 3 X 3 x 3 x\n"]], ["block_8", ["Suppose you have just ordered a tank car of polybutadiene from your friendly rubber company. By\nsome miracle, all the polymers in the sample have M 254,000. The question we would like to\nconsider is this: what are the chances that any two molecules in this sample have exactly the same\nchemical structure?\n"]], ["block_9", ["24\nIntroduction to Chain Molecules\n"]], ["block_10", ["Solution\n"]], ["block_11", ["(a)\n(b)\n"]], ["block_12", ["1.7\nMolecular Weights and Molecular Weight Averages\n"]], ["block_13", ["Example 1.4\n"]], ["block_14", ["Figure 1.4\n1,4-Polyisoprene (a) all-cis isomer (natural rubber) and (b) all\u2014trans isomer (gutta-percha).\n"]], ["block_15", [{"image_3": "40_3.png", "coords": [174, 57, 325, 96], "fig_type": "molecule"}]]], "page_41": [["block_0", [{"image_0": "41_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The probability x, is the number fraction or mole fraction of i-mer. We can use this quantity to\ndefine a particular average molecular weight, called the number-average molecular weight, Mn.\nWe do this by multiplying the probability of \ufb01nding an i-mer with its associated molecular weight,\nx,M,-, and adding all these up:\n"]], ["block_2", ["descriptions are appropriate for the resulting product. The situation is reminiscent of the distribu-\n"]], ["block_3", ["tion of molecular \nenergy to some reducing the energy of others. Therefore, we talk about the\nmolecular weight of a polymer, we mean some characteristic average molecular weight. It turns out\nthere are several distinct averages that may be defined, and that may be measured experimentally; it\nis therefore spend time on this topic. Furthermore, one might well \nsamples ofin terms of one kind of \nterms of another; this, in turn, can lead to the situation where the two polymers behave identically\nin terms of some important properties, but differently in terms of others.\nIn Chapter 2 through Chapter 4 we shall examine the expected distribution of molecular weights\nfor condensation and addition polymerizations in some detail. For the present, our only concern is\nhow such distribution of molecular weights is described. We will define the most commonly\nencountered averages, and how they relate to the distribution as a whole. We will also relate them\nto the standard parameters used for characterizing a distribution: the mean and standard deviation.\nAlthough these students are familiar with them only results\nprovided by a calculator, and so we will describe them in some detail.\n"]], ["block_4", ["n,- (we could also refer to n,- as the number of moles of i-mer, but again this just involves a factor of\nAvogadro\u2019s number). The first question we ask is this: if we choose a molecule at random from our\nsample, what is the probability of obtaining an i\u2014mer? The answer is straightforward. The total\nnumber of molecules is 2, m, and thus this probability is given by\n"]], ["block_5", ["Molecular Weights Averages\n25\n"]], ["block_6", ["The other expressions on the right-hand side of Equation 1.7.2 are equivalent, and will prove useful\nsubsequently. You should convince yourself that this particular average is the one you are familiar\nwith in everyday life: take the value of the property of interest, M,- in this case, add it up for all the\n(m) objects that possess that value of the property, and divide by the total number of objects.\nSo far, so good. We return for a moment to our hypothetical sample, but instead of choosing\n"]], ["block_7", ["Suppose we have a polymer sample containing many molecules with a variety of degrees of\npolymerization. We will call a molecule with degree of polymerization i an \u201ci-mer\u201d, and the\nassociated molecular weight MiziMo, where M0 is the molecular weight of the repeat unit.\n(Conversion between a discussion couched in terms of i or in terms of M,- is therefore straightfor\u2014\nward, and we will switch back and forth when convenient.) The number of i-mers we will denote as\n"]], ["block_8", ["a molecule at random, we choose a repeat unit or monomer at random, and ask about the molecular\nweight of the molecule to which it belongs. We will get a different answer, as a simple argument\nillustrates. Suppose we had two molecules, one a 10-mer and another a 20\u2014mer. If we choose\nmolecules at random, we would choose each one 50% of the time. However, if we choose mono-\nmers at random, 2/3 of the monomers are in the 20-mer, so we would pick the larger molecule\ntwice as often as the smaller. The total number of monomers in a sample is 2,- in;, and the chance of\n"]], ["block_9", ["1.7.1\nNumber-, Weight-. and z-Average Molecular Weights\n"]], ["block_10", [{"image_1": "41_1.png", "coords": [42, 489, 216, 526], "fig_type": "molecule"}]], ["block_11", [{"image_2": "41_2.png", "coords": [46, 397, 91, 429], "fig_type": "molecule"}]], ["block_12", ["x,- \n(1.7.1)\n\u201d1\nZr \u201d1'\n"]], ["block_13", ["M \u2014:x,-M- =Z\u2014Z':\u2014:I:4:M\nMO%\u2014:\u201c\n(1.7.2)\n"]]], "page_42": [["block_0", [{"image_0": "42_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Although M2 is not directly related to a simple fraction like x, or w,, it does have some experimental\nrelevance. We could continue this process indefinitely, just by incrementing the power of i by one in\nboth numerator and denominator of Equation 1.7.5, but it will turn out that there is no real need to do\nso. However, there is a direct relationship between this process and something well-known in\nstatistical probability, namely the moments of a distribution, as we will see in the next section.\n"]], ["block_2", ["Of course, mass\u2014average would be the preferred descriptor, but it is not in common usage.\nQualitatively, we can say that Mn is the Characteristic average molecular weight of the sample\nwhen the number of molecules is the crucial factor, whereas MW is the characteristic average\nmolecular weight when the size of each molecule is the important feature. Although knowledge of\nMW and Mn is not suf\ufb01cient to provide all the information about a polydisperse system, these two\naverages are by far the most important and most commonly encountered, as we shall see\nthroughout the book.\nComparison of the last expressions in Equation 1.7.2 and Equation 1.7.4 suggests a trend; we\ncan define a new average by multiplying the summation terms in the numerator and denominator\nby t'. The so-called z-average molecular weight, M2, is constructed in just such a way:\n"]], ["block_3", ["The PDI is always greater than 1, unless the sample consists of exactly one value of M, in which case\nthe PDI :1; such a sample is said to be monodisperse. We will see in Chapter 2 and Chapter 3 that\ntypical polymerization schemes are expected to give PDIs near 2, at least in the absence of various\nside reactions; in industrial practice, such side reactions often lead to PDIs as large as 10 or more. In\nChapter 4, in contrast, so-ealled living polymerizations give rise to PDIs of 1.1 or smaller. Thus,\ndistributions for which the PDI < 1.5 are said to be \u201cnarrow,\u201d whereas those for which PDI > 2 are\nsaid to be \u201cbroad,\u201d of course, such designations are highly subjective. As a very simple illustration,\nthe two\u2014molecule example given in the previous section consisting of a lO-mer and a 20\u2014mer has a\nnumber-average degree of polymerization of 15 and a weight average of 16.7; thus its PDI 1.11,\nwhich in polymer terms would be considered \u201cnarrow.\u201d This trivial example actually underscores\nan important point to bear in mind: polymer samples with \u201cnarrow\u201d distributions will still contain\nmolecules that are quite different in size (see Problem 1.8 for another instance).\n"]], ["block_4", ["Although the values MW or MH tell us something useful about a polymer sample, individually they\ndo not provide information about the breadth of the distribution. However, the ratio of the two\nturns out to be extremely useful in this regard, and it is given a special name, the polydispersity\nindex (PDI) or just the polydispersity:\n"]], ["block_5", ["Accordingly, we define the weight\u2014average molecular weight of the sample, MW, by\n"]], ["block_6", ["picking a particular i\u2014mer will be determined by the product 171,-. The resulting ratio is, in fact, the\nweight fraction. or mass fraction of i-mer in the sample, w,\u2014:\n"]], ["block_7", ["26\nIntroduction to Chain Molecules\n"]], ["block_8", ["1.7.2\nPolydispersity Index and Standard Deviation\n"]], ["block_9", [{"image_1": "42_1.png", "coords": [40, 297, 120, 342], "fig_type": "molecule"}]], ["block_10", ["Z; 53\u201d!\u201c\nM2 =M0 2 {212,-\n(1.7.5)\n"]], ["block_11", ["Mw\nMn\nPDI \n(1.7.6)\n"]], ["block_12", ["(1.7.3)\nWf=\n"]], ["block_13", [{"image_2": "42_2.png", "coords": [47, 139, 280, 171], "fig_type": "molecule"}]], ["block_14", ["2\n\u2014Z_WiM 2,a\nZ \u201d\u2018M\u2019\nMow:\nI \"\u2019\n(1.7.4)\n"]], ["block_15", ["23-171;\n"]], ["block_16", ["in,-\n"]], ["block_17", ["ZZZ in,\nZ, ngM-z M02. in,-\n"]]], "page_43": [["block_0", [{"image_0": "43_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "43_1.png", "coords": [26, 199, 259, 252], "fig_type": "molecule"}]], ["block_2", [{"image_2": "43_2.png", "coords": [28, 551, 326, 623], "fig_type": "figure"}]], ["block_3", [{"image_3": "43_3.png", "coords": [32, 110, 202, 147], "fig_type": "molecule"}]], ["block_4", ["deviation. We will now develop relationshipsbetween these quantities and M\u201c, in\n"]], ["block_5", ["so doing justify useful of the breadth \nmean of any distribution (i), is de\ufb01ned as\n"]], ["block_6", ["where both discrete (x,) continuous (P(i)) versions are considered. From this definition, and\nEquation 1.7.2, more than the number-average \nization, thus Mn weight.\nThe standard deviation 0' quantifies the width of the distribution. It is defined as\n"]], ["block_7", ["Note that a2 has the significance of being the mean of the square of the deviations of\nindividual values from the mean. Accordingly, a is sometimes called the root mean square (rms)\ndeviation.\nFrom a computational point of view, the standard deviation may be written in a more conveni-\nent form by carrying out the following operations. First both sides of Equation 1.7.8 are squared,\nand then the difference i\u2014(i) is squared to give\n"]], ["block_8", [{"image_4": "43_4.png", "coords": [35, 454, 134, 476], "fig_type": "molecule"}]], ["block_9", ["Recalling the definition of the mean, we recognize the first term on the right\u2014hand side of Equation\n1.7.9 to be the mean value of i2 and write\n"]], ["block_10", ["Molecular \n27\n"]], ["block_11", ["where in this case the standard deviation will have the units of molecular weight. If we expand\nEquation 1.7.12 we find\n"]], ["block_12", ["It is, of course, important to realize that (1'2) # (1)2. An alternative to Equation 1.7.8 as a definition\nof standard deviation is, therefore,\n"]], ["block_13", ["We can factor out Mn Z,- x,M,- to obtain\n"]], ["block_14", ["Similarly, the standard deviation can be written\n"]], ["block_15", [{"image_5": "43_5.png", "coords": [35, 498, 171, 532], "fig_type": "molecule"}]], ["block_16", [{"image_6": "43_6.png", "coords": [39, 395, 214, 424], "fig_type": "molecule"}]], ["block_17", [{"image_7": "43_7.png", "coords": [39, 328, 191, 362], "fig_type": "molecule"}]], ["block_18", [{"image_8": "43_8.png", "coords": [40, 626, 172, 668], "fig_type": "molecule"}]], ["block_19", ["In most fields of science, distributions are generally characterized by a mean and a standard\n"]], ["block_20", ["1'\u2014\n1\n2 \u201d2\n\u201d2\na E\n(Eingnf\n\u00bb\n)\n:\n(Xxx;\n_\n"]], ["block_21", [".72 (i2) 2(7)2 + (7)2 (72) (7)2\n(1.7.10)\n"]], ["block_22", ["a 2 [Z x,(M,.2 214,114,, + M3,)\n"]], ["block_23", ["a 2 ((1-2) _ (ml/2\n(1.7.11)\n"]], ["block_24", ["0' : [Zia-(M, 114,02] \n(1.7.12)\n"]], ["block_25", ["(i) 2\nix,- 3 J: iP(i)di\n(1.7.7)\n"]], ["block_26", ["2 _ 2,71,?\n. Zinii\n.2\n"]], ["block_27", ["'\n2\n1/2\n111M\n1]\n(1,-1.1...\n(Zl\u2018xiMi)\n"]], ["block_28", ["1/2\n1/2\n:\n[(Zxr\u2018M?)\na\nMg]\n(1.7.13)\n"]], ["block_29", ["(0)2)\n(1.7.8)\n"]]], "page_44": [["block_0", [{"image_0": "44_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "44_1.png", "coords": [11, 539, 453, 673], "fig_type": "figure"}]], ["block_2", [{"image_2": "44_2.png", "coords": [32, 67, 165, 112], "fig_type": "molecule"}]], ["block_3", ["This result shows that the square root of the amount by which the ratio MW/Mn exceeds unity equals\nthe standard deviation of the distribution relative to the number-average molecular weight. Thus if\n"]], ["block_4", ["MW/Mn 1.50, then the standard deviation is 71% of the value of Mn. This shows that reporting the\nmean and standard deviation of a distribution or the values of Mn and MW/Mn gives equivalent\ninformation.\nWe can define the quantities known as moments of a distribution, again either in the discrete or\ncontinuous forms. The k-th moment 11;, is given by\n"]], ["block_5", ["150130) di 1. From Equation 1.7.17 we can also see that the mean is equivalent to the first moment.\nFrom Equation 1.7.4 and Equation 1.7.5 it is apparent that MW and Mz are proportional to the ratio of\nthe second to the \ufb01rst moment and the third to the second, respectively. More generally, moments\n"]], ["block_6", ["The first and second columns of Table 1.4 give the number of moles of polymer in six different\nmolecular weight fractions. Calculate Mw/Mn for this polymer and evaluate 0- using both Equation\n1.7.12 and Equation 1.7.16.\n"]], ["block_7", ["Table 1.4\nSome Molecular Weight Data for a Hypothetical Polymer Used in Example 1.5\n"]], ["block_8", [{"image_3": "44_3.png", "coords": [34, 258, 169, 292], "fig_type": "molecule"}]], ["block_9", ["and, finally, by recognizing from Equation 1.7.4 that\n"]], ["block_10", ["we reach\n"]], ["block_11", ["a distribution is characterized by Mn 10,000 and (r 3000, then MW/Mn 1.09. Alternatively, if\n"]], ["block_12", ["In this definition both x,- and P(i) are normalized distributions, which mean that 2, x,- and\n"]], ["block_13", ["can be referred to a particular value, such as the k\u2014th moment about the mean, 121,:\n"]], ["block_14", ["2:0.049\n2:734\n2: 113\n2:70\n2:28.10\n"]], ["block_15", ["28\nIntroduction to Chain Molecules\n"]], ["block_16", ["First, consider the following numerical example in which we apply some of the equations of this\nsection to hypothetical data.\n"]], ["block_17", ["0.003\n10,000\n30\n3.0\n25\n7.50\n"]], ["block_18", ["0.008\n12,000\n96\n1 1.5\n9\n7.20\n"]], ["block_19", ["0.011\n14,000\n154\n21.6\n1\n1.10\n"]], ["block_20", ["0.009\n18,000\n162\n29.2\n9\n8.10\n"]], ["block_21", ["0.001\n20,000\n20\n4.0\n25\n2.50\n"]], ["block_22", ["From this expression we see that 02 is the second moment about the mean.\n"]], ["block_23", ["0.017\n16,000\n272\n43.5\n1\n1.70\n"]], ["block_24", ["24. (mol)\nM1 (g/mol)\nm.- (g)\n(gz/mol)\n(g\u2018Z/molz)\n(nmol)\n"]], ["block_25", ["Example 1.5\n"]], ["block_26", ["1.7.3\nExamples of Distributions\n"]], ["block_27", [{"image_4": "44_4.png", "coords": [44, 358, 133, 391], "fig_type": "molecule"}]], ["block_28", ["MW\n1/2\n=Mn\n_1\n.7.1\n\u20187\n[Mn\ni\n(1\n6)\n"]], ["block_29", ["00\n,1, 2m\u201c \n(1.7.17)\n"]], ["block_30", ["m. 2w >t\n(1.7.13)\n"]], ["block_31", [".\n.2\n.\n.\n.2\n"]], ["block_32", [{"image_5": "44_5.png", "coords": [129, 573, 445, 658], "fig_type": "figure"}]], ["block_33", ["\u201d1,114,. x 10\u20185\n(Mt\u2014Mn)2 x 10-6\nn,(M,-\u2014Mn)2 x 10\u20144\n"]]], "page_45": [["block_0", [{"image_0": "45_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "45_1.png", "coords": [21, 412, 194, 446], "fig_type": "molecule"}]], ["block_2", ["The proof of Equation 1.7.21 is left to Problem 1.9 at the end of the chapter. Now we can\nuse Equation 1.7.19 and Equation 1.7.20 to generate distributions for specified values of Mn and\nMw/Mn.\nFigure 1.5a shows the mole fraction and weight fraction as a function of M, with the particular\nchoice of Mn: 10,000, M0: 100, and 2: 1. Thus, according to Equation 1.7.21, the PDI=2\nand Mw=20,000. Both MW and MD are indicated by vertical lines on the plot. There are\nseveral remarkable features to point out. First, x, is a continuously decreasing function of\nM (and therefore i). We shall see in Chapter 2 that this is to be expected in step-growth\npolymerizations. It means, for example, that there are more unreacted monomers (i 1) than\nany other particular i-mer. The weight fraction, however, has a distinct but broad maximum.\nNotice also how many different values of M are present in significant amounts. For example, there\nis certainly a significant mass of the sample that is five times smaller than MW, or five times larger\nthan Mn.\n"]], ["block_3", ["Evaluate the product m, for each class; this is required for the calculation of both Mn and\nMW. Values of quantity are listed in the third column of Table 1.4. From 2,111,114,- and 2,11,,\nMn 734/0049 15,000. (The matter of significant figures will not be strictly adhered to in this\nexample. As a general rule, one has to work pretty hard to obtain more than two significant \ufb01gures\nin an experimental determination of M.)\nThe products miM, are mass-weighted contributions and are listed in the fourth column of\nTable 1.4. From 2.1%,. and 2,11,14,14,, 113 x 105/734 15,400.\nThe ratio MW/Mn is found to be 15,400/15,000 1.026 for these data. Using Equation 1.7.16, we\nhave a/a (1.026 1)\u201d2 0.162 or 0': 0.162(15,000) 2430.\nTo evaluate 0' via Equation 1.7.12, differences between M,- and Mn must be considered. The fifth\nand sixth columns in Table 1.4 list (M, Mn)2 and 11,-(M, Mn)2 for each class of data. From 2,11,-\nand 2,n,(M, Mn)2, 02 28.1 x 104/0049 5.73 x 106 and 0' 2390.\nThe discrepancy between the two values of 0' is not meaningful in terms of significant figures;\nthe standard deviation is 2400.\nAs polymers go, this is a very narrow molecular weight distribution.\n"]], ["block_4", ["The utility of this distribution arises in part because of a very simple relationship between the\nparameter 2 and the polydispersity:\n"]], ["block_5", ["When we consider particular polymerization schemes in Chapter 2 through Chapter 4, we\nwill derive explicit expressions for the expected distributions x,- and 142,. For now, however, let us\nconsider a particular mathematical function known as the Schulz\u2014Zimm distribution. It has the virtue\nthat by varying a single parameter, 2, it is possible to obtain reasonable descriptions for typical\nnarrow or moderately broad samples. We will use it to illustrate graphically how the distribution\nmight appear for a given polydispersity. The Schulz\u2014Zimm distribution can be expressed as\n"]], ["block_6", ["where Hz 1) is the so\u2014called gamma function (which is tabulated in many mathematical\nreferences). For integer values ofz, F(z+ 1) 2!, where z! (z factorial) z X (z 1) x (z 2) x x 1.\nFrom Equation 1.7.2 and Equation 1.7.3 we can see that w,- = x,M,/Mn, and thus\n"]], ["block_7", ["Molecular Weights and Molecular Weight Averages\n29\n"]], ["block_8", ["Solution\n"]], ["block_9", [{"image_2": "45_2.png", "coords": [40, 336, 209, 370], "fig_type": "molecule"}]], ["block_10", ["Mw_z+1\n"]], ["block_11", ["P(M\n\u2014\u2014\n22\u201c\nM\u2018H\nZM\u201d\n1719\n"]], ["block_12", ["1 \n--\u2014-\u2014-\u2014---\u2014--\u2014\u2014-\u2014\u2014\u2014\u2014\u2014\nI\n\u2014\nI\n10702\nW\nF(z+ 1) Mg\u201c exP<\nMn)\n(\nO)\n"]], ["block_13", ["Mn\n2\n(1.7.21)\n"]], ["block_14", ["zz+1\nM?\n2M.\n"]]], "page_46": [["block_0", [{"image_0": "46_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "46_1.png", "coords": [23, 118, 294, 394], "fig_type": "figure"}]], ["block_2", [{"image_2": "46_2.png", "coords": [31, 395, 277, 594], "fig_type": "figure"}]], ["block_3", [{"image_3": "46_3.png", "coords": [33, 147, 254, 315], "fig_type": "figure"}]], ["block_4", ["0.002\n"]], ["block_5", ["30\nIntroduction to Chain Molecules\n"]], ["block_6", ["0.004 \n_\nz\u2014 1\nj\n"]], ["block_7", ["Figure 1.5b shows the analogous curves, but now with PDI =4 and 220.333. Although the\nmole fraction looks superficially similar to the previous case, the mass distribution is exceedingly\nbroad, with a long tail on the high M side of Mn. (Note that the coincidence between Mn and the\npeak in w,- is a feature of this distribution, but not a general result in polymers; see Problem 1.10.)\n"]], ["block_8", ["Figure 1.5\nNumber and weight distributions for the Schulz\u2014Zimm distribution with the indicated poly-\ndispersities, and Mn 2 10,000.\n"]], ["block_9", ["0.0083 Mn\nMw\n0\nx,\nJ\nL\u2018\n_\ni\n_\n0\nWI\ni\n0006\nl\u2014\n\u2014\n__\n"]], ["block_10", ["0.008\n"]], ["block_11", ["0\n4x104\n8x104\n0))\nM\n"]], ["block_12", ["0.004 \n"]], ["block_13", ["l\n\u2018\u2014\u2018\n7\nD\n0\n2x104\n4x104\n5x10\u201cl\n8x104\n1x105\n"]], ["block_14", ["\u2014\n1\n"]], ["block_15", ["L\ni\n"]], ["block_16", [".\n4\n"]], ["block_17", [{"image_4": "46_4.png", "coords": [179, 423, 234, 483], "fig_type": "figure"}]], ["block_18", [{"image_5": "46_5.png", "coords": [180, 173, 234, 230], "fig_type": "figure"}]], ["block_19", [{"image_6": "46_6.png", "coords": [195, 525, 244, 563], "fig_type": "molecule"}]], ["block_20", ["I_I\n"]], ["block_21", ["_l_l_l_\n"]]], "page_47": [["block_0", [{"image_0": "47_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "47_1.png", "coords": [16, 45, 324, 330], "fig_type": "figure"}]], ["block_2", [{"image_2": "47_2.png", "coords": [32, 67, 265, 226], "fig_type": "figure"}]], ["block_3", ["Figure 1.5c shows the opposite extreme, or a narrow distribution with z :10 and PDI :1.11. Here\nboth x,- and w,- are much narrower, and have distinct maxima that are also close to Mlrl and MW,\nrespectively. (Note also that the ordinate scale has been truncated.) Even though this is a narrow\ndistribution by polymer standards, there are still significant numbers of molecules that are 50%\nlarger and 50% smaller than the mean.\n"]], ["block_4", ["The measurement of molecular weight is clearly the most important step in characterizing a\npolymer sample, although the previous sections have introduced many other aspects of polymer\nstructure necessary for a full analysis. A rich variety of experimental techniques have been\ndeveloped and employed for this purpose, and the major ones are listed in Table 1.5. As indicated\nin the rightmost column, we will describe end group analysis and matrix-assisted laser desorption/\nionization (MALDI) mass spectrometry in this section, while deferring treatment of size exclu\u2014\nsion chromatography (SEC), osmotic pressure, light scattering, and intrinsic viscometry to\nsubsequent chapters. Other techniques, such as those involving sedimentation, we will omit\nentirely. Before considering any technique in detail, some general comments about the entries in\nTable 1.5 are in order.\nThe first two techniques listed, SEC and MALDI, can provide information on the entire\ndistribution of molecular weights. To the extent that either one can do this reliably, accurately,\nand conveniently, there is a diminished need for any other approach. Over the past 30 years, SEC\nhas unquestionably emerged as the dominant method. Automated analysis of a few milligrams of\nsample dissolved in a good solvent can be achieved in half an hour. It is hard to imagine any\nserious polymer laboratory that does not have SEC capability. Nevertheless, as will be discussed in\n"]], ["block_5", ["0.008 \n00\n.0\n\u2014\n"]], ["block_6", ["J\n0004 \n.0\n0.2),\nPD|=1.11\nd\n"]], ["block_7", ["Measurement of Molecular Weight\n31\n"]], ["block_8", ["(C)\nM\n"]], ["block_9", ["Figure 1.5 (continued)\n"]], ["block_10", ["1.8\nMeasurement of Molecular Weight\n"]], ["block_11", ["1.8.1\nGeneral Considerations\n"]], ["block_12", ["OJ\n0\n1x104\n2x104\n3x104\n4x104\n"]], ["block_13", ["r\u2014\n00\n00\n\ufb02\n\u2014\n"]], ["block_14", ["_\no\noo\n_\n_\n.0\n\u20190\n_\n_\n.0\n0%\nO\nW,\u201c\n_\n.0\n.0\n_\n00\n_\n"]], ["block_15", ["_I\n"]], ["block_16", [":\n3%\n:\n"]], ["block_17", ["\u201c\n08\n0%\no\nX,\n-\n"]], ["block_18", ["_\n.0\n.0\n_\n"]], ["block_19", ["_\n00\n_\n00\noo\n_\no\n.0\n\"\n"]], ["block_20", ["_\n"]], ["block_21", [".0\n.0\n_\n_\no\n_\n"]], ["block_22", ["_\n.0\n:2:\n"]], ["block_23", ["_\n.0\n.00\n_\no\n\u2014\n'o\n'.o\n2:10\n-\n"]], ["block_24", ["_\n.0\ng???\n"]], ["block_25", ["..\n_\n. o\n_\n"]], ["block_26", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI-\n"]], ["block_27", [{"image_3": "47_3.png", "coords": [188, 89, 259, 159], "fig_type": "figure"}]]], "page_48": [["block_0", [{"image_0": "48_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "48_1.png", "coords": [17, 64, 431, 253], "fig_type": "figure"}]], ["block_2", ["Chapter 9, SEC has some serious limitations; one of these is a lack of resolution. Resolution in this\ncontext refers to the difference in M values that can be determined; SEC would struggle to help you\ndecide whether your sample was narrowly distributed with Mn 2 50,000, or was bimodal with peak\nmolecular weights at 40,000 and 60,000. In contrast, resolution is the real strength of MALDI, as\nwe will see below. MALDI is a relative newcomer among the techniques listed in Table 1.5, but it\nwill undoubtedly grow in importance as its scope expands.\nThe next group of techniques provides measurements of Mn. They do so by being sensitive to\nthe number of solute molecules in solution, as is inherent in the so-called colligative properties\n(osmotic pressure, freezing point depression, boiling point elevation, etc.). Of these, osmotic\npressure is the most commonly employed. It has the virtue (shared with light scattering) of\nbeing a technique based on equilibrium thermodynamics that can provide an absolute measurement\nwithout resorting to calibration against other polymer samples. End group analysis, to be discussed\nbelow, includes any of a number of analytical tools that can be used to quantify the presence of the\nunique structure of the polymer chain ends. For a linear chain, with two and only two ends,\ncounting the number of ends is equivalent to counting the number of molecules.\nLight scattering is suf\ufb01ciently important that it merits a full chapter; determination of MW is\nonly one facet of the information that can be obtained. Sedimentation experiments will not be\ndiscussed further in this text, although they play a crucial role for biOpolymer analysis in general,\nand proteins in particular. Similarly, gel electrophoresis (not listed in Table 1.5 or discussed further\nhere) is an analytical method of central importance in biological sciences, and especially for the\nseparation and sequencing of DNA in the Human Genome Project.\nThe intrinsic viscosity approach holds a place of particular historical importance; in the\ndays before routine use of SEC (up to about 1970) it was by far the easiest way to obtain\nmolecular weight information. The viscosity average molecular weight defined in Table 1.5 is not\na simple moment of the distribution, but involves the Mark\u2014Houwink exponent a, which needs to be\nknown based on other information. As 0.5 g a S 0.8 for most \ufb02exible polymers in solution, Mn 3 MV\n5 MW. The relation between the viscosity of a dilute polymer solution and the molecular weight of the\npolymer actually rests on some rather subtle hydrodynamics, as we will explore in Chapter 9.\n"]], ["block_3", ["As indicated previously, the end groups of polymers are inherently different in chemical structure\nfrom the repeat units of the chain, and thus provide a possible means of counting the number of\n"]], ["block_4", ["Z. \"5114 ,2\n.\n.\nMW\n'\nLight scattering\n8\na'M\u2018\nSedimentation velocity\n\u2014\n"]], ["block_5", ["2.31504?\n.\n.\n.\n.\n.\nMz\n2\nM2\nSedimentation equilibrium\n~\u2014\n"]], ["block_6", ["\"I\nl+a\n1/0\nMv\n(Li\u20149:7)\nIntrinsic viscosity\n9.3\n"]], ["block_7", ["Table 1.5\nSummary of the Molecular Weight Averages Most Widely Encountered in Polymer Chemistry\n"]], ["block_8", ["Zr :1e\n-\nMn\n081110t pressure\n7.4\n2\u201d\"\nOther colligative properties\n\u2014\n"]], ["block_9", ["1.8.2\nEnd Group Analysis\n"]], ["block_10", ["Information\nDefinition\nMethods\nSections\n"]], ["block_11", ["Full distribution\nx,, w,-\nSize exclusion chromatography\n9.8\nMALDI mass spectrometry\n1.8\n"]], ["block_12", ["32\nIntroduction to Chain Molecules\n"]], ["block_13", [{"image_2": "48_2.png", "coords": [137, 83, 326, 248], "fig_type": "figure"}]], ["block_14", [".n;\nr\n"]], ["block_15", [{"image_3": "48_3.png", "coords": [222, 80, 430, 236], "fig_type": "figure"}]], ["block_16", ["End group analysis\n1.8\n"]]], "page_49": [["block_0", [{"image_0": "49_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "49_1.png", "coords": [28, 617, 232, 678], "fig_type": "figure"}]], ["block_2", ["As an example, consider condensation polymers such as polyesters and polyamides. They are\nespecially well suited to this molecular weight determination, because they tend to have lower\nmolecular weights than addition polymers, and because they naturally have unreacted functional\ngroups at each end. Using polyamides as an example, we can readily account for the following\npossibilities:\n"]], ["block_3", ["molecules in a sample. Any analytical technique that can reliably quantify the concentration of end\ngroups can be used in this manner, and over the years many have been so employed.\nIt should section that will\nmeasure of Mn. The experiment willinvolve preparing\n"]], ["block_4", ["In this case, there is one functional group of each kind per molecule and could be detected for\nexample by titration with a strong base (for \u2014COOH) or strong acid (for \u2014NH2).\n2.\nIf a polyamide is prepared in the presence of a large excess of diamine, the average chain will\nbe capped by an amino group at each end:\n"]], ["block_5", ["In this case, only the amine can be titrated, and two ends are counted per molecule.\n3.\nSimilarly, if a polyamide is prepared in the presence of a large excess of dicarboxylic acid,\nthen the average chain will have a carboxyl group at each end:\n"]], ["block_6", ["a known mass of sample, probably in solution, which given M0 corresponds to a certain number\nof repeat units. The number of end groups is directly proportional to the number of polymers\nand the ratio of the number to the number of polymers is the number-average degree\nof polymerization.\nSeveral general principles apply to end group analysis:\n"]], ["block_7", ["1.\nThe chemical structure of the end group must be sufficiently different from that of the repeat\nunit for the chosen analytical technique to resolve the two clearly.\n2.\nThere must be a well-defined number of end groups per polymer, at least on average. For a linear\npolymer, there will be two and only two end groups per molecule, which may or may not be\ndistinct from each other. For branched polymers, the relation of the number of end groups to the\nnumber of polymers is ambiguous, unless the total number of branching points is also known.\n"]], ["block_8", ["3.\nThe technique is limited to relatively low molecular weights, as the end groups become more\nand more dilute as N increases. This is an obvious corollary of the fact that we can ignore end\ngroups in considering the structure of high molecular weight chains. How low is low in this\ncontext? The answer will depend on the particular system and analytical technique, but as a\nrule of thumb end groups present at the 1% level (corresponding to degrees of polymerization\nof 100 for a single end group, 200 for both end groups) can be reliably determined; those at the\n0.1% level cannot.\n"]], ["block_9", ["Measurement of \n33\n"]], ["block_10", ["I.\nA linear molecule has a carboxyl group at one end and an amino group at the other, such as\npoly(a-caprolactam):\n"]], ["block_11", [{"image_2": "49_2.png", "coords": [37, 615, 187, 676], "fig_type": "molecule"}]], ["block_12", [{"image_3": "49_3.png", "coords": [39, 424, 158, 471], "fig_type": "molecule"}]], ["block_13", [{"image_4": "49_4.png", "coords": [43, 524, 179, 567], "fig_type": "molecule"}]], ["block_14", ["H\nH\nHO\nFl\"\nN,\nI.N\nFl'\nOH\nv {\ufb01r\nR rd v\n0\n0\nOn\n0\n"]], ["block_15", ["3. ii\nRI\n/Rl\n/\nN\nFl\"\nN\nNH\nH2N\n{H\nH\n)1\n2\n"]], ["block_16", ["O\n0\nHOAm\n5 \u201dMn\n"]]], "page_50": [["block_0", [{"image_0": "50_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Peak (c) corresponds to the methylene protons adjacent to the hydroxyl end group; there are\ntwo such protons per polymer. Peak (a) re\ufb02ects the single olefinic proton per 1,4 repeat unit,\nwhereas peak (b) shows the two vinyl protons per 3,4 repeat unit. If we represent the integration of\npeak i as 1,, then the degree of polymerization is pr0portional to (Iat + 113/2) 26.9 + 2.61 29.5.\nThe number of polymers is proportional to 1J2 1. Thus the number-average degree of poly-\nmerization is 29.5, which gives an Mn=29.5\nx\n68 =2,000. Peak ((1) indicates the six methyl\nprotons on the initiator fragment. The peak integration Id 5.95 should be 31., which is within\nexperimental error. For this particular molecular structure, therefore, either end group could be\nused. An additional conclusion is that essentially 100% of the polymers were terminated with a\nhydroxyl group.\n"]], ["block_2", ["In this case, the acid group can be titrated, and again two ends should be counted per molecule.\nThe preceding discussion illustrated how classical acid\u2014base titration could be used for\nmolecular weight determination. In current practice, nuclear magnetic resonance (NNIR) spectroscopy\nis probably the most commonly used analytical method for end group analysis, especially proton\n(1H) NMR. An additional advantage of this approach is the possibility of obtaining further\ninformation about the polymer structure from the same measurement, as illustrated in the following\nexample.\n"]], ["block_3", ["The 1H NMR spectrum in Figure 1.6 corresponds to a sample of polyisoprene containing a sec-\nbutyl initiating group and a hydroxyl terminating end group. The relative peak integrations are (a)\n"]], ["block_4", ["34\nIntroduction to Chain Molecules\n"]], ["block_5", ["26.9, (b) 5.22, (c) 2.00, and (d) 5.95. What is Mn for this polymer? What is the relative percentage\nof 1,4 and 3,4 addition?\n"]], ["block_6", ["Figure 1.6\nlH NMR spectrum of a polyisoprene sample, discussed in Example 1.6. (Data courtesy of\nN. Lynd and MA. Hillmyer.)\n"]], ["block_7", ["Solution\n"]], ["block_8", ["1 HgC\n/\n,OH\n'\n"]], ["block_9", [".4\n.-\nc\nCH3\n"]], ["block_10", ["d\n..\n"]], ["block_11", ["Example 1.6\n"]], ["block_12", ["C\n\u00e9\n_\ni\n.J\na... 1.\nL\nl\nI\nl\nI\nI\n6\n5\n4\n3\n2\n1\n0\n"]], ["block_13", [{"image_1": "50_1.png", "coords": [45, 398, 285, 638], "fig_type": "figure"}]], ["block_14", ["3\n"]], ["block_15", ["I\n|_\n_|_\n_l\nI\n"]], ["block_16", ["E\n"]], ["block_17", ["Ha\nm\n:1\nCH3\n"]], ["block_18", ["ppm\n"]]], "page_51": [["block_0", [{"image_0": "51_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["phase without degradation. However, progress in recent years has been quite rapid, with two\ngeneral approaches being particularly productive. In one, electrospray ionization, a polymer\nsolution is ejected through a small ori\ufb01ce into a vacuum environment; an electrode at the exit\ndeposits a charge onto each drop of solution. The solvent then evaporates, leaving behind a charged\nmacromolecule in the gas phase. In the other, the polymer sample is dispersed in a particular\nmatrix on a solid substrate. An intense laser pulse is absorbed by the matrix, and the resulting\nenergy transfer vaporizes both polymer and matrix. For uncharged synthetic polymers, the\nnecessary charge is usually complexed with the polymer in the gas phase, after a suitable salt\nhas been codissolved in the matrix. This technique, matrix-assisted laser desorption/ionization\nmass spectrometry, or MALDI for short, is already a standard approach in the biopolymer arena,\nand is making substantial inroads for synthetic polymers as well.\nIt is worth recalling the basic ingredients of a mass spectrometric experiment. A sample molecule\nof mass m is first introduced into the gas phase in a high vacuum, and at some point in the process it\nmust acquire a net charge 2. The resulting ion is accelerated along a particular direction by suitably\nplaced electrodes, and ultimately collected and counted by a suitable analyzer. The ion acquires\nkinetic energy in the applied field, which depends on the net charge 2 and the applied voltage. This\nenergy will result in a mass-dependent velocity v, as the kinetic energy v/Z. This allows for the\ndiscrimination of different masses, by a variety of possible schemes. For example, if the ions\nexperience an orthogonal magnetic field B, their trajectories will be curved to different extents, and\nit is possible to tune the magnitude ofB to allow a particular mass to pass through an aperture before\nthe detector. For polymers it turns out to be more effective to use time-of-\ufb02ight (TOF) analysis.\nFor a given applied field and \ufb02ight path to the detector, larger masses will take longer to reach\nthe detector. As long as all the molecules are introduced into the gas phase at the same instant in\ntime, the time of arrival can be converted directly into a value of m/z.\nThe preceding discussion may give the misleading impression that MALDI is rather a straight-\nforward technique. This is not, in fact, the case, especially for synthetic polymers. A great deal\nremains to be learned about both the desorption and ionization processes, and standard practice\nis to follow particular recipes (matrix and salt) that have been found to be successful for a\ngiven polymer. For example, polystyrene samples are most often dispersed in dithranol (1,8,9-\nanthracenetriol) with a silver salt such as silver tri\ufb02uoroacetate. This mixture is co-dissolved in a\nvolatile common good solvent such as tetrahydrofuran, to ensure homogeneity; after depositing\na drop on a sample plate, the solvent is then allowed to evaporate. An intense pulse from a nitrogen\nlaser ()1 = 337 nm) desorbs some portion of the sample, and some fraction of the resulting gas-\nphase polystyrene molecules are complexed with a single silver cation.\nTwo examples of MALDI spectra on narrow distribution of polystyrene samples are shown in\nFigure 1.7a and Figure 1.7b. In the former, the average molecular weight is in the neighborhood\nof 5000, and different i-mers are clearly resolved. Each peak is separated by 104 g/mol, which is\nthe repeat unit molecular weight. The absolute molecular weight of each peak should correspond\n"]], ["block_2", ["This typical prepared by anionic polymerization in a \n"]], ["block_3", ["solvent \n"]], ["block_4", ["Mass spectrometry offers unprecedented resolution in the analysis of gas phase ions, and is a\nworkhorse of chemical analysis. Its application to synthetic polymers has been limited until rather\nrecently, primarily due to the difficulty of transferring high molecular weight species into the gas\n"]], ["block_5", ["Measurement of Molecular Weight\n35\n"]], ["block_6", ["1.8.3\nMALDI Mass Spectrometry\n"]], ["block_7", ["The percent 1,4 addition can be computed as follows:\n"]], ["block_8", ["26.9\n%1,4=\nx100=91.2%.\n[a + (lb/2)\n"]]], "page_52": [["block_0", [{"image_0": "52_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 1.7\nMALDI spectra for anionically prepared polystyrenes with (a) MW 2 4,971, Mn 2 4,620,\nPDI 1.076, and (b) MW 49,306, Mn 48,916, PDI \n1.008. Inset shows that even in (b) there is\nsome resolution of the different i-mers. (Data courtesy of K. Fagerquist, T. Chang, and T.P. Lodge.)\n"]], ["block_2", ["W\n'hMI-um\n"]], ["block_3", ["to the following sum: (1041' + 108 + 1 -|\u2014 57) :1041' + 166, where 108, 1, and 57 are the contribu-\ntions of the silver ion, terminal proton, and sec\u2014butyl initiator fragment, respectively. Using the\nformulas given in Equation 1.7.2 and Equation 1.7.4, Mn and MW can be calculated as 4620 and\n4971, respectively, and the polydispersity is 1.076. This sample, which was prepared by living\nanionic polymerization as described in Chapter 4, is thus quite narrow. Nevertheless, the plot\nshows distinctly how many different i-mers are present, and in what relative proportion. The\nassumption is made that the height of each peak is proportional to its number concentration in\nthe sample, and thus the yards corresponds to an unnormalized form of mole fraction x,. This\nimage, perhaps more than any other, underscores the point we have already made several times:\neven the best of polymer samples is quite heterogeneous. Recalling Example 1.4 and the tank car\nof polybutadiene, it is worth pointing out that each peak in Figure 1.7a corresponds to many\nstructurally different molecules, in terms of the stereochemical sequence along the backbone.\nThe MALDI spectrum in Figure 1.7b corresponds to a sample about 10 times higher in\nmolecular weight. At this point it is not possible to see any structure between different i-mers,\nalthough the expanded version shows that there is still a hint of resolution of distinct molecular\nweights. This serves to point out one limitation with MALDI, namely that its main attribute, high\nresolution, is diminished as M increases. Not apparent from this plot, but even more troublesome,\nis the fact that the absolute amplitude of the signal is greatly reduced compared to Figure 1.7a. It is\n"]], ["block_4", ["I\nI\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n|\nI\nI\nl\n|\nI\nI\nl\nI\nl\u2014I\nI\nI\nI\nI\n\u2014|\nI\u2018T llllllllllllllll\nI llllllllllll\n40900\n50,000\n60000\n48,000\n49,000\n50,000\n51,000\n"]], ["block_5", ["36\nIntroduction to Chain Molecules\n"]], ["block_6", ["IIIIII\nIIIIIIIIIIIII IIIIIIllillllllllllltlllllll\nII\n(a)\n2,000\n3,000\n4,000\n5,000\n6,000\n7,000\n"]], ["block_7", ["(b)\n"]], ["block_8", [{"image_1": "52_1.png", "coords": [39, 442, 336, 600], "fig_type": "figure"}]], ["block_9", [{"image_2": "52_2.png", "coords": [48, 284, 315, 412], "fig_type": "figure"}]], ["block_10", ["...||111\n,\n,._.11111111111111\u201c..\n"]], ["block_11", ["m/z\n"]], ["block_12", [{"image_3": "52_3.png", "coords": [230, 478, 399, 579], "fig_type": "figure"}]], ["block_13", ["IV;\n"]], ["block_14", ["80\u2018\u201d g no\nv38 :2 2\na? 22\u2018\n"]], ["block_15", ["V\n"]]], "page_53": [["block_0", [{"image_0": "53_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nMultiply charged ions can present a problem, because one cannot distinguish between a\nmolecule with a charge of l and a molecule of twice the molecular weight but with a charge\nof 2. In fact, if the technique were called \u201cmass-to-charge ratio spectrometry\u201d it would be a\nmouthful, but it would serve as a constant reminder of this important complication.\n4.\nIt is difficult to compare the amplitudes of peaks from one laser pulse to another, and from one\nsample drop to another. This presumably re\ufb02ects the microscopic details of the spot on the\nsample that is actually at the focus of the laser beam. As a consequence any quantitative\ninterpretation should be restricted to a given spectrum.\n"]], ["block_2", ["The contents of this book may be considered to comprise three sections, each containing four\nseparate chapters. The first section, including Chapter 2 through Chapter 5, addresses the synthesis\nof polymers, the various reaction mechanisms and kinetics, the resulting molecular weight\ndistributions, and some aspects of molecular characterization. In particular, Chapter 2 concerns\nstep-growth (condensation) polymerization and Chapter 3 chain\u2014growth (free radical) polymeriza-\ntion. Chapter 4 describes a family of particular polymerization schemes that permit a much higher\ndegree of control over molecular weight, molecular weight distribution, and molecular architecture\nthan those in the preceding two chapters. Chapter 5 addresses some of the factors that control the\nstructural details within polymers, especially copolymers and stereoregular polymers, and aspects\nof their characterization.\nThe second section takes up the behavior of polymers dissolved in solution. The conformations\nof polymers, and especially random coils, are treated in Chapter 6. Solution thermodynamics are\nthe subject of Chapter 7, including the concepts of solvent quality, osmotic pressure, and phase\nbehavior. The technique of light scattering, which provides direct information about molecular\nweight, solvent quality, and chain conformations, is covered in detail in Chapter 8. Chapter 9\nexplores the various hydrodynamic properties of polymers in solution, and especially as they\nimpact viscosity, diffusivity, and SEC.\nThe concluding section addresses the properties of polymers in the bulk, with a particular\nemphasis on the various solid states: rubber, glass, and crystal. Thus Chapter 10 considers polymer\nnetworks and their characteristic and remarkable elasticity. Chapter 11 treats the unusual visco-\nelastic behavior of polymer liquids, in a way that combines central concepts from both Chapter 9\nand Chapter 10. Chapter 12 introduces the phenomenon of the glass transition, which is central to\nall polymer materials yet relatively unimportant in most atomic or small molecule\u2014based materials.\nFinally, the rich crystallization properties of polymers are taken up in Chapter 13. The text\n"]], ["block_3", ["1.\nGenerally, the more polar a polymer, the easier it is to analyze by MALDI. Thus poly(ethylene\noxide) is relatively easy; poly(methyl methacrylate) is easier than polystyrene; polyethylene is\nalmost impossible.\n2.\nAn important, unresolved issue is relating the amplitude of the signal of a particular peak to\nthe relative abundance of that molecule in the sample. For example, are all molecular weights\ndesorbed to the same extent within a given laser pulse (unlikely), and are all molecular\nweights equally likely to be ionized once in the gas phase (no)? Consequently, it can be\ndangerous to extract MW and M,1 as we did for the samples in Figure 1.7a and Figure 1.7b,\nbecause the signals have an unknown sensitivity to molecular weight. (In this instance, this\nproblem is mitigated because the distributions are quite narrow.) In general, lower molecular\nweights have a much higher yield.\n"]], ["block_4", ["data in Figure 1.7b give MW 49,306, and a polydispersity of 1.008. This turns out to\nbe nearly as narrow as the theoretical limit for this class of polymerizations (see Chapter 4), yet it is\nstill obviously quite \nWe conclude this section with some further general observations about MALDI:\n"]], ["block_5", ["simply weight molecules into the gas phase. the\n"]], ["block_6", ["Preview of Things to Come\n37\n"]], ["block_7", ["1.9\nPreview of Things to Come\n"]]], "page_54": [["block_0", [{"image_0": "54_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["To a signi\ufb01cant extent the problems in this book are based on data from the original literature. In\nmany instances the values given have been estimated from graphs, transformed from other\nfunctional representations, or changed in units. Therefore, these quantities do not necessarily\nre\ufb02ect the accuracy of the original work, nor is the given number of significant figures always\njustified. Finally, the data may be usedfor purposes other than were intended in the original study.\n"]], ["block_2", ["lR.E. Cohen and AR. Ramos Macromolecules, 12, 131 (1979).\n"]], ["block_3", ["1.\nRE. Cohen and AR. RamosJr describe phase equilibrium studies of block copolymers of\nbutadiene (B) and isoprene (1). One such polymer is described as having a 2:1 molar ratio of B\nto I with the following microstructure:\nB\u201445% cis-1,4; 45% trans\u20141,4; 10% vinyl.\nI-\u2014-over 92% cis\u20141,4.\nDraw the structure of a portion of this polymer consisting of about 15 repeat units, and having\napproximately the composition of this polymer.\n2.\nHydrogenation of polybutadiene converts both cis and trans isomers to the same linear\nstructure, and vinyl groups to ethyl branches. A polybutadiene sample of molecular weight\n168,000 was found by infrared spectroscopy to contain double bonds consisting of 47.2% cis,\n"]], ["block_4", ["1.\nThe most important feature of a polymer is its degree of polymerization or molecular weight.\nFor example, even though polyethylene has the same chemical formula as the n-alkanes, it has\nremarkably different physical properties from its small molecule analogs.\n2.\nThe statistical nature of polymerization schemes inevitably leads to a distribution of molecular\nweights. These can be characterized via specific averages, such as the number-average and\nweight-average molecular weights, or by the full distribution, which can be determined by\nSEC or MALDI mass spectrometry.\n3.\nPolymers can exhibit many different architectures, such as linear, randomly branched, or\nregularly branched chains, and networks. Homopolymers contain only one type of repeat unit,\nwhereas copolymers contain two or more.\n4.\nThere are many possible variations in local structure along a polymer chain, which we have\nclassified as positional, stereochemical, or geometrical isomers. Given these possibilities, and\nthose identified in the previous point, it is unlikely that any two polymer molecules within a\nparticular sample have exactly the same chemical structure, even without considering differ-\nences in molecular weight.\n5.\nNatural polymers such as polysaccharides, proteins, and nucleic acids share many of the\nattributes of their synthetic analogs, and as such are an important part of the subject of this\nbook. On the other hand, the specific biological functions of these macromolecules, especially\nproteins and nucleic acids, fall outside our scope.\n"]], ["block_5", ["concludes with an Appendix that reviews some of the mathematical manipulations encountered\nthroughout the book.\n"]], ["block_6", ["In this chapter, we have introduced the central concept of chain molecules, and identified various\nways in which polymers may be classified. The importance of molecular weight and its distribution\nwas emphasized, and associated averages defined. Examples were given of the many possible\nstructural variations that commonly occur in synthetic polymers:\n"]], ["block_7", ["Problems\n"]], ["block_8", ["38\nIntroduction to Chain Molecules\n"]], ["block_9", ["1.10\nChapter Summary\n"]]], "page_55": [["block_0", [{"image_0": "55_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4.\nSome polymers are listed below using either IUPAC (1) names or acceptable trivial (T) names.\nDraw structural formulas for the repeat units in these polymers, and propose an alternative\nname in the system other than the one given:\nPolymethylene (I)\nPolyforrnaldehyde (T)\nPoly(pheny1ene oxide) (T)\nPoly[(2-propy1-1,3-dioxane-4,6-diyl)methylene] (I)\nPoly(1-acetoxyethylene) (I)\nPoly(methy1 acrylate) (T)\n"]], ["block_2", ["5.\nStar polymers are branched molecules with a controlled number of linear arms anchored to\none central molecular unit acting as a branch point. Schaefgen and Flory\u00a7 prepared poly\n(a-caprolactam) four-\nand eight-arm\nstars using cyclohexanone tetrapropionic\nacid\nand\ndicyclohexanone octapropionic acid as branch points. The authors present the following stoi-\nchiometric de\ufb01nitions/relations to relate the molecular weight of the polymer to the concentration\nof unreacted acid groups in the product. Provide the information required for each of the\nfollowing steps:\n(a)\nThe product has the formula R\u2014{\u2014CO[\u2014NH(CH2)5CO\u2014\u2014]y\u2014OH }b. What is the significance\nof R, y, and b?\n(b)\nIf Q is the number of equivalents of multifunctional reactant which react per mole of\nmonomer and L represents the number of equivalents of unreacted (end) groups per mole\nof monomer, then (1\u2014L)/(Q+L). Justify this relationship, assuming all functional\ngroups are equal in reactivity.\n(c)\nIf M0 is the molecular weight of the repeat unit and Mg, is the molecular weight of the\noriginal branch molecule divided by b, then the number-average molecular weight of\nthe star polymer is\n"]], ["block_3", ["3.\nLandel used a commercial material called Vulcollan\n18/40 to study the rubber-to-glass\ntransition of a polyurethane}\u2014 This material is described as being \u201cprepared from a low\nmolecular weight polyester which is extended and cross-linked by reacting it with naphtha-\nlene\u20141,4,-diisocyanate and 1,4-butanediol. The polyester is prepared from adipic acid and a\nmixture of ethylene and propylene glycols.\u201d Draw the structural formula of a portion of the\ncross\u2014linked polymer which includes the various possible linkages that this description\nincludes. Remember that isocyanates react with active hydrogens; use this fact to account\nfor the cross-linking.\n"]], ["block_4", ["Problems\n39\n"]], ["block_5", ["TW.E. Rochefort, (10. Smith, H. Rachapudy, V.R. Raju, and W.W. Graessley, J. Polym. Sci, Polym. Phys, 17, 1197 (1979).\n*R.F. Landel, J. Colloid Scan, 12, 308 (1957).\n\u00a7 J.R. Schaefgen and RI. Flory, J. Am. Chem. Soc., 70, 2709 (1948).\n"]], ["block_6", ["44.9% trans, and 7.9% vinyl.T After hydrogenation, what is the average number of backbone\ncarbon atoms between ethyl side chains?\n"]], ["block_7", ["((1)\nEvaluate MD for the following molecules:\n"]], ["block_8", [{"image_1": "55_1.png", "coords": [58, 479, 182, 520], "fig_type": "molecule"}]], ["block_9", ["l\u2014L\nMn=b M\u2014+M\n{\u00b0Q+L\nb}\n"]], ["block_10", ["Justify this result and evaluate M0 and Mb for the b 4 and b 8 stars.\n"]], ["block_11", ["4\n0.2169\n0.0018\n"]], ["block_12", ["o\nQ\nL\n"]], ["block_13", ["8\n0.134\n0.00093\n"]]], "page_56": [["block_0", [{"image_0": "56_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["40\n"]], ["block_2", ["tH. Batzer, Makromol. Chem, 5, 5 (1950).\nISN. Chinai, J.D. Matlock, A.L. Resnick, and R.J. Samuels, J. Polym. sci, 17, 391 (1955).\n"]], ["block_3", ["12.\n"]], ["block_4", ["10.\n"]], ["block_5", ["11.\n"]], ["block_6", [{"image_1": "56_1.png", "coords": [43, 472, 150, 520], "fig_type": "molecule"}]], ["block_7", [{"image_2": "56_2.png", "coords": [53, 222, 269, 281], "fig_type": "molecule"}]], ["block_8", [{"image_3": "56_3.png", "coords": [54, 486, 221, 613], "fig_type": "figure"}]], ["block_9", ["(a)\nH2CA\\n/OH\nO\n"]], ["block_10", ["(b)\nO\nO\n"]], ["block_11", ["O\n(c)\nHON/LLOH\n"]], ["block_12", ["(d)\nHsN\n"]], ["block_13", ["Consider a set consisting of 4\u20148 family members, friends, neighbors, etc. Try to select a\nvariety of ages, genders, and other attributes. Take the mass of each individual (a rough\nestimate is probably wiser than asking directly) and calculate the number- and weight\u2014\naverage masses for this set. Does the resulting PDI indicate a rather \u201cnarrow\u201d distribution?\nIf you picture this group in your mind, do you imagine them all to be roughly the same size,\nas the PDI probably suggests?\nProve that the polydispersity of the Schulz-Zimm distribution is given by Equation 1.7.21.\nYou may want to look up the general solution for integrals of the type fxae\u2019bxdx.\nIn Figure 1.5a through Figure 1.5c it appears that the maximum in w; corresponds closely to\nM\". Differentiate Equation 1.7.20 with respect to M,- to show why this is the case.\nThe MALDI spectrum in Figure 1.7b resembles a Gaussian or normal distribution. One\nproperty of a Gaussian distribution is that the half\u2014width at half-height of the peak is\napproximately equal to 120\u2018. Use this relation to estimate 0' from the trace, and compare it\nto the value you would get from Equation 1.7.16.\nGive the overall chemical reactions involved in the polymerization of these monomers, the\nresulting repeat unit structure, and an acceptable name for the polymer.\n"]], ["block_14", ["The Mark\u2014Houwink exponent a for poly(methyl methacrylate) at 25\u00b0C has the value 0.69\nin acetone and 0.80 in chloroform. Calculate (retaining more significant figures than\nstrictly warranted) the value of MV that would be obtained for a sample with the following\nmolecular weight distribution if the sample were studied by viscometry in each of these\nsolvents.i Compare the values of MV with MH and MW.\n"]], ["block_15", ["Batzer reported the following data for a fractionated polyester made from sebacic acid and\n1,6-hexanediolff evaluate Mn, MW, and M2.\n"]], ["block_16", ["M,x10\"5(g/mol)\n2.0\n4.0\n6.0\n8.0\n10.0\n"]], ["block_17", ["Mass (g)\n1.15\n0.73\n0.415\n0.35\n0.51\n0.34\n1.78\n0.10\n0.94\nM X 10'4\n1.25\n2.05\n2.40\n3.20\n3.90\n4.50\n6.35\n4.10\n9.40\n"]], ["block_18", ["Fraction\n1\n2\n3\n4\n5\n6\n7\n8\n9\n"]], ["block_19", ["n,x103(mol)\n1.2\n2.7\n4.9\n3.1\n0.9\n"]], ["block_20", [{"image_4": "56_4.png", "coords": [62, 509, 214, 593], "fig_type": "molecule"}]], ["block_21", [{"image_5": "56_5.png", "coords": [71, 515, 211, 560], "fig_type": "molecule"}]], ["block_22", [{"image_6": "56_6.png", "coords": [83, 528, 215, 552], "fig_type": "molecule"}]], ["block_23", [{"image_7": "56_7.png", "coords": [86, 476, 140, 517], "fig_type": "molecule"}]], ["block_24", ["HQN\u2018E/ENH? CIMCI\n"]], ["block_25", [{"image_8": "56_8.png", "coords": [99, 480, 207, 624], "fig_type": "molecule"}]], ["block_26", ["Me\n"]], ["block_27", ["Introduction to Chain Molecules\n"]]], "page_57": [["block_0", [{"image_0": "57_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["H.R. Allcock, F.W. Lampe, and J.E. Mark, Contemporary Polymer Chemistry, 3rd ed., Prentice Hall, Engle\u2014\nwood Cliffs, NJ, 2003.\nR]. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\nA.Y. Grosberg, and AR. Khokhlov, Giant Molecules, Here, There and Everywhere, Academic Press, San\nDiego, CA, 1997.\nAL. Lehninger, Biochemistry, 2nd ed., Worth Publishers, New York, 1975.\nP. Munk and TM. Aminabhavi, Introduction to Macromolecular Science, 2nd ed., Wiley Interscience, New\nYork, 2002.\nG. Odian, Principles of Polymerization, 4th ed., Wiley, New York, 2004.\nRC. Painter and M.M. Coleman, Fundamentals of Polymer Science, Technomic, Lancaster, PA, 1994.\nRB. Seymour and CE. Carraher, Jr., Polymer Chemistry: An Introduction, Marcel Dekker, New York, 1981.\nL.H. Sperling, Introduction to Physical Polymer Science, Wiley Interscience, New York, 1986.\nR]. Young and PA. Lovell, Introduction to Polymers, 2nd ed., Chapman and Hall, London, 1991.\n"]], ["block_2", ["Further Readings\n41\n"]], ["block_3", ["Further Readings\n"]], ["block_4", ["References\n"]], ["block_5", ["l\u2014n 0\nChemical cf: Engineering News, 58, 29 (1980).\nThe Graduate, directed by Mike Nichols, Embassy Pictures Corporation, 1967.\nNo enterprise as rich as polymer science has only one \u201cfather.\u201d Herman Mark is one of those to whom the\ntitle could readily be applied. An interesting interview with Professor Mark appears in the Journal of\nChemical Education, 56, 38 (1979).\nH. Morawetz, Polymers: The Origins and Growth of a Science, Wiley, New York, 1985.\nRB. Seymour, History of Polymer Science and Technology, Marcel Dekker, New York, 1982.\nCS. Marvel, another pioneer in polymer chemistry, reminisced about the early days of polymer chemistry\nin the United States in the Journal of Chemical Education, 58, 535 (1981).\nSee IUPAC Macromolecular Nomenclature Commission, Macromolecules, 6, 149 (1973).\n"]], ["block_6", ["13.\n"]], ["block_7", ["14.\n"]], ["block_8", ["15.\n"]], ["block_9", ["16.\n"]], ["block_10", [{"image_1": "57_1.png", "coords": [44, 38, 232, 130], "fig_type": "figure"}]], ["block_11", [{"image_2": "57_2.png", "coords": [52, 109, 146, 154], "fig_type": "molecule"}]], ["block_12", ["(e)\nQ\n\"'\nHOHOH\n"]], ["block_13", ["(f)\nH2N WOH\n"]], ["block_14", ["A MALDI-TOF analysis of a polystyrene sample exhibited a peak (one of many) at 1206.\nThe sample was prepared in dithranol, with silver nitrate as the salt. Assuming no head-to-\nhead defects, how many distinct chemical structures could this peak represent? Propose\nstructures for the end groups of the polymer as well.\nProton NMR is used to attempt to quantify the molecular weight of a poly(ethylene oxide)\nmolecule with methyoxy end groups at each terminus. If the integration of the methyl protons\nrelative to the methylene protons gave a ratio of 1:20, what can you say about the molecular\nweight?\nWhat would be Mw and Mn for a sample obtained by mixing\n10 g of polystyrene\n"]], ["block_15", ["(MW 100,000, Mn 70,000) with 20 g of another polystyrene (Mw 60,000, Mn 20,000)?\nWhat would MW and Mn be for an equimolar mixture of tetradecane and decane? (Ignore\nisotope effects.)\n"]], ["block_16", [{"image_3": "57_3.png", "coords": [88, 52, 223, 119], "fig_type": "molecule"}]]], "page_58": [["block_0", [{"image_0": "58_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In Section 1.4, we discussed the classi\ufb01cation of polymers into the categories of addition or\ncondensation. At that time we noted that these classifications could be based on the following:\n"]], ["block_2", ["It is the third of these criteria that offers the most powerful insight into the nature of the\npolymerization process for this important class of materials. We shall sometimes use the terms\nstep-growth and condensation polymers as synonyms, although step-growth polymerization encom\u2014\npasses a wider range of reactions and products than either criteria (1) or (2) above would indicate.\nThe chapter is organized as follows. First, we examine how the degree of polymerization and its\ndistribution vary with the progress of the polymerization reaction, with the latter defined both in\nterms of stoichiometry and time (Section 2.2 through Section 2.4). Initially we consider these\ntopics for simple reaction mixtures, that is, those in which the proportions of reactants agree\nexactly with the stoichiometry of the reactions. After this, we consider two important classes of\ncondensation or step\u2014growth polymers: polyesters and polyamides (Section 2.5 and Section 2.6).\nFinally we consider nonstoichiometric proportions of reactants (Section 2.7). The important case of\nmultifunctional monomers, which can introduce branching and cross-linking into the products, is\ndeferred until Chapter 10.\n"]], ["block_3", ["1.\nStoichiometry of the polymerization reaction (small molecule eliminated?)\n2.\nComposition of the backbone of the polymer (atoms other than carbon present?)\n3.\nMechanism of the polymerization (stepwise or chain reaction?)\n"]], ["block_4", ["As the name implies, step-growth polymers are formed through a series of steps, and high\nmolecular weight materials result from a large number of steps. Although our interest lies in\nhigh molecular weight, long-chain molecules, a crucial premise of this chapter is that these\nmolecules can be effectively discussed in terms of the individual steps that lead to the formation\nof the polymer. Thus, polyesters and polyamides are substances that result from the occurrence of\nmany steps in which ester or amide linkages are formed between the reactants. Central to our\ndiscussion is the idea that these steps may be treated in essentially the same way, whether they\noccur between small molecules or polymeric species. We shall return to a discussion of the\nimplications and justification of this assumption of equal reactivity throughout this chapter.\n"]], ["block_5", ["2.2\nCondensation Polymers: One Step at a Time\n"]], ["block_6", ["2.1\nIntroduction\n"]], ["block_7", ["2.2.1\nClasses of Step-Growth Polymers\n"]], ["block_8", ["Here are examples of important classes of step-growth polymers:\n"]], ["block_9", ["1.\nPolyesters\u2014successive reactions between diols and dicarboxylic acids:\n"]], ["block_10", ["RI\nn Ho/HT'OH n HOJLR'JLOH\n\u2014\"\" {O\u2018H/Om/ \ufb02\n"]], ["block_11", [{"image_1": "58_1.png", "coords": [55, 575, 349, 629], "fig_type": "molecule"}]], ["block_12", ["Step-Growth Polymerization\n"]], ["block_13", ["0\n0\n"]], ["block_14", [{"image_2": "58_2.png", "coords": [146, 584, 337, 621], "fig_type": "molecule"}]], ["block_15", ["2\n"]], ["block_16", ["43\n"]], ["block_17", ["+ 2n H20\n(2A)\n0 n\nO\n"]]], "page_59": [["block_0", [{"image_0": "59_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "59_1.png", "coords": [19, 553, 280, 653], "fig_type": "figure"}]], ["block_2", [{"image_2": "59_2.png", "coords": [29, 297, 172, 331], "fig_type": "molecule"}]], ["block_3", ["aA and B represent two different functional groups and ab is the product\nof their reaction with each other. Consult the text for a discussion of the\nline-by\u2014line development of the reaction.\n"]], ["block_4", ["Of course, in Reaction (2.A) and Reaction (2B) the hydrocarbon sequences R and R\u2019 can be the\nsame or different, contain any number of carbon atoms, be linear or cyclic, and so on. Likewise, the\ngeneral reactions, Reaction (2.C) and Reaction (2E), certainly involve hydrocarbon sequences\nbetween the reactive groups A and B. The notation involved in these latter reactions is particularly\nconvenient, however, and we shall use it extensively in this chapter. It will become clear as we\nproceed that the stoichiometric proportions of reactive groups\u2014A and B in the above notation\u2014\nplay an important role in determining the characteristics of the polymeric product. Accordingly, we\nshall confine our discussion for the present to reactions of the type given by Reaction (2E), since\nequimolar proportions of A and B are assured by the structure of the monomer.\n"]], ["block_5", ["Table 2.1\nHypothetical Step\u2014Growth Polymerization\nof 10 AB Moleculesa\n"]], ["block_6", ["44\nStep-Growth Polymerization\n"]], ["block_7", ["Table 2.] presents a hypothetical picture of how Reaction (2.E) might appear if we examined the\ndistribution of product molecules in detail. Row 1 of Table 2.1 shows the initial pool of monomers,\n"]], ["block_8", ["3.\nAbababaB AbaB AbabaB AB\n4.\nAbababaB AbabababaB AB\n5.\nAbababaB AbababababaB\n6,\nAbababababababababaB\n"]], ["block_9", ["Row\nMolecular species present\n"]], ["block_10", ["1.\nAB AB AB AB AB AB AB AB AB AB\n2.\nAbaB AbaB AbaB AbaB AB AB\n"]], ["block_11", ["2.\nPolyamides\u2014successive reactions between diamines and dicarboxylic acids:\n"]], ["block_12", ["Since the two reacting functional groups can be located in the same reactant molecule, we add\nthe following:\n4.\nPoly(amino acid)\n"]], ["block_13", ["2.2.2\nFirst Look at the Distribution of Products\n"]], ["block_14", ["5.\nGeneral\n"]], ["block_15", ["3.\nGeneral\u2014successive reactions between difunctional monomer A\u2014A and difunctional mono-\n"]], ["block_16", [{"image_3": "59_3.png", "coords": [39, 214, 246, 286], "fig_type": "figure"}]], ["block_17", [{"image_4": "59_4.png", "coords": [48, 223, 245, 281], "fig_type": "molecule"}]], ["block_18", ["mer B\u2014B:\n"]], ["block_19", ["n H,N/'\\c/\u00b0H fN\n+ n H20\n(2D)\n"]], ["block_20", ["n A\u2014B_yl(a\\b/)H+\n(2.13)\n"]], ["block_21", ["0\nO\nH\nH\nR\nN\nHI\nn H2N/\n\\NH2\n+\n\u201dHOJLHIJ%H\n"]], ["block_22", ["n A\u2014A + n B\u2014B\u2014hta\\a/b\\b/)n+\n(2.0\n"]], ["block_23", ["Fl\nF!\n"]], ["block_24", ["H\nH\no\nn\nO\n"]], ["block_25", [{"image_5": "59_5.png", "coords": [145, 79, 336, 117], "fig_type": "molecule"}]], ["block_26", ["_+ H2O\n(2.B)\n00:1\n"]]], "page_60": [["block_0", [{"image_0": "60_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["10 molecules in a possible composition after a certain amount of\nreaction Section that the particular that account\nfor the rows are not highly probable. Our objective here is\nnot to assess the probability of certain reactions, but rather to consider some possibilities.\nStoichiometrically, we can still account for the initial set of 10 A groups and 10 B groups; we\nindicate those that have reacted with each other as ab groups. The same conservation of atom\ngroupings would be obtained if row 2 showed one trimer, two dimers, and three monomers instead\nof the four dimers and two monomers indicated in Table 2.1. Other combinations could also\n"]], ["block_2", ["be assembled. These possibilities indicate one of the questions that we shall answer in this chapter:\nHow do the distribute among the different possible species as the reaction\nproceeds?\nRow 3 of Table 2.1 shows the mixture after two more reaction steps have occurred. Again,\nthe components we have elected to show are an arbitrary possibility. For the monomer system\nwe have chosen, the concentration of A and B groups in the initial monomer sample are equal\nto each other and equal to the concentration of monomer. In this case, an assay of either A groups\nor B groups in the mixture could be used to monitor the progress of the reaction. Choosing the\nnumber of A groups for this purpose, we see that this quantity drops from 10 to 6 to 4, respectively,\nas we proceed through rows 1, 2, and 3 of Table 2.1. What we wish to point out here is the fact that\nthe 10 initial monomers are now present in four molecules, so the number average degree of\npolymerization is only 2.5, even though only 40% of the initial reactive groups remain. Another\nquestion is thus raised: In general, how does the average molecular weight vary with the extent of\nthe reaction?\nThe reaction mixture in the fourth row of Table 2.1 is characterized by a number average degree\nof polymerization Nn 10/3 3.3, with only 30% of the functional groups remaining. This means\nthat 70% of the possible reactions have already occurred, even though we are still dealing with a\nvery low average degree of polymerization. Note that the average degree of polymerization would\n"]], ["block_3", ["Condensation Polymers: One Step at a Time\n45\n"]], ["block_4", ["be the same if the 70% reaction of functional groups led to the mixture AbababababababaB and\ntwo AB\u2019s. This is because the initial 10 monomers are present in three molecules in both instances,\nand we are using number averages to talk about these possibilities. The weight average would be\ndifferent in the two cases. This poses still another question: How does the molecular weight\ndistribution vary with the extent of reaction?\nBy the fifth row, the reaction has reached 80% completion and the number average value of the\ndegree of polymerization Nn is 5. Although we have considered this slowly evolving polymer\nin terms of the extent of reaction, another question starts to be worrisome: How long is this going\nto take?\nThe sixth row represents the end of the reaction as far as linear polymer is concerned. Of the 10\ninitial A groups, 1 is still unreacted, but this situation raises the possibility that the decamer shown\nin row 6\u20140r for that matter, some other i-mer, including monomer\u2014might form a ring or cyclic\ncompound, thereby eliminating functional groups without advancing the polymerization. Through-\nout this chapter we will assume that the extent of ring formation is negligible.\nIt is an easy matter to generalize the procedure we have been following and express the number\naverage degree of polymerization in terms of the extent of reaction, regardless of the initial sample\nsize. We have been dividing the initial number of monomers present by the total number of\nmolecules present after any extent of reaction. Each molecular species\u2014whether monomer or\npolymer of any length\u2014contains just one A group. The total number of monomers is therefore\nequal to the initial (superscript 0) number of A groups, 12%; the total number of molecules at any\nextent of reaction (no superscript) is equal to the number of A groups, VA, present at that point. The\nnumber average degree of polymerization is therefore given by\n"]], ["block_5", ["V0\nNn : A\n(2.2.1)\nVA\n"]]], "page_61": [["block_0", [{"image_0": "61_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Most of the questions raised in the past few paragraphs will be answered during the course of this\nchapter, some for systems considerably more involved than the one considered here. Before\nproceeding further, we should reemphasize one premise that underlies the entire discussion of\nTable 2.1: How do the chemical reactivities of A and B groups depend on the degree of\npolymerization of the reaction mixture? In Table 2.1, successive entries were generated by simply\nlinking together at random those species present in the preceding row. We have thus assumed that,\nas far as reactivity is concerned, an A reacts as an A and a B reacts as a B, regardless of the size of\nthe molecule to which the group is attached. If this assumption of equal reactivity is valid, it results\nin a tremendous simplification; otherwise we shall have to characterize reactivity as a function of\ndegree of polymerization, extent of reaction, and so on.\nOne of the most sensitive tests of the dependence of chemical reactivity on the size of the\nreacting molecules is the comparison of the rates of reaction for compounds that are members of a\nhomologous series with different chain lengths. Studies by Flory and others on the rates of\nesteri\ufb01cation and saponification of esters were the first investigations conducted to clarify the\ndependence of reactivity on molecular size [1]. The rate constants for these reactions are observed\nto converge quite rapidly to a constant value that is independent of molecular size, after an initial\ndependence on molecular size for small molecules. In the esterification of carboxylic acids, for\nexample, the rate constants are different for acetic, propionic, and butyric acids, but constant for\ncarboxylic acids with 4\u201418 carbon atoms. This observation on nonpolymeric compounds has been\n"]], ["block_2", ["This expression is consistent with the analysis of each of the rows in Table 2.1 as presented above\nand provides a general answer to one of the questions posed there. It is often a relatively easy\nmatter to monitor the concentration of functional groups in a reaction mixture; Equation 2.2.4\nrepresents a quantitative summary of an end group method for determining Nn. We reiterate that\nEquation 2.2.4 assumes equal numbers of A and B grOUps, with none of either lost in nonpolymer\nreactions.\nFrom row 6 in Table 2.1, we see that Nn=10 when p=0.9. The fact that this is also the\nmaximum value for N is an artifact of the example. In a larger sample of monomers higher average\ndegrees of polymerization are attainable. Equation 2.2.4 enables us to calculate that Nn becomes\n20, 100, and 200 for extents of reaction 0.950, 0.990, and 0.995, respectively. These results reveal\nwhy condensation polymers are often of relatively modest molecular weight: it may be very\ndifficult to achieve the extents of reaction required for very high molecular weights. As p increases\nthe concentration of H20 (or other small molecule product) will increase, and the law of mass\naction will Oppose further polymerization. Consequently, steps must be taken to remove the small\nmolecule as it is formed, if high molecular weights are desired.\n"]], ["block_3", ["It is convenient to define the fraction of reacted functional groups in a reaction mixture by a\nparameter p, called the extent of reaction. Thus, p is the fraction of A groups that have reacted at\nany stage of the process, and 1\u2014 p is the unreacted fraction:\n"]], ["block_4", ["46\nStep-Growth Polymerization\n"]], ["block_5", ["Comparison of Equation 2.2.1 and Equation 2.2.2 enables us to write very simply:\n"]], ["block_6", ["2.2.3\nA First Look at Reactivity and Reaction Rates\n"]], ["block_7", ["01\u2018\n"]], ["block_8", ["p \n(2.2.3)\n\u201dA\n"]], ["block_9", ["N11 \u2014\u20141\u2014\u2014\n(2.2.4)\n1 p\n"]], ["block_10", ["1 p 3\u20193\n(2.2.2)\nVA\n"]]], "page_62": [["block_0", [{"image_0": "62_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "62_1.png", "coords": [25, 29, 337, 182], "fig_type": "figure"}]], ["block_2", [{"image_2": "62_2.png", "coords": [26, 46, 183, 165], "fig_type": "figure"}]], ["block_3", ["where the parentheses represent the caged pair, as in Figure 2.1, and the ks are the rate constants for\nthe individual steps: k, and k0 for diffusion into and out of the cage, respectively, and k, for the\nreaction itself.\nSince this is the first occasion we have had to examine the rates at which chemical reactions\noccur, a few remarks about mechanistic steps and rate laws seem appropriate. The reader who feels\nthe need for additional information on this topic should consult any introductory physical chem-\nistry text.\nAs a brief review we recall the following:\n"]], ["block_4", ["generalized to polymerization reactions as well. The latter are subject to several complications that\nare not involved in the study of simple model compounds, but when these complications are\nproperly considered, the independence of reactivity on molecular size has been repeatedly veri\ufb01ed.\nThe foregoing conclusion does not mean that a constant rate of reaction persists throughout\nTable 2.1. The rate of reaction depends on the concentrations of reactive groups, as well as on their\nreactivities. Accordingly, rate of the reaction decreases as the extent of reaction progresses.\nWhen the rate law for the reaction is extracted from proper kinetic experiments, speci\ufb01c reactions\nare found to be characterized by \ufb01xed rate constants over a range of M, values.\nAmong the further complications that can interfere with this conclusion is the possibility that\nthe polymer becomes insoluble beyond a critical molecular weight or that the low molecular\nweight by-product molecules accumulate and thereby shift the equilibrium to favor reactants. It is\nalso possible that the transport of reactants will be affected by the increasing viscosity of the\npolymerization medium, which is a very complicated issue.\nFigure 2.] suggests that reactive end-groups may be brought into contact by rotation around\nonly a few bonds, an effect which is therefore independent of chain length. Once in close\nproximity, the A and B groups may be thought of as being in the same \u201ccage\u201d defined by near\nneighbors. It may take some time for the two reactive groups to diffuse together, but it will also\ntake some time for them to diffuse apart; this provides the opportunity to react. The rate at which\nan A and a B group react to form an ab linkage therefore depends on the relative rates of three\nprocesses: the rate to diffuse together, the rate at which they diffuse apart, and the rate at which\n\u201ctrapped\u201d A and B groups react. These considerations can be expressed more quantitatively by\nwriting the process in terms of the following mechanism:\n"]], ["block_5", ["Figure 2.1\nThe reaction of A and B groups at the ends of two different chains. Note that rotations around\nonly a few bonds will bring A and B into the same cage of neighboring groups, indicated by the dashed-line\nenclosure.\n"]], ["block_6", [{"image_3": "62_3.png", "coords": [35, 500, 208, 531], "fig_type": "molecule"}]], ["block_7", ["Condensation \n47\n"]], ["block_8", ["1.\nThe rate of a process is expressed by the derivative of a concentration (square brackets) with\nrespect to time, d[A]/dt. If the concentration of reaction product is used, this quantity is\n"]], ["block_9", ["3 . ~\\\\u201c\\'\\X\\\n.\u2018\n\\\\:-\\\\\"\u2018-_I\n\\\\\\\\\\\\\\\\\\\u00a7\n\\\n'\n\u20183.\n.\n..'*\n\\\\\n\\ '-\n\u2018\n\\\\u2018uVR\nu x \\ ~\n.\u2018\\\\ ..\n\\\n.\n-_u\u2018\n-\nx.\nk\n-\\\nn\n'-\n.o'\n\\\\u2018 \\\\\u2018~. \u2018A\u2018\nQ\n\\ mu o.:.\\\\\\.. o...\u2018x\\\\\nu~_. you u o;\n.\nuu\u2018\\\\\u201c~\\\\\u2018\\\\Q\u2018\\\u00a7\u2018t\\.\\\\2u\\\\u2018uu\\yu\u20183\\\\\u201c\\\n{on ' i}.-.\\<=.u\\\u00a7oo.- . .\\ \u2018i'u {Q\\ \ufffd\\.\\. \\.\\~:~:u\\ \u2018nuu \\\\\\\\u2018 -\\\\u2018 xxx\u00bb.-\n\\A. . 3-.\\\"\\\\\\\n"]], ["block_10", ["Qk\u201cubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ub>\u2018ub>mubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>u2018ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>\n.. ..ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>u2018ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>~ubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>\n:ubub>>ub>\\ubub>>ub>\na\nubub>>ub>-ubub>>ub>\nNo\n\u201c ub>\u2018ub>\nubub>>ub>um:ubub>>ub>\nI\u201c\nubub>>ub>\\ubub>>ub>\nub>>ub>\n\u201dH.\nub>\u2018ub>\nubub>>ub>\\ubub>>ub>ubub>>ub>\\ubub>>ub>\nu\nub>\u2018ub>ubub>>ub>\\ubub>>ub> \u201c\n"]], ["block_11", ["\u2014A + \u2014B\n$11;\n(\u2014A + B\u2014) \u00a35 \n(2F)\n"]], ["block_12", ["0\n"]], ["block_13", ["..\nx\u2018\n\\\\u2018\n\\\n\"._ \\\\_\\\n. \n"]], ["block_14", ["Q.\n"]], ["block_15", [{"image_4": "62_4.png", "coords": [196, 56, 372, 162], "fig_type": "figure"}]]], "page_63": [["block_0", [{"image_0": "63_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["positive; if a reactant is used, it is negative and a minus sign must be included. Also, each\nderivative d[A]/dt should be divided by the coefficient of that component in the chemical\nequation that describes the reaction, so that a single rate is described whichever component in\nthe reaction is used to monitor it.\n2.\nA rate law describes the rate of reaction as the product of a constant k, called the rate constant,\nand various concentrations, each raised to specific powers. The power of an individual\nconcentration term in a rate law is called the order with respect to that component, and the\nsum of the exponents of all concentration terms gives the overall order of the reaction. Thus in\nthe rate law Rate :k[X]1[Y]2, the reaction is first order in X, second order in Y, and third\norder overall.\n3.\nA rate law is determined experimentally and the rate constant evaluated empirically. There\nis no necessary connection between the stoichiometry of a reaction and the form of the\nrate law.\n4.\nA mechanism is a series of simple reaction steps that, when added together, account for the\noverall reaction. The rate law for the individual steps of the mechanism may be written by\ninspection of the mechanistic steps. The coefficients of the reactants in the chemical equation\ndescribing the step become the exponents of these concentrations in the rate law for that step.\n5.\nFrequently it is possible to write more than one mechanism that is compatible with an\nobserved rate law. Thus, the ability to account for an experimental rate law is a necessary\nbut not a sufficient criterion for the correctness of the mechanism.\n"]], ["block_2", ["48\nStep-Growth Polymerization\n"]], ["block_3", ["These ideas are readily applied to the mechanism described by Reaction (2.F). To begin with, the\nrate at which ab links are formed is first order with reSpect to the concentration of entrapped pairs.\nIn this sense, the latter behaves as a reaction intermediate or transition state according to this\nmechanism. Therefore\n"]], ["block_4", ["We shall return to this type of kinetic analysis in Chapter 3 where we discuss chain-growth\npolymerization.\nAccording to the mechanism provided by Reaction (2F) and the analysis given by Equation\n2.2.8, the rate of polymerization is dependent upon the following:\n"]], ["block_5", ["1.\nThe concentrations of both A and B, hence the reaction slows down as the conversion to\npolymer progresses.\n2\nThe three constants associated with the rates of the individual steps in Reaction (2.F).\n"]], ["block_6", ["The concentration of entrapped pairs is assumed to exist at some stationary\u2014state (subscript 5) level\nin which the rates of formation and loss are equal. In this stationary state d[(FA + B\ufb02)]/dt 0 and\nEquation 2.2.6 becomes\n"]], ["block_7", ["where the subscript s reminds us that this is the stationary\u2014state value. Substituting Equation 2.2.7\ninto Equation 2.2.5 gives\n"]], ["block_8", ["These entrapped pairs, in turn, form at a rate given by the rate at which the two groups diffuse\ntogether minus the rate at which they either diffuse apart or are lost by reaction:\n"]], ["block_9", ["k0 + kr\nRate of ab formation :\n[A][B]\n(22.8)\n"]], ["block_10", ["Rate of ab formation :kr[(\u2014A + B\u2014)]\n(2.2.5)\n"]], ["block_11", ["d[(eA + B\u2014)]\ndt\n= kilAllB] a Icon\u2014A + 3\u2014)] \u2014 kin\u2014A + 3\u2014)]\n(2.2.6)\n"]], ["block_12", ["k1\n[(\u2014A + 3\u2014)]5 \nk0 +16: \n(22-7)\n"]], ["block_13", ["kik,\n"]]], "page_64": [["block_0", [{"image_0": "64_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4.\nThe two constants k0 describe exactly the same kind of diffusional process, and differ\nonly in direction. Hence they should have the same dependence on molecular size, whatever\nthat might be, and that dependence therefore cancels out.\n5.\nThe mechanism in Reaction (2.F) is entirely comparable to the same reaction in low molecular\nweight systems. Such reactions involve considerably larger activation energies than physical\nprocesses like diffusion and, hence, do proceed slowly.\n"]], ["block_2", ["Note that the rate only on and any size dependence for this constant\nwould not cancel out.\nBoth Equation 2.2.9 and Equation 2.2.10 predict rate laws that are first order with respect to the\nconcentration of each of the reactive groups; the proportionality constant has a different signifi-\ncance in the two cases, however. The observed rate laws, which suggest a reactivity that is\nindependent of molecular size and the a priori expectation cited in item 5 above regarding\nthe magnitudes of different kinds of k values, lend credibility to the version presented in\nEquation 2.2.9.\nOur objective in the preceding argument has been to justify the attitude that each ab linkage\nforms according to the same rate law, regardless of the extent of the reaction. While our attention is\nfocused on the rate laws, we might as well consider the question, raised above, about the actual\nrates of these reactions. This is the topic of the next section.\n"]], ["block_3", ["In this section we consider the experimental side of condensation kinetics. The kinds of ab links\nthat have been most extensively studied are ester and amide groups, although numerous additional\nsystems could also be cited. In many of these systems the carbonyl group is present and believed to\nplay an important role in stabilizing the actual chemical transition state involved in the reactions.\nThe situation can be represented by the following schematic reaction:\n"]], ["block_4", ["3.\nIf the rate of chemical reaction is slow compared to the rate of group diffusion (kr << ki, k0),\nthen Equation 2.2.8 reduces to\n"]], ["block_5", ["6.\nIf k, >> ki, k0, then Equation 2.2.8 reduces to\n"]], ["block_6", ["2.3\nKinetics of Step-Growth Polymerization\n"]], ["block_7", ["in which the intermediate is stabilized by coordination with protons, metal ions, or other Lewis\nacids. The importance of this is to emphasize that the kinds of reactions we are considering are\noften conducted in the presence of an acid catalyst, frequently something like a sulfonic acid or a\nmetal oxide. The purpose of a catalyst is to modify the rate of a reaction, so we must be attentive to\nthe situation with respect to catalysts. At present, we assume a constant concentration of catalyst\nand attach a subscript c to the rate constant to remind us of the assumption. Accordingly, we write\n"]], ["block_8", ["Kinetics of Step-Growth Polymerization\n49\n"]], ["block_9", ["which is consistent with both Equation 2.2.9 and Equation 2.2.10. We expect the constant kc to be\ndependent on the concentration of the catalyst in some way which means that Equation 2.3.1 may\n"]], ["block_10", [{"image_1": "64_1.png", "coords": [38, 484, 300, 536], "fig_type": "molecule"}]], ["block_11", ["0\no\n_\n0\nRJLX\n+\nY\u201c\n\u2014\u2014h-\nR\u2014l\u2014Y\n-\u2014-\u2014\u2014-)-\nJL\n+\nx\u2014\n(Z\u2018G)\nx\nFl\nY\n"]], ["block_12", ["\u2014 99\u2014] = kc[A][B]\n(2.3.1)\ndt\n"]], ["block_13", ["Rate of ab formation ki[A][B]\n(2.2.10)\n"]], ["block_14", ["Rate of ab formation \nk\u2014kr[A][B]\n(2.2.9)\n"]], ["block_15", [{"image_2": "64_2.png", "coords": [179, 486, 281, 530], "fig_type": "molecule"}]]], "page_65": [["block_0", [{"image_0": "65_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["that after cancellation and rearrangement gives\n"]], ["block_2", ["be called a pseudo-second\u2014order rate law. We shall presently consider these reactions in the\nabsence of external catalysts. For now it is easier to proceed with the catalyzed case.\n"]], ["block_3", ["now differentiate with respect to t, noting that only x is a function of r:\n"]], ["block_4", ["Neither [A]0 nor [B]0 are functions of 1, although both [A] and [B] are. We write the latter two as\n[A] [A]0 x and [B] [B]0 x. Substitute these results into Equation 2.3.3 and rearrange:\n"]], ["block_5", ["Since d[A]/dt= Fdx/dt by the definition of x, this proves Equation 2.3.3 to be a solution to\nEquation 2.3.1. Equation 2.3.3 is undefined in the event [A]0 [B]0, but in this case the expression\nis anyhow inapplicable. Since A and B react in a 1:1 proportion, their concentrations are identical\nat all stages of reaction if they are equal initially. In this case, Equation 2.3.1 would reduce to a\nsimpler second-order rate law, which integrates to Equation 2.3.2.\n"]], ["block_6", ["We shall now proceed on the assumption that [A]0 and [B]0 are equal. As noted above, having\nboth reactive groups on the same molecule is one way of enforcing this condition. Accordingly, we\n"]], ["block_7", ["Both of these results are readily obtained; we examine the less obvious relationship in Equation\n2.3.3 in the following example. The consequences of different A and B concentrations on the\nmolecular weight of the polymer will be discussed in Section 2.7.\n"]], ["block_8", ["Equation 2.3.1 is the differential form of the rate law that describes the rate at which A groups are\nused up. To test a proposed rate law and evaluate the rate constant it is preferable to work with the\nintegrated form of the rate law. The integration of Equation 2.3.1 yields different results, depend\u2014\ning on whether the concentrations of A and B are the same or different:\n"]], ["block_9", ["2.3.1\nCatalyzed Step-Growth Reactions\n"]], ["block_10", ["the reaction, and [A]0 and [B]0 as the concentrations of these groups at I: 0.\n2.\nIf [A]0 [B]0, the integration of Equation 2.3.1 yields\n"]], ["block_11", ["By differentiation, verify that Equation 2.3.3 is a solution to Equation 2.3.1 for the conditions\ngiven.\n"]], ["block_12", ["3.\nIf [Ale 75 [B]O, \n"]], ["block_13", ["Solution\n"]], ["block_14", ["50\nStep\u2014Growth Polymerization\n"]], ["block_15", ["Example 2.1\n"]], ["block_16", ["1.\nWe define [A] and [B] as the instantaneous concentrations of these groups at any time t during\n"]], ["block_17", ["dx\nd? \n"]], ["block_18", ["[A10 + 1,12% = ([A]0 \u2014 [310) kcr\n1\n_______\n\u201c[Bn\u2014x\n[Ah\n"]], ["block_19", ["([B]0\n"]], ["block_20", ["1\n1\n\u201d\u2014 _ \u2014 : k\u00b0\n2.3.2\n[AllAb\n\u2019\n(\n)\n"]], ["block_21", ["1\n[AllBlo\nl\n= kc\n2.3.3\n[No [Blo n([A]0[B])\nI\n(\n)\n"]], ["block_22", ["'\u2014\nI) (\u201c([Blo\n- x) + ([No \u2014 x)\ndx\n) dz: ([Alo [1310) kc\n[Ab-x\nawe\u2014n?\n"]]], "page_66": [["block_0", [{"image_0": "66_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a plot for adipic acid reacted with 1,10-decamethylene glycol, catalyzed by p-toluene sulfonic\nacid. The reaction had already been run to consume 82% of the reactive groups before this\nexperiment was conducted. Interpreting the slope of the line in terms of Equation 2.3.6 and in\nthe light of actual initial concentrations gives a value of kc20.097 kg eq\u20141 min\u2018l. Note that\n"]], ["block_2", ["Figure 2.2\nPlot of l/(l\u2014p) versus time for the late stages of esterification of adipic acid with 1,10-\ndecamethylene glycol at 161\u00b0C, catalyzed by p-toluene sulfonic acid. The reaction time t 0 corresponds\nto a previous extent of reaction in which 82% of the COOH groups had been consumed. (Data from Hamann,\nS.D., Solomon, DH, and Swift, J.D., J. Macromol. Sci. Chem, A2, 153, 1968.)\n"]], ["block_3", ["At this point, itto recall the extent of reaction parameter, p, defined by Equation\n2.2.3. If we and Equation 2.3.4, we obtain\n\u201cFm\n(2'35)\n"]], ["block_4", ["01'\n"]], ["block_5", ["where we incorporated Equation 2.2.4 into the present discussion. These last expressions provide\ntwo very useful views of the progress of a condensation polymerization reaction with time.\nEquation 2.3.4 describes how the concentration of A groups asymptotically approaches zero at\nlong times; Equation 2.3.6 describes how the number average degree of polymerization increases\nlinearly with time.\nEquation 2.3.6 predicts a straight line when 1/(1\u2014p) is plotted against 1\u2018. Figure 2.2 shows such\n"]], ["block_6", ["rearrange Equation 2.3.2 to give the instantaneous concentrations of unreacted A groups as a\nfunction \n"]], ["block_7", ["Kinetics \n51\n"]], ["block_8", ["WHO)\n40\n"]], ["block_9", ["\u2014\u20141\u2014 :NIn 1 + kC[A]Ot\n(2.3.6)\n1 p\n"]], ["block_10", ["[A] \n[A10\n\u2014\n\u2014\u2014\u2014\u2014-\u20141\n+ kc[A]0t\n(2.3.4)\n"]], ["block_11", [{"image_1": "66_1.png", "coords": [48, 354, 290, 594], "fig_type": "figure"}]], ["block_12", ["1 00\n"]], ["block_13", ["80\n"]], ["block_14", ["20\n"]], ["block_15", ["60\n"]], ["block_16", ["0\nI\nI\nl\nI\nI\nI_I_\nI\nI\nI\nI_\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_17", ["0\n20\n40\n60\n30\n100\n120\n140\n"]], ["block_18", ["1\u2018 (min)\n"]]], "page_67": [["block_0", [{"image_0": "67_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["these units imply group concentrations expressed as equivalents per kilogram. Mass rather than\nvolume units are often used for concentration, as substantial volume changes may occur during\npolymerization.\n"]], ["block_2", ["a wide portion of the concentration change that occurs during a reaction. Each of these criteria\nseeks to maximize the region of fit, but the former emphasizes maximizing the range of while the\nlatter maximizes the range of p. Both standards tolerate deviations from their respective ideals at\nthe beginning or the end of the experiment. Deviations at the beginning of a process are\nrationalized in terms of experimental uncertainties at the point of mixing or modelistic difficulties\non attainment of stationary-state conditions.\nThe existence ofthese two different standards for success would be only of academic interest ifthe\nanalysis we have discussed applied to experimental results over most of the time range and over most\nextents of reactions as well. Unfortunately, this is not the case in all of the systems that have been\ninvestigated. Ref. [2], for example, shows one particular set of data\u2014adipic acid and diethylene\nglycol at 166\u00b0C, similar reactants as the system in Figure 2.2\u2014analyzed according to two different\nrate laws. This system obeys one rate law between p 0.50 and 0.85 that represents 15% of the\nduration of the experiment, and another rate law between p 0.80 and 0.93, which spans 45% of\nthe reaction time. These would be interpreted differently by the two standards above. This sort of\ndilemma is not unique to the present problem, but arises in many situations where one variable\nundergoes a large percentage of its total change while the other variable undergoes only a small\nfraction of its change. In the present context one way out of the dilemma is to take the view that only\nthe latter stages of the reaction are significant, as it is only beyond, say,p 0.80, that it makes sense to\nconsider the process as one of polymerization. Thus, it is only at large extents of reaction that\npolymeric products are formed and, hence, the kinetics of polymerization should be based on a\ndescription of this part of the process. This viewpoint intentionally focuses attention on a relatively\nmodest but definite range of p values. Since the reaction is necessarily slow as the number of\nunreacted functional groups decreases, this position tends to maximize the time over which the\nrate law fits the data. Calculation from the ordinate of Figure 2.2 shows that the data presented there\nrepresent only about the last 20% of the range of p values. The zero of the timescale has thus been\nshifted to pick up the analysis of the reaction at this point.\nWe commented above that the deviations at the beginning or the end of kinetic experiments can\nbe rationalized, although the different schools of thought would disagree as to what constitutes\n\u201cbeginning\u201d and \u201cend.\u201d Now that we have settled upon the polymer range, let us consider\nspecifically why deviations occur from a simple second-order kinetic analysis in the case of\ncatalyzed polymerizations. At the beginning of the experiment, say, up to p m 0.5, the concentra-\ntions of A and B groups change dramatically, even though the number average degree of\npolymerization has only changed from monomer to dimer. By ordinary polymeric standards, we\nare still dealing with a low molecular weight system that might be regarded as the solvent medium\nfor the formation of polymer. During this transformation, however, 50% of the very\u2014polar A groups\nand 50% of the very-polar B groups have been converted to the less\u2014polar ab groups. Thus, a\nsignificant change in the polarity of the polymerization medium occurs during the first half of the\nchange in p, even though an insignificant amount of true polymer has formed. In view of the role of\n"]], ["block_3", ["Although the results presented in Figure 2.2 appear to verify the predictions of Equation 2.3.6, this\nverification is not free from controversy. This controversy arises because various workers in this\nfield employ different criteria in evaluating the success of the relationships we have presented in\nfitting experimental polymerization data. One school of thought maintains that an adequate kinetic\ndescription of a process must apply to the data over a large part of the time of the experiment.\nA second point of view maintains that a rate law correctly describes a process when it applies over\n"]], ["block_4", ["2.3.2\nHow Should Experimental Data Be Compared with Theoretical Rate Laws?\n"]], ["block_5", ["52\nStep-Growth Polymerization\n"]]], "page_68": [["block_0", [{"image_0": "68_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "68_1.png", "coords": [17, 214, 270, 271], "fig_type": "molecule"}]], ["block_2", [{"image_2": "68_2.png", "coords": [34, 466, 155, 518], "fig_type": "figure"}]], ["block_3", ["on the assumption that A represents carboxyl groups. In this case, ku is the rate constant for the\nuncatalyzed reaction. This differential rate law is the equivalent of Equation 2.3.1 for the catalyzed\nreaction. Equation 2.3.7 is readily integrated when [A]0 = [B]0, in which case it becomes\n"]], ["block_4", ["Until now we have been discussing the kinetics of catalyzed reactions. Losses due to volatility and\nside reactions also raise questions as to the validity of assuming a constant concentration of\ncatalyst. Of course, one way of avoiding this issue is to omit an outside catalyst; reactions\ninvolving carboxylic acids can be catalyzed by these compounds themselves. Experiments\nconducted under these conditions are informative in their own right and not merely as a means\nof eliminating errors in the catalyzed case. As noted in connection with the discussion of Reaction\n(2G), the intermediate is stabilized by coordination with a proton from the catalyst. In the case of\nautOprotolysis by the carboxylic intermediate,\n"]], ["block_5", ["as this intermediate involves an additional equivalent of acid functional groups, the rate law for the\ndisappearance of A groups becomes\n"]], ["block_6", ["In order to achieve large p\u2019s and high molecular weights, it is essential that these equilibria be\nshifted to the right by removing the by-product molecule, water in these reactions. This may be\naccomplished by heating, imposing a partial vacuum, or purging with an inert gas, or some\ncombination of the three. These treatments also open up the possibility of reactant loss due to\nvolatility, which may accumulate to a significant source of error for reactions that are carried out to\nlarge values of p.\n"]], ["block_7", ["ionic ofmight \nwell in\ufb02uence ofreaction.\nAt the other end of the reaction, deviations from idealized rate laws are attributed to secondary\nreactions alcohols, amines through decarboxylation, dehydration,\nand deamination, respectively. The step-growth polymers that have been most widely studied are\nsimple condensation suchas polyesters and polyamides. Although we shall take up these\nclasses of polyamides\u2014specifically in Section 2.5 and Section 2.6,\nrespectively, it is appropriate to mention here that these are typically equilibrium reactions.\n"]], ["block_8", ["2.3.3\nUncatalyzed Step-Growth Reactions\n"]], ["block_9", ["and\n"]], ["block_10", ["Kinetics \n53\n"]], ["block_11", [{"image_3": "68_3.png", "coords": [40, 157, 251, 204], "fig_type": "molecule"}]], ["block_12", [{"image_4": "68_4.png", "coords": [43, 227, 256, 256], "fig_type": "molecule"}]], ["block_13", ["amJLOH + HO/Rzm \ufb01 ,. Rio/Hz + H20\n(2H)\n"]], ["block_14", ["d A\n__[_] __k[A]2[3]\n(2.3.7)\ndt\n"]], ["block_15", ["\u2014[__A]3\nd[A]\u2014\n(2.3.8)\n"]], ["block_16", ["4.\nOH\nR1\u2014l\u2014OH\n_ \n"]], ["block_17", ["o\no\n2\n"]], ["block_18", ["0...\nR2\n"]], ["block_19", [{"image_5": "68_5.png", "coords": [101, 475, 145, 513], "fig_type": "molecule"}]]], "page_69": [["block_0", [{"image_0": "69_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "69_1.png", "coords": [22, 27, 300, 295], "fig_type": "figure"}]], ["block_2", ["6i\n- 95.9 I\nO\nO\n0.1\nO\no\nc\n\u2018-\n.9\nX\n_\n*6\n"]], ["block_3", ["This integrates to\n"]], ["block_4", ["which shows that Nn increases more gradually with t than in the catalyzed case, all other things\nbeing equal.\nFigure 2.3 shows data for the uncatalyzed polymerization of adipic acid and 1,10~decamethylene\nglycol at 161\u00b0C plotted according to Equation 2.3.10. The various provisos of the catalyzed\ncase apply here also, so it continues to be appropriate to consider only the final stages of the\nconversion to polymer. From these results, ku is about 4.3 x 10\u20183 kg2 eq\u20182 min\u20181 at 161\u00b0C.\nWe conclude this section with a numerical example that serves to review and compare some of\nthe important relationships we have considered.\n"]], ["block_5", ["Thus for the uncatalyzed reaction, we have the following:\n"]], ["block_6", ["o.\nE\nI\nC\nCD\n3.:\n2\nCD\n0.\n\u2014 92.9\n2 ..\n"]], ["block_7", ["Assuming that 117C210\"1 kg eq\u2018l min\u2014l, ku=10_3 kg2 eq\u2018l min\"1, and [A]0:: 10 eq kg\u2018l,\ncalculate the time required for p to reach values 0.2, 0.4, 0.6, and so on, for both catalyzed and\n"]], ["block_8", ["Figure 2.3\nPlot of 1/(1\u201419)2 (left ordinate) and p (right ordinate) versus time for an uncatalyzed esterifi\u2014\ncation. (Data from Hamann, S.D., Solomon, DH, and Swift, J.D., J. Macromol. Sci. Chem, A2, 153, 1968.)\n"]], ["block_9", ["and this shows that a plot of (1\u2014,1'))_2 increases linearly with t.\n3.\nSince [A]/[A]O UN\u201c, Equation 2.3.10 becomes\n"]], ["block_10", ["Example 2.2\n"]], ["block_11", ["54\nStep-Growth Polymerization\n"]], ["block_12", ["1.\nThe rate law is third order.\nSince [A]/[A]O l\ufb02p, Equation 2.3.9 may be rewritten as\n"]], ["block_13", ["8 i-\n\u2014 96.5\n"]], ["block_14", ["N\ufb01 1 + 2ku[A]021\n(2.3.11)\n"]], ["block_15", ["(1 ~10)2\n"]], ["block_16", ["1\n1\n1\n1\nl\nJ\u2014 _L\n0\n200\n400\n600\n800\n1000\n1200\n1400\nr(min)\n"]], ["block_17", ["1\n1\n"]], ["block_18", ["1\n"]], ["block_19", ["\u2014[\nl\u2014\n\u2014i_\ni\nI\nl\nl\n"]], ["block_20", ["= 1 + 2ku[A]022\u2018\n(2.3.10)\n"]], ["block_21", ["\u2014 82.0\n"]]], "page_70": [["block_0", [{"image_0": "70_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "70_1.png", "coords": [0, 236, 339, 382], "fig_type": "figure"}]], ["block_2", ["p\n[A]/[A]0\nNn\nTime (min) catalyzed\nTime (min) uncatalyzed\n"]], ["block_3", ["1.\nEquation 2.2.4:\nNn 1/(1\u2014p).\n2.\nEquation 2.2.2:\n[A]/[A]0 l\u2014p.\n3.\nEquation 2.3.6:\nr=(1v,,\u20141)/k,[A]0 Nn\u2014l if catalyzed, since 10\u201c1(10) =1.\n4.\nEquation 2.3.11:\n2\u2018: (Nn2\u20141)/2 ku[A]02 (Nn2\u20141)(5) if uncatalyzed, since\n2(10\u20143) (10)2= 0.2.\n"]], ["block_4", ["uncatalyzed 2.3.2 and Equation 2.3.9, respectively, \n"]], ["block_5", ["In this section we turn our attention to two other questions raised in Section 2.2, namely, how do\nthe molecules distribute themselves among the different possible species, and how does this\ndistribution vary with the extent of reaction? Since a range of species is present at each stage of\npolymerization, it is apparent that a statistical answer is required for these questions. This time, our\nanswer begins, \u201cOn the average. . . .\u201d\nWe shall continue basing our discussion on the step-growth polymerization of the hypothetical\nmonomer AB. In Section 2.7, we shall take a second look at this problem for the case of unequal\nconcentrations of A and B groups. For now, however, we assure this equality by considering a\nmonomer that contains one group of each type. In a previous discussion of the polymer formed\nfrom this monomer, we noted that remnants of the original functional groups are still recognizable,\nalthough modified, along the backbone of the polymer chain. This state of affairs is emphasized by\nthe notation Ababa. . . abaB in which the as and b\u2019s of the ab linkages are groups of atoms carried\nover the initial A and B reactive groups. In this type of polymer molecule, then, there are i\u2014l a\u2019s\n"]], ["block_6", ["Using these relationships, the accompanying table is developed.\n"]], ["block_7", ["0.2\n0.8\n1.25\n0.25\n2.8\n0.4\n0.6\n1.67\n0.67\n8.9\n0.6\n0.4\n2.50\n1.5\n26\n0.8\n0.2\n5.00\n4.0\n120\n0.9\n0.1\n10.0\n9.0\n500\n0.95\n0.05\n20.0\n19\n2.0 x 103\n0.99\n0.01\n100\n99\n5.0 x 104\n0.992\n0.008\n120\n119\n7.2 x 104\n0.998\n0.002\n500\n499\n1.3 x 106\n"]], ["block_8", ["Since for specific values of \nsummarize the following relationships:\n"]], ["block_9", ["to the entire ofof\nand the fraction of unreacted A groups as a function of time.\n"]], ["block_10", ["A graphical comparison of the trends appearing here is presented in Figure 2.4. The importance\nof the catalyst is readily apparent in this hypothetical but not atypical system: To reach Nl1l 5\nrequires 4 min in the catalyzed case and 120 min without any catalyst, assuming that the\nappropriate rate law describes the entire reaction in each case.\n"]], ["block_11", ["Distribution of Molecular Sizes\n55\n"]], ["block_12", ["The question posed in Section 2.2\u2014how long will it take to reach a certain extent of reaction or\ndegree of polymerization?\u2014is now answered. As is often the case, the answer begins, \u201cIt all\ndepends. . .\n"]], ["block_13", ["Solution\n"]], ["block_14", ["2.4\nDistribution of Molecular Sizes\n"]]], "page_71": [["block_0", [{"image_0": "71_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "71_1.png", "coords": [22, 527, 180, 561], "fig_type": "molecule"}]], ["block_2", ["We now turn to the question of evaluating the fraction of i\u2014mers in a mixture as a function of\np. The fraction of molecules of a particular type in a population is just another way of describing\nthe probability of such a molecule. Hence our restated objective is to find the probability of an\ni\u2014mer in terms of p; we symbolize this quantity as the mole fraction x,(p). Since the i-mer consists\nof i\u2014l a\u2019s and 1 A, its probability is the same as the probability of finding i\u2014l a\u2019s and 1 A in the\nsame molecule. Recalling from Chapter 1 how such probabilities are compounded, we write\n"]], ["block_3", ["where pa and pA are the probabilities of individual a and A groups, respectively, and pa=p and\npA=1\u2014p. Equation 2.4.1 is known as the most probable distribution, and it arises in several\ncircumstances in polymer science, in particular free radical polymerization (see Chapter 3). The\nprobability of an i-mer can be converted to the number of i-mer molecules in the reaction mixture,\nm, by multiplying by the total number of molecules m in the mixture after the reaction has occurred\nto the extent p:\n"]], ["block_4", ["and 1 A if the degree of polymerization of the polymer is i. The a\u2019s differ from the A\u2019s precisely\nin that the former have undergone reaction whereas the latter have not. At any point during the\npolymerization reaction the fraction of the initial number of A groups that have reacted to become\na\u2019s is given by p, and the fraction that remains as A\u2019s is given by 1\u20141:). In these expressions p is\nthe same extent of reaction defined by Equation 2.2.3.\n"]], ["block_5", ["Figure 2.4\nComparison of catalyzed (solid lines) and uncatalyzed (dashed lines) polymerizations using\nresults calculated in Example 2.2. Here l\u2014p (left ordinate) and Nn (right ordinate) are plotted versus time.\n"]], ["block_6", ["2.4.1\nMole Fractions of Species\n"]], ["block_7", ["56\nStep-Growth Polymerization\n"]], ["block_8", ["2\n<1\n.3\n- 30\ni\nN\"\n"]], ["block_9", [{"image_2": "71_2.png", "coords": [45, 31, 269, 295], "fig_type": "figure"}]], ["block_10", ["x:(p) \ufb01lm pi\"(1 1))\n(2.4.1)\n"]], ["block_11", ["n,- m pi\u2014\u2018(l p)\n(2.4.2)\n"]], ["block_12", ["\\ \\ a\n\u201c20\n\u201ca ____. -...,_\n"]], ["block_13", ["\u2014 \u2014 \u2014 : Uncatalyzed\n\u2014: Catalyzed\n-40\n"]], ["block_14", ["150\n20\n250\n"]], ["block_15", ["_l\nI\n"]], ["block_16", ["- 50\n"]], ["block_17", ["- 10\n"]]], "page_72": [["block_0", [{"image_0": "72_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "72_1.png", "coords": [16, 421, 279, 641], "fig_type": "figure"}]], ["block_2", ["where m0 is present initially; m0: [A]0 \n"]], ["block_3", ["elect itanswer \nposed earlier.\nFigure 2.5 is a plot of the ratio fig/m versus1' for several values of p. Several features are apparent\nfrom Figure of\npresent:\n"]], ["block_4", ["Note that \n"]], ["block_5", ["Incorporating \n"]], ["block_6", ["result \n"]], ["block_7", ["seen before, contains\n"]], ["block_8", ["1.\nOn a number basis, the fraction of molecules always decreases with increasing regardless of\nthe value of p. The distributions in Table 2.1 are unrealistic in this regard.\n2.\nAs p increases, the proportion of molecules with smaller 1' values decreases and the proportion\nwith larger i values increases.\n"]], ["block_9", ["each multiplied by appropriate factor by Equation \n"]], ["block_10", ["Note that the upper limit of the second summation has been shifted from to 00 for mathematical\nreasons, namely that the answer is simple and known (see Appendix). The change is of little\npractical significance, since Equation 2.4.1 drops off for very large values of i. In particular, the\nresult derived in the Appendix is\n"]], ["block_11", ["003\nIlllll\u2014l'lIllllllllllllllllllll\n"]], ["block_12", ["The number average degree of polymerization for these mixtures is easily obtained by recalling\nthe de\ufb01nition of this average from Section 1.7. It is given by the sum of all possible i values, with\n"]], ["block_13", ["3.\nThe combination of effects described in item (2) tends to \ufb02atten the curves as p increases, but\nnot to the extent that the effect of item (1) disappears.\n"]], ["block_14", ["Distribution \n57\n"]], ["block_15", ["032\n"]], ["block_16", ["031\n"]], ["block_17", ["Figure 2.5\nMole fraction of i\u2014mer as a function of i for several values of p.\n"]], ["block_18", [{"image_2": "72_2.png", "coords": [40, 322, 202, 354], "fig_type": "molecule"}]], ["block_19", ["m p\u201c'(1 P)2m0\n(2.4.3)\n"]], ["block_20", ["mo\n00\n.\nNn Z 3x10?) Z arr\u2014la p)\n(2.4.4)\ni=1\ni=l\n"]], ["block_21", ["o\n100\nf\n200\n300\n"]], ["block_22", ["p=095\n"]], ["block_23", ["JllllllIlIIIIIIIIIIIIIIIIIJI\n"]]], "page_73": [["block_0", [{"image_0": "73_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "73_1.png", "coords": [29, 50, 138, 87], "fig_type": "molecule"}]], ["block_2", ["The weight fraction of i-mers is plotted as a function of i in Figure 2.6 for several large values of\np. Inspection of Figure 2.6 and comparison with Figure 2.5 reveals the following:\n"]], ["block_3", ["Next we turn our attention to the distribution of the molecules by weight among the various\nspecies. This will lead directly to the determination of the weight average molecular weight and the\nratio MW/Mn.\nWe begin by recognizing that the weight fraction w; of i\u2014mers in the polymer mixture at any\nvalue of p equals the ratio of the mass of i-mer in the mixture divided by the mass of the total\nmixture. The former is given by the product i rig-M0, where M0 is the molecular weight of the repeat\nunit; the latter is given by mOMO. Therefore we write\n"]], ["block_4", ["0.02\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\ni\n|\nI T\n|\n[TI\nI\u2014l\u2014I\nI\u2014I\u2014I\n'\n1\np 0.95\n"]], ["block_5", ["Of course, this is the same result that was obtained more simply in Equation 2.2.4. The earlier\nresult, however, was based on purely stoichiometric considerations and not on the detailed\ndistribution as is the present result.\n"]], ["block_6", ["i\n0.01 \n"]], ["block_7", ["into which Equation 2.4.3 may be substituted to give\n"]], ["block_8", ["Figure 2.6\nWeight fraction of i-mer as a function of i for several values of p.\n"]], ["block_9", ["Simplification of the summation in Equation 2.4.4 thus yields\n"]], ["block_10", ["2.4.2\nWeight Fractions of Species\n"]], ["block_11", ["58\nStep-Growth Polymerization\n"]], ["block_12", ["-\n1\nT\n\u201cl\nW;\n'\n-\n-\ni\n"]], ["block_13", [{"image_2": "73_2.png", "coords": [42, 431, 241, 640], "fig_type": "figure"}]], ["block_14", ["0\n"]], ["block_15", ["NI] \n(2.4.5)\n"]], ["block_16", ["w.- 53\n(2.4.6)\nm0\n"]], ["block_17", ["w.- ipirla p)2\n(2.4.7)\n"]], ["block_18", ["gm.\u2014 :(1\u2014p)2\n"]], ["block_19", ["0\n100\nI\nJ300\n"]], ["block_20", ["1\ni\np = 0.97\n_\ni\nr\np 0.99\n"]], ["block_21", ["-\n"]], ["block_22", ["I\nI\na\na\nI\nI\nI\nI\n|\ni\n-\n.\nI\n"]], ["block_23", ["p 0.995\n1\n"]]], "page_74": [["block_0", [{"image_0": "74_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "74_1.png", "coords": [25, 184, 221, 259], "fig_type": "figure"}]], ["block_2", [{"image_2": "74_2.png", "coords": [27, 285, 143, 324], "fig_type": "molecule"}]], ["block_3", [{"image_3": "74_3.png", "coords": [34, 496, 288, 681], "fig_type": "figure"}]], ["block_4", ["P\nNn\nNW\nNw/Nn\n"]], ["block_5", ["The weight by averaging the contributions \n"]], ["block_6", ["where the upper limit on i has been extended to infinity as before. The new summation that we\nneed is also evaluated in Appendix:\n"]], ["block_7", ["Table 2.2\nValues of N\u201c, NW, and NW/Nn for Various Large\nValues of p\n"]], ["block_8", ["which is the desired result.\nWe saw in Chapter 1 that the ratio Mw/Mn, or polydispersity index, is widely used in polymer\nchemistry as a measure of the width of a molecular weight distribution. If the effect of chain ends is\ndisregarded, this ratio is the same as the corresponding ratio of i values:\n"]], ["block_9", ["where the ratio of Equation 2.4.9 to Equation 2.2.4 has been used. Table 2.2 lists values of NW, N\u201c,\n"]], ["block_10", ["Using this in Equation 2.4.8 gives\n"]], ["block_11", ["0.90\n10.0\n19.0\n1.90\n0.92\n12.5\n24.0\n1.92\n0.94\n16.7\n32.3\n1.94\n0.96\n25.0\n49.0\n1.96\n0.98\n50.0\n99.0\n1.98\n0.990\n100\n199\n1 .990\n0.992\n125\n249\n1.992\n0.994\n167\n332\n1.994\n0.996\n250\n499\n1.996\n0.998\n500\n999\n1.998\n"]], ["block_12", ["1,\nAt any p, very small and very large values of i contribute a lower weight fraction to the\nmixture than do intermediate values of i. This arises because of the product in,- in Equation\n2.4.6: n; is large for monomers, in which case i is low, and then n,- decreases as i increases.\nAt intermediate values of i, w,- goes through a maximum.\n2,\nAs p increases, the maximum in the curves shifts to larger i values and the tail of the curve\nextends to higher values of i.\n"]], ["block_13", ["Distribution \n59\n"]], ["block_14", ["1' values using in the averaging procedure:\n"]], ["block_15", ["3.\nThe effect in item (2) is not merely a matter of shifting curves toward higher i values as p\nincreases, but re\ufb02ects a distinct broadening of the distribution of i values as p increases.\n"]], ["block_16", ["and NW/Nn for a range of high p values. Note that NW/Nn \u2014> 2 as p \u2014+ 1; this is a characteristic result\n"]], ["block_17", [{"image_4": "74_4.png", "coords": [42, 529, 265, 648], "fig_type": "table"}]], ["block_18", [{"image_5": "74_5.png", "coords": [43, 187, 203, 255], "fig_type": "molecule"}]], ["block_19", ["mo\n00\nZiw;\n2310910 -19)2\n_ \n{:1\nNw\n(2.4.8)\n00\nin\u00bb.\nZipi\u2014lu \u2014p)2\n"]], ["block_20", ["MW\nNW\n\u2014\u2014 = \u2014 = 1 +\n2.4.10\nM,\nNn\n\u20180\n(\n)\n"]], ["block_21", ["_1+p\nNW\u2014 l\u2014p\n(2.4.9)\n"]], ["block_22", ["0\u00b0.\n,-_\n1+1)\n:12? \n(lvp)\ni=1\n"]], ["block_23", ["121\ni=1\n"]]], "page_75": [["block_0", [{"image_0": "75_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We have not attempted to indicate the conditions of temperature, catalyst, solvent, and so on, for\nthese various reactions. For this type of information, references that deal specifically with synthetic\npolymer chemistry should be consulted. In the next few paragraphs we shall comment on\nthe various routes to polyester formation in the order summarized above and followed in Table\n2.3. The studies summarized in Figure 2.2 and Figure 2.3 are examples of Reaction 2 in Table 2.3.\n"]], ["block_2", ["of the most probable distribution. In light of Equation 1.7.16, the standard deviation of the\nmolecular distribution is equal to Mn for the polymer sample produced by this polymerization.\nIn a manner of speaking, the molecular weight distribution is as wide as the average is high. The\nbroadening of the distribution with increasing p is dramatically shown by comparing the values\nin Table 2.2 with the situation at a low p value, say p205. At p205, Nn=2, Nw=3, and\nNW/Nn 1.5.\nSince Equation 2.3.5 and Equation 2.3.11 give p as a function of time for the catalyzed and\nuncatalyzed polymerizations, respectively, the distributions discussed in the last few paragraphs\ncan also be expressed with time as the independent variable instead of p.\nThe results we have obtained on the basis of the hypothetical monomer AB are also applicable\nto polymerizations between monomers of the AA and BB type, as long as the condition [A] [B]\nis maintained. In Section 2.7, we shall extend the arguments of this section to conditions in which\n[A] 79 [B]. In the meanwhile, we interrupt this line of reasoning by considering a few particular\ncondensation polymers as examples of step-growth systems. The actual systems we discuss will\nserve both to verify and reveal the limitations of the concepts we have been discussing. In addition,\nthey point out some of the topics that still need clarification. We anticipate some of the latter points\nby noting the following:\n"]], ["block_3", ["Polyesters and polyamides are two of the most-studied step-growth polymers, as well as\nsubstances of great commercial importance. We shall consider polyesters in Section 2.5 and\npolyamides in Section 2.6.\n"]], ["block_4", ["The preceding discussions of the kinetics and molecular weight distributions in the step-growth\npolymerizations of AB monomers are exemplified by esterification reactions between such\nmonomers as glycolic acid and w\u2014hydroxydecanoic acid. Therefore one method of polyester\nsynthesis is the following:\n"]], ["block_5", ["Several other chemical reactions are also widely used for the synthesis of these polymers. This\nlist enumerates some of the possibilities, and Table 2.3 illustrates these reactions by schematic\nchemical equations.\n"]], ["block_6", ["Esteri\ufb01cation of a diacid and a diol\nEster interchange with alcohol\nEster interchange with ester\nEsterification of acid chlorides\nLactone polymerization\n9\u201899?!\u201d\n"]], ["block_7", ["2.5\nPolyesters\n"]], ["block_8", ["1.\nWhen [A] 7\u00e9 [B], both ends of the growing chain tend to be terminated by the group that is\npresent in excess. Subsequent reaction of such a molecule involves reaction with the limiting\ngroup. The effect is a decrease in the maximum attainable degree of polymerization.\n2.\nWhen a monofunctional reactant is present\u2014-one containing a single A or B group\u2014the effect\nis also clearly a decrease in the average degree of polymerization. It is precisely because this\ntype of reactant can only react once that it is sometimes introduced into polymer formulations,\nthereby eliminating the possibility of long-term combination of chain ends, and/or restricting\nthe average molecular weight.\n"]], ["block_9", ["60\nStep-Growth Polymerization\n"]], ["block_10", ["1.\nEsterification of a hydroxycarboxylic acid\n"]]], "page_76": [["block_0", [{"image_0": "76_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["While up to now we have emphasized bifunctional reactants, both monofunctional compounds and\nmonomers with functionality greater than 2 are present in some polymerization processes, either\nintentionally or adventitiously. The effect of the monofunctional reactant is clearly to limit chain\ngrowth. As noted above, a functionality greater than 2 results in branching. A type of polyester that\nincludes mono-, di-, and trifunctional monomers is the so-called alkyd resin. A typical example is\nbased on the polymerization of phthalic acid (or anhydride), glycerol, and an unsaturated mono-\ncarboxylic acid. The following suggests the structure of a portion of such a polyester:\n0 \n0 \n\u201d\u20194.\nHo/Y\\0H + H0\n+ H0\nm \u2014..\nO\n2.J\nOH\nI\ng/\\O(\\O \n(\n)\n0\n\ufb02\n"]], ["block_2", ["Table 2.3\nSome Schematic Reactions for the Formation of Polyesters\n"]], ["block_3", ["2. Esterification \n"]], ["block_4", ["6. Lactone polymerization:\n"]], ["block_5", ["4. Ester interchange with ester (\u201ctransesterification\u201d):\n"]], ["block_6", ["3. Ester \n"]], ["block_7", ["5. Esterification of acid chlorides (Schotten-Baumann reaction):\n0 \u00b0\n9 it \nR1\n1\nn HO/\n\\OH\n+\nCIJLRELQ =\n{0/\n\\0\nHrs/\u201c9n +\n2n HCI\n"]], ["block_8", ["Polyesters\n51\n"]], ["block_9", ["1. Esterification \n"]], ["block_10", [{"image_1": "76_1.png", "coords": [40, 588, 255, 643], "fig_type": "molecule"}]], ["block_11", [{"image_2": "76_2.png", "coords": [45, 435, 260, 488], "fig_type": "molecule"}]], ["block_12", [{"image_3": "76_3.png", "coords": [71, 89, 324, 138], "fig_type": "figure"}]], ["block_13", [{"image_4": "76_4.png", "coords": [76, 231, 328, 281], "fig_type": "molecule"}]], ["block_14", [{"image_5": "76_5.png", "coords": [81, 442, 245, 473], "fig_type": "molecule"}]], ["block_15", [{"image_6": "76_6.png", "coords": [87, 93, 321, 142], "fig_type": "molecule"}]], ["block_16", [{"image_7": "76_7.png", "coords": [87, 302, 351, 354], "fig_type": "molecule"}]], ["block_17", [{"image_8": "76_8.png", "coords": [89, 303, 209, 344], "fig_type": "molecule"}]], ["block_18", [{"image_9": "76_9.png", "coords": [95, 372, 274, 412], "fig_type": "molecule"}]], ["block_19", ["n HO/ 1\\OH\n+\nn HOJLHQJLOH\n{0/ \\D/ILHQ/K)\u201d 20 \n"]], ["block_20", ["0\nO\nR1\non2\n+\nFia\u2014OH\nR1/L|\\OR3\n+\nR2_0H\n"]], ["block_21", ["R1/U\\OR2\n+\nR3/U\\0Fi4\n\u2014_'\"\nR1/U\\0Fi4\nRa/LkOFiz\n"]], ["block_22", [{"image_10": "76_10.png", "coords": [96, 167, 326, 204], "fig_type": "molecule"}]], ["block_23", ["O\nn\n"]], ["block_24", ["Ft\n0\n"]], ["block_25", ["O\n0\nO\n"]], ["block_26", ["H\no\n0\nH1\n0\no\n"]], ["block_27", [{"image_11": "76_11.png", "coords": [144, 589, 368, 672], "fig_type": "molecule"}]], ["block_28", ["0\nO\n|\n"]], ["block_29", [{"image_12": "76_12.png", "coords": [144, 434, 235, 483], "fig_type": "molecule"}]], ["block_30", [{"image_13": "76_13.png", "coords": [147, 449, 225, 476], "fig_type": "molecule"}]], ["block_31", [{"image_14": "76_14.png", "coords": [150, 235, 314, 273], "fig_type": "molecule"}]], ["block_32", [{"image_15": "76_15.png", "coords": [194, 304, 339, 348], "fig_type": "molecule"}]], ["block_33", [{"image_16": "76_16.png", "coords": [202, 165, 415, 206], "fig_type": "molecule"}]], ["block_34", [{"image_17": "76_17.png", "coords": [222, 368, 403, 420], "fig_type": "molecule"}]], ["block_35", [{"image_18": "76_18.png", "coords": [235, 378, 400, 405], "fig_type": "molecule"}]]], "page_77": [["block_0", [{"image_0": "77_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["It has been hypothesized that cross\u2014linked polymers would have better mechanical properties if\ninterchain bridges were located at the ends rather than the center of chains. To test this, low\n"]], ["block_2", ["The ethylene glycol liberated by Reaction (2L) is removed by lowering the pressure or purging\nwith an inert gas. Because the ethylene glycol produced by Reaction (2L) is removed, proper\nstoichiometry is assured by proceeding via the intermediate bis(2\u2014hydroxyethyl) terephthalate;\notherwise the excess glycol used initially would have a deleterious effect on the degree of\npolymerization. Poly(ethylene terephthalate) is more familiar by some of its trade names: Mylar\nas a film and Dacron, Kodel, or Terylene as fibers; it is also known by the acronym PET.\nEster interchange reactions like that shown in Reaction 4 in Table 2.3 (transesterification) can\nbe carried out on polyesters themselves to produce a scrambling between the two polymers.\nStudies of this sort between high and low molecular weight prepolymers result in a single polymer\nwith the same molecular weight distribution as would have been obtained from a similarly\nconstituted diol\u2014diacid mixture by direct polymerization. This is true when the time-catalyst\nconditions allow the randomization to reach equilibrium. If the two prepolymers are polyesters\nformed from different monomers, the product of the ester interchange reaction will be a copolymer\nof some sort. If the reaction conditions favor esterification, the two chains will merely link together\nand a block copolymer results. If the conditions favor the ester interchange reaction, then a\nscrambled 00polymer molecule results. These possibilities underscore the idea that the derivations\nof the preceding sections are based on complete equilibrium among all molecular species present\nduring the condensation reaction.\n"]], ["block_3", ["The presence of the unsaturated substituent along this polyester backbone gives this polymer cross-\nlinking possibilities through a secondary reaction of the double bond. These polymers are used in\npaints, varnishes, and lacquers, where the ultimate cross-linked product results from the oxidation of\nthe double bond as the coating cures. A cross-linked polyester could also result from Reaction (2.1)\nwithout the unsaturated carboxylic acid, but the latter would produce a gel in which the entire\nreaction mass solidified, and is therefore not as well suited to coating applications as a polymer that\ncross-links upon \u201cdrying.\u201d\nMany of the reactions listed at the beginning of this section are acid catalyzed, although a number\nof basic catalysts are also employed. Esterifications are equilibrium reactions, and often carried out\nat elevated temperatures for favorable rate and equilibrium constants and to shift the equilibrium in\nfavor of the polymer by volatilization of the by-product molecules. An undesired feature of higher\npolymerization temperatures is the increased possibility of side reactions, such as the dehydration of\nthe diol or the pyrolysis of the ester. Basic catalysts tend to produce fewer undesirable side reactions.\nEster exchange reactions are valuable, since, say, methyl esters of dicarboxylic acids are often\nmore soluble and easier to purify than the diacid itself. The methanol by-product is easily removed\nby evaporation. Poly(ethylene terephthalate) is an example of a polymer prepared by double\napplication of Reaction 4 in Table 2.3. The first stage of the reaction is conducted at temperatures\nbelow 200\u00b0C and involves the interchange of dimethyl terephthalate with ethylene glycol.\n"]], ["block_4", ["The rate of this reaction is increased by using excess ethylene glycol, and removal of the methanol\nis assured by the elevated temperature. Polymer is produced in the second stage after the\ntemperature is raised above the melting point of the polymer, about 260\u00b0C.\n"]], ["block_5", ["Example 2.3\n"]], ["block_6", [{"image_1": "77_1.png", "coords": [35, 278, 239, 317], "fig_type": "molecule"}]], ["block_7", ["62\nStep-Growth Polymerization\n"]], ["block_8", [{"image_2": "77_2.png", "coords": [39, 373, 368, 417], "fig_type": "molecule"}]], ["block_9", ["Me\u2014O\nO\u2014Me\n"]], ["block_10", ["0\no\no\no\nn HO_\\_0>_\u00a9_n\n"]], ["block_11", ["0\nO\nO\nO\nW\n+\n2 \n+ 2 MeOH\n"]], ["block_12", [{"image_3": "77_3.png", "coords": [190, 274, 428, 321], "fig_type": "molecule"}]], ["block_13", [{"image_4": "77_4.png", "coords": [296, 375, 439, 411], "fig_type": "molecule"}]], ["block_14", ["(2K)\n"]]], "page_78": [["block_0", [{"image_0": "78_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "78_1.png", "coords": [15, 98, 411, 264], "fig_type": "figure"}]], ["block_2", ["-\u2014_\u2014\n"]], ["block_3", ["On the basis of these facts, do the following:\n"]], ["block_4", ["2.\n3.\n"]], ["block_5", ["molecular weight polyesters were synthesized from a diol and two different diacids: one saturated\n"]], ["block_6", ["Endene\nCentrene\nStep 1:\n8 h at about 150\u00b0C\u2014200\u00b0C\nMaleic anhydride (mol)\n0\n2.0\nSuccinic anhydride (mol)\n2.0\n0\nDiethylene glycol (mol)\n3.0\n3.0\nStep 2:\nAbout 1/2 h at about 120\u00b0C\u2014130\u00b0C\nMaleic anhydride (mol)\n2.0\n0\nSuccinic anhydride (mol)\n0\n2.0\nCatalyst\n0\n0\nStep 3:\n30% Styrene + catalyst\n16 h at 55\u00b0C -l- 1 h at 110\u00b0C\n"]], ["block_7", ["located \n"]], ["block_8", ["and the other unsaturated. The synthetic procedure was such that the unsaturated acid units were\n"]], ["block_9", ["aspects of the overall experiment are listed below:\n"]], ["block_10", ["Polyesters\n63\n"]], ["block_11", ["Solution\n"]], ["block_12", ["1.\nSince the reaction conditions are mild in step 2 (only 6% as much time allowed as in step 1 at a\nlower temperature) and no catalyst is present, it is unlikely that any significant amount of ester\nscrambling occurs. Isomerization of maleate to fumarate is also known to be insigni\ufb01cant\nunder these conditions.\nThe idealized structures of these molecules are\n"]], ["block_13", ["1.\n"]], ["block_14", [{"image_2": "78_2.png", "coords": [39, 382, 284, 601], "fig_type": "figure"}]], ["block_15", [{"image_3": "78_3.png", "coords": [48, 504, 285, 554], "fig_type": "molecule"}]], ["block_16", [{"image_4": "78_4.png", "coords": [50, 516, 262, 598], "fig_type": "molecule"}]], ["block_17", [{"image_5": "78_5.png", "coords": [51, 413, 282, 499], "fig_type": "figure"}]], ["block_18", ["Comment on the likelihood that the comonomers are segregated as the names of these\npolymers suggest.\nSketch the structure of the average endene and centrene molecules.\nComment on the results in terms of the initial hypothesis.\n"]], ["block_19", ["Endene\n"]], ["block_20", ["A cross-linked product with unsaturation at the chain ends does, indeed, have a higher modulus.\nThis could be of commercial importance and indicates that industrial products might be formed\nby a nonequilibrium process precisely for this sort of reason. A fuller discussion of the factors\nthat contribute to the modulus will be given in Chapter 10 and Chapter 12.\n"]], ["block_21", ["Centrene\n"]], ["block_22", ["O\nO\nO\nO\nl\\\\u2018\nHOJJWOwOA/ONo/VO\n"]], ["block_23", ["O\nO\nO\nO\n\\\nHowowom/ONONO\n"]], ["block_24", [{"image_6": "78_6.png", "coords": [62, 412, 279, 495], "fig_type": "molecule"}]], ["block_25", [{"image_7": "78_7.png", "coords": [65, 418, 271, 458], "fig_type": "molecule"}]], ["block_26", ["HO\n"]], ["block_27", ["Elastic modulus (Pa)\n21,550\n16,500\n"]], ["block_28", [{"image_8": "78_8.png", "coords": [89, 107, 307, 233], "fig_type": "figure"}]], ["block_29", ["O\nO\n0\nO\n"]], ["block_30", ["0\nO\n"]], ["block_31", ["O\nO\n"]], ["block_32", ["O\nO\n"]]], "page_79": [["block_0", [{"image_0": "79_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Acid chlorides are generally more reactive than the parent acids, so polyester fonnation via\nReaction 5 in Table 2.3 can be carried out in solution and at lower temperatures, in contrast with\nthe bulk reactions of the melt as described above. Again, the by-product molecules must be\neliminated either by distillation or precipitation. The method of interfacial condensation, described\nin Section 2.6, can be applied to this type of reaction.\nThe formation of polyesters from the polymerization of lactones (Reaction 6 in Table 2.3) is a\nring\u2014opening reaction that may follow either a step-growth or chain mechanism, depending on\nconditions. For now our only concern is to note that the equilibrium representing this reaction in\nTable 2.3 describes polymerization by the forward reaction and ring formation by the back reaction.\nRings clearly compete with polymers for monomer in all polymerizations. Throughout the chapter\nwe have assumed that all competing side reactions, including ring formation, could be neglected.\n"]], ["block_2", ["Table 2.4\nSome Schematic Reactions for the Formation of Polyamides\n"]], ["block_3", ["64\nStep-Growth Polymerization\n"]], ["block_4", ["Additional synthetic routes that closely resemble polyesters are also available. Several more of\nthese are listed below and are illustrated by schematic reactions in Table 2.4:\n"]], ["block_5", ["The discussion of polyamides parallels that of polyesters in many ways. To begin with, polyamides\nmay be fonned from an AB monomer, in this case amino acids:\n"]], ["block_6", ["2.6\nPolyamides\n"]], ["block_7", ["1.\nAmidation of amino acids\n"]], ["block_8", ["4. Amidation of acid chlorides:\n"]], ["block_9", ["3. Interchange reactions:\n"]], ["block_10", ["1. Amidation of amino acids:\nj:\n..\nO\n"]], ["block_11", ["2. Amidation of diamine and diacid:\n"]], ["block_12", ["5. Lactam polymerization:\n"]], ["block_13", [{"image_1": "79_1.png", "coords": [58, 605, 264, 656], "fig_type": "molecule"}]], ["block_14", [{"image_2": "79_2.png", "coords": [80, 335, 312, 397], "fig_type": "figure"}]], ["block_15", [{"image_3": "79_3.png", "coords": [102, 613, 252, 646], "fig_type": "molecule"}]], ["block_16", [{"image_4": "79_4.png", "coords": [102, 338, 313, 393], "fig_type": "molecule"}]], ["block_17", ["n\nH2N._R\nOH\n\u2014'__\u2018_'\u2014\"\"-\n\u2018(N\u2018R/Ugn\n4..\n\u201dH20\n"]], ["block_18", ["n H2N\u2019R1\u201cNH2 \nn HOJ-LRZJkOH = {HERE/Lag\u201c H20\n"]], ["block_19", [".Fi\n.\nn H2N\n1\"NH2\n4'\nn CIJLRELCI\n\u2014--'--...\u2014\u2014\n{NRLH\nR\ufb01n+\n2n HCI\n"]], ["block_20", [{"image_5": "79_5.png", "coords": [113, 543, 312, 582], "fig_type": "molecule"}]], ["block_21", [{"image_6": "79_6.png", "coords": [119, 477, 394, 523], "fig_type": "molecule"}]], ["block_22", ["R1/LSHR2\n+\nPlain/R4\nmin/R4\n+\nF\u2018s/\u201c\\H\u2019Fb\n"]], ["block_23", [{"image_7": "79_7.png", "coords": [137, 346, 303, 379], "fig_type": "molecule"}]], ["block_24", ["0\nO\nO\nO\n"]], ["block_25", ["O\n0\nj:\n0\n"]], ["block_26", ["O\nO\nO\nO\n"]], ["block_27", [{"image_8": "79_8.png", "coords": [217, 474, 378, 514], "fig_type": "molecule"}]], ["block_28", [{"image_9": "79_9.png", "coords": [224, 411, 420, 452], "fig_type": "molecule"}]], ["block_29", [{"image_10": "79_10.png", "coords": [233, 543, 417, 583], "fig_type": "molecule"}]]], "page_80": [["block_0", [{"image_0": "80_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "80_1.png", "coords": [21, 356, 243, 413], "fig_type": "molecule"}]], ["block_2", ["Reaction (2.N) describes the nylon salt $ nylon equilibrium. Reaction (2.0) and Reaction (2.P)\nshow proton transfer with water between carboxyl and amine groups. Since proton transfer\nequilibria are involved, the self\u2014ionization of water, Reaction (2.Q), must also be included.\nEspecially in the presence of acidic catalysts, Reaction (2R) and Reaction (2.8) are the equilibria\nof the acid-catalyzed intermediate described in general in Reaction (2.G). The main point in\nincluding all of these equilibria is to indicate that the precise concentration of A and B groups in\n"]], ["block_3", ["compounds first number after the name gives the number ofatoms in\nthe diamine, of carbons in the diacid.\nThe diacid\u2014diamine amidation described in Reaction 2 in Table 2.4 has been widely studied in\nthe melt, solution, and state. When equal amounts of two functional groups are present,\nboth the rate laws and weight distributions are given by the treatment of the\npreceding sections. The stoichiometric balance between reactive groups is readily obtained\nby precipitating the 1:1 ammonium salt from ethanol:\n"]], ["block_4", ["We only need to recall the trade name of synthetic polyamides, nylon, to recognize the importance\nof these to prepare them. Recall the System naming these\n"]], ["block_5", ["This compound is sometimes called a nylon salt. The salt :3 polymer equilibrium is more\nfavorable to the production of polymer than in the case of polyesters, so this reaction is often\ncarried out in a sealed tube or autoclave at about 200\u00b0C until a fairly high extent of reaction is\nreached; then the temperature is raised and the water driven off to attain high molecular weight\npolymer. Also in contrast to polyesters, Reaction 1 and Reaction 2 in Table 2.4 can be conducted\nrapidly without an acid catalyst.\nThe process represented by Reaction 2 in Table 2.4 actually entails a number of additional\nequilibrium reactions. Some of the equilibria that have been considered include the following:\n"]], ["block_6", ["a diacid-diamine reaction mixture is a complicated function of the moisture content and the pH, as\nwell as the initial amounts of reactants introduced. Because of the high affinity for water of the\n"]], ["block_7", ["Amidation of a diacid and a diamine\nInterchange reactions\nAmidation of acid chlorides\nLactam polymerization\nP\u2018P\u2018E\u2019JP\n"]], ["block_8", ["Polyamides\n65\n"]], ["block_9", ["r\"JQOH\n+ HZNM\n~\u2014.\u2014\u2014\u2014\u2014-=\nALNJ\u2018\"\n+ H30+\n(Tl-S)\nH\n"]], ["block_10", ["H\u2018JLo'\n+\nHal-5M\n\u201cAw\u2014~\u2014\nFJLNF\n+\nH2O\n(2N)\nH\n"]], ["block_11", ["0\n0\n2.0\nH\u2018JLOH\n+\nH20\n._..__.._a...._\nrpi\u2018H\\O-\n+\nH30+\n"]], ["block_12", ["2 H20 \ufb02\nH30+\n+ \u2018OH\n(2Q)\n"]], ["block_13", ["WNH2\n+\nH20 \nnHs , -0H\n(2.P)\n"]], ["block_14", ["0\nO\nO\nO\nR\nHQN/ 1\\NH2\n+ HO/lLRQ/LkOH\n\u2014\u2014-\u2014\u2014r-\nHera/RKIQHC,\n+ -0J\\R{LL0\u2014\n(2M)\n"]], ["block_15", ["o\no\n"]], ["block_16", ["OH\n0\n"]], ["block_17", ["0\nOH\n(2R)\n"]], ["block_18", [{"image_2": "80_2.png", "coords": [54, 407, 227, 442], "fig_type": "molecule"}]], ["block_19", [{"image_3": "80_3.png", "coords": [62, 221, 187, 256], "fig_type": "molecule"}]], ["block_20", [{"image_4": "80_4.png", "coords": [77, 524, 234, 558], "fig_type": "molecule"}]], ["block_21", [{"image_5": "80_5.png", "coords": [78, 365, 234, 402], "fig_type": "molecule"}]], ["block_22", [{"image_6": "80_6.png", "coords": [113, 218, 373, 264], "fig_type": "molecule"}]], ["block_23", [{"image_7": "80_7.png", "coords": [115, 221, 177, 256], "fig_type": "molecule"}]], ["block_24", [{"image_8": "80_8.png", "coords": [237, 220, 356, 258], "fig_type": "molecule"}]], ["block_25", ["(\n)\n"]]], "page_81": [["block_0", [{"image_0": "81_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "81_1.png", "coords": [21, 399, 191, 635], "fig_type": "figure"}]], ["block_2", ["Figure 2.7\nSketch of an interfacial polymerization with the collapsed polymer film being withdrawn from\nthe surface between the immiscible phases. (From Morgan, P.W. and Kwolek, S.L. J. Chem. Educ, 36, 182,\n1959. With permission.)\n"]], ["block_3", ["various functional groups present, the complete removal of water is impossible: the equilibrium\nmoisture content of molten nylon-6,6 at 290\u00b0C under steam at\n1 atm is 0.15%. Likewise, the\nvarious ionic possibilities mean that at both high and low pH values the concentration of unionized\ncarboxyl or amine groups may be considerably different from the total concentration\u2014without\nregard to state of ionization\u2014of these groups. As usual, upsetting the stoichiometric balance of the\nreactive groups lowers the degree of polymerization attainable. The abundance of high-quality\nnylon products is evidence that these complications have been overcome in practice.\nAmide interchange reactions of the type represented by Reaction 3 in Table 2.4 are known to\noccur more slowly than direct amidation; nevertheless, reactions between high and low molecular\nweight polyamides result in a polymer of intermediate molecular weight. The polymer is initially a\nblock copolymer of the two starting materials, but randomization is eventually attained.\nAs with polyesters, the amidation reaction of acid chlorides may be carried out in solution because\nof the enhanced reactivity of acid chlorides compared with carboxylic acids. A technique known as\ninterfacial polymerization has been employed for the formation of polyamides and other step-growth\npolymers, including polyesters, polyurethanes, and polycarbonates. In this method, the polymeriza-\ntion is carried out at the interface between two immiscible solutions, one of which contains one of the\ndissolved reactants, while the second monomer is dissolved in the other. Figure 2.7 shows a\npolyamide film forming at the interface between layers of an aqueous diamine solution and a solution\nof diacid chloride in an organic solvent. In this form, interfacial polymerization is part of the standard\nrepertoire of chemical demonstrations. It is sometimes called the \u201cnylon rope trick\u201d because of the\nfilament of nylon that can be produced by withdrawing the collapsed film.\nThe amidation of the reactive groups in interfacial polymerization is governed by the rates at\nwhich these groups can diffuse to the interface where the growing polymer is deposited. Accord-\ningly, new reactants add to existing chains rather than interacting to form new chains. This is\n"]], ["block_4", ["66\nStep-Growth Polymerization\n"]], ["block_5", ["4-1\u201c- Collapsed\n"]], ["block_6", ["at interface\n"]]], "page_82": [["block_0", [{"image_0": "82_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "82_1.png", "coords": [34, 227, 154, 268], "fig_type": "molecule"}]], ["block_2", ["By definition of the problem, this ratio cannot exceed unity.\n"]], ["block_3", ["We now turn to one of the problems we have sidestepped until now\u2014the polymerization of\nreactants in which a stoichiometric imbalance exists in the numbers of reactive groups A and B.\nIn earlier sections dealing with the quantitative aspects of step-growth polymerization, we focused\nattention on monomers of the AB type to assure equality of reactive groups. The results obtained\nabove also apply to AA and BB polymerizations, provided that the numbers of reactive groups are\nequal. There are obvious practical difficulties associated with the requirement of stoichiometric\nbalance. Rigorous puri\ufb01cation of monomers is difficult and adds to the cost of the final product.\nThe effective loss of functional groups to side reactions imposes restrictions on the range of\nexperimental conditions at best and is unavoidable at worst. These latter considerations apply even\nin the case of the AB monomer. We have already stated that the effect of the imbalance of A and\nB groups is to lower the eventual degree of polymerization of the product. A quantitative\nassessment of this limitation is what we now seek.\nWe define the problem by assuming that the polymerization involves AA and BB monomers\nand that the B groups are present in excess. We define 12A and V3 to be the numbers of A and B\nfunctional groups, respectively. The number of either of these quantities in the initial reaction\nmixture is indicated by a superscript o; the numbers at various stages of reaction have no\nsuperscript. The stoichiometric imbalance is defined by the ratio r, where\n"]], ["block_4", ["The combination of strong intermolecular forces and high chain-stiffness accounts for the high\nmelting points of polyamides (see Chapter 13). The remarks of this section and Section 2.5\nrepresent only a small fraction of what could be said about these important materials. We have\ncommented on aspects of the polymerization processes and of the polymers themselves that have a\ndirect bearing on the concepts discussed throughout this volume. This material provides an\nexcellent example of the symbiosis between theoretical and application-oriented viewpoints.\nEach stimulates and reinforces the other with new challenges, although it must be conceded that\nmany industrial processes reach a fairly high degree of empirical refinement before the conceptual\nbasis is quantitatively deve10ped.\n"]], ["block_5", ["different elsewhere in it\nis evident that polymer should result from this difference. The HCl\nby-product of the amidation reaction is neutralized by also dissolving an inorganic base in the\naqueous layer The choice of the organic solvent plays a role in\ndetermining \nquality for Since this reaction is carried out at low temperatures, the\ncomplications be kept to a minimum. yield\nincreased between the two solutions by stirring.\nLactam polymerization represented by Reaction 5 in Table 2.4 is another example of a ring-\nopening reaction, the reverse of which is a possible competitor with polymer for reactants. The\nvarious mechanical properties of polyamides may be traced in many instances to the possibility of\nintermolecular between the polymer molecules, and to the relatively stiff chains\nthese substances possess. latter, in turn, may be understood by considering still another\nequilibrium, along the chain backbone:\n"]], ["block_6", ["2.7\nStoichiometric Imbalance\n"]], ["block_7", ["Stoichiometric Imbalance\n67\n"]], ["block_8", ["O\nr\n_=_\n3%\n(2.7.1)\n\u201dB\n"]], ["block_9", ["O\n0\u2014\nMIL...\u00bb =2: .3}s\n(2T)\n"]], ["block_10", ["H\nH\n"]]], "page_83": [["block_0", [{"image_0": "83_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "83_1.png", "coords": [26, 424, 198, 465], "fig_type": "molecule"}]], ["block_2", ["4.\nThe total number of repeat units distributed among these chains is the number of monomer\nmolecules present initially:\n"]], ["block_3", ["As a check that we have done this correctly, note that Equation 2.7.7 reduces to the previously\nestablished Equation 2.4.5 when r 1.\nOne distinction that should be pointed out involves the comparison of Equation 2.2.1 and\nEquation 2.7.7. In the former we considered explicitly the AB monomer, whereas the latter is based\non the polymerization of AA and BB monomers. In both instances Nn is obtained by dividing the\ntotal number of monomer molecules initially present by the total number of chains after the\nreaction has occurred to extent p. Following the same procedure for different reaction mixtures\nresults in a different de\ufb01nition of the repeat unit. In the case of the AB monomer, the repeat unit is\nthe ab entity, which differs from AB by the elimination of the by-product molecule. In the case\nof the AA and BB monomers, the repeat unit in the polymer is the aabb unit, which differs from\nAA + BB by two by\u2014product molecules. Equation 2.2.1 counts the number of ab units in the\npolymer directly. Equation 2.7.7 counts the number of aa + bb units. The number of an + bb units\nis twice the number of aabb units. Rather than attempting to formalize this distinction by\nintroducing more complex notation, we simply point out that application of the formulas of this\nchapter to Specific systems must be accompanied by a re\ufb02ection on the precise meaning of the\ncalculated quantity for the system under consideration.\n"]], ["block_4", [{"image_2": "83_2.png", "coords": [34, 240, 187, 274], "fig_type": "molecule"}]], ["block_5", ["As with other problems with stoichiometry, it is the less-abundant reactant that limits the\nproduct. Accordingly, we define the extent of reaction p to be the fraction of A groups that have\nreacted at any point. Since A and B groups react in a 1:1 proportion, the number of B groups that\nhave reacted when the extent of reaction has reached p equals p123, which in turn equals prv\u2018\ufb01. The\nproduct pr gives the fraction of B groups that have reacted at any point. With these definitions in\nmind, the following relationships are readily obtained:\n"]], ["block_6", ["The number average degree of polymerization is given by dividing the number of repeat units by\nthe number of chains, or\n"]], ["block_7", ["3.\nThe total number of chains is half the number of chain ends:\n"]], ["block_8", ["2.\nThe total number of chain ends is given by the sum of Equation 2.7.2 and Equation 2.7.3:\n"]], ["block_9", ["68\nStep-Growth Polymerization\n"]], ["block_10", ["1.\nThe number of unreacted functional groups after the reaction reaches extent p is\n"]], ["block_11", [{"image_3": "83_3.png", "coords": [45, 291, 181, 326], "fig_type": "molecule"}]], ["block_12", ["1+1/r\n_\nl+r\nNu:\n1+1/rh2p \n1+r-\u20142pr\n(2.7.7)\n"]], ["block_13", ["and\n"]], ["block_14", ["1\n1\n0\nVchains \nE\n1 + ;\n"]], ["block_15", ["I\nl\nl\nl\nI\"repeat units \n'2\u2018\nVOA + 5\nV2, \n'2\u2018 (I + ;) V:\n(2.76)\n"]], ["block_16", ["1 _\nVends : (I p +\npr) VOA\n(2.7.4)\nr\n"]], ["block_17", ["VB (1 mpg (1 \n(2.7.3)\n,.\n"]], ["block_18", ["VA (1 \u2014p)v\u2019\u00b0A\n(2.7.2)\n"]], ["block_19", [{"image_4": "83_4.png", "coords": [54, 362, 225, 394], "fig_type": "molecule"}]], ["block_20", ["\u201c\u2014 2p\nVA\n(2H75)\n"]]], "page_84": [["block_0", [{"image_0": "84_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 2.5\nSome Values of NH Calculated by Equation 2.7.7\nfor Values of r and p Close to Unity\n"]], ["block_2", ["0.95\n13.5\n13.2\n23.3\n39.0\n0.97\n15.5\n22.3\n39.9\n65.7\n0.99\n13.3\n23.7\n66.3\n199\n1.00\n20.0\n33.3\n100\n00\n"]], ["block_3", ["The parameter r\u2019 continues to measure the ratio of the number of A and B groups; the factor 2\nenters since the monofunctional reagent has the same effect on the degree of polymerization as a\ndifunctional molecule with two B groups, hence, is doubly effective compared to the latter. With\nthis modification taken into account, Equation 2.7.7 enables us to evaluate quantitatively the effect\nof stoichiometric imbalance or monofunctional reagents, whether these are intentionally intro-\nduced to regulate Nn or whether they arise from impurities or side reactions.\nThe parameter r varies between 0 and 1; as such it has the same range as p. Although the\nquantitative effect of r and p on Nn is different, the qualitative effect is similar for each: the closer\neach of these fractions is to unity, higher degrees of polymerization are obtained. Table 2.5 shows\nsome values of NH calculated from Equation 2.7.7 for several combinations of (larger values of)\nr and p. Inspection of Table 2.5 reveals the following:\n"]], ["block_4", ["The distinction pointed out in the last paragraph carries over to the evaluation of MD from N\u201c.\nWe assume that the chain of the polymer is great enough to render unnecessary any\ncorrection case, the molecular ofpolymer\nis obtained the of polymerization by multiplying the latter by the molecular weight of\nthe repeat unit. The following examples illustrate the distinction under consideration:\n"]], ["block_5", ["Equation 2.7.7 also applies to the case when some of the excess B groups present are in the form\nof monofunctional reagents. In this latter situation the definition of r is modified somewhat\n(and labeled with a prime) to allow for the fact that some of the B groups are in BB-type monomers\n(unprimed) and some are in monofunctional (primed) molecules:\n"]], ["block_6", ["r\np 0.95\np 0.97\np 0.99\np 1.00\n"]], ["block_7", ["and M0 58. Neglecting end groups, we have Mn 58 ND with Nn given by Equation 2.2.1.\n2.\nPolymerization of AA and BB monomers is illustrated by butane-1,4-diol and adipic acid. The\naabb repeat unit in the polymer has an M0 value of 200. If Equation 2.2.4 is used to evaluate\nNn, it gives the number of aa + bb units; therefore Mn (200 Nn)/2.\n3.\nAn equivalent way of looking at the conclusion of item (2) is to recall that. Equation 2.7.7 gives\nthe (number average) number of monomers of both kinds in the polymer; we should multiply\nthis quantity by the average molecular weight of the two kinds of units in the structure: (88 +\n112)/2 100.\n"]], ["block_8", ["Stoichiometric \n59\n"]], ["block_9", ["1.\nPolymerization of an AB monomer is illustrated by the polyester formed from glycolic acid.\nThe repeat unit in this polymer has the structure\n"]], ["block_10", ["\u2019 \u2014 \u2014\u2014\u2014V3\n2 7 3\n_vg+2v%,\u2019\n(H)\n"]], ["block_11", ["(05.7\n"]]], "page_85": [["block_0", [{"image_0": "85_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4.\nThe monofunctional reagent B\u2019 is the acetic acid in this case and the number of monofunc\u2014\ntional carboxyl groups is 2(0.010) 0.020 125. The number of B groups in BB monomers is\n"]], ["block_2", ["The various expressions we have developed in this section relating p to the size of the polymer\nare all based on N\u201c. Accordingly, we note that the average reactant molecule in this mixture\nhas a molecular weight of 100 as calculated above. Therefore the desired polymer has a value of\nNH 50.\n"]], ["block_3", ["1.\nCalculate the value of p at which the reaction should be stopped to obtain this polymer,\nassuming perfect stoichiometric balance and neglecting end group effects on Mn.\n2.\nAssuming that 0.5 mol% of the diol is lost to polymerization by dehydration to ole\ufb01n, what\nwould be the value of Mn if the reaction were carried out to the same extent as in (1)?\nHow could the loss in (2) be offset so that the desired polymer is still obtained?\n4.\nSuppose the total number of carboxyl groups in the original mixture is 2 mol, of which 1.0% is\npresent as acetic acid to render the resulting polymer inert to subsequent esterification. What\nvalue of p would be required to produce the desired polymer in this case, assuming no other\nstoichiometric imbalance?\n"]], ["block_4", ["1.\nWe use Equation 2.4.5 for the case of equal numbers of A and B groups and find that\np 1\u2014 I/Nn 0.980. Even though Equation 2.4.5 was derived for an AB monomer, it applies\nto this case with the \u201caverage monomer\u201d as the repeat unit.\n2.\nComponent AA is the diol in this case and v3, 0.995 mol; therefore r: 0.995/ 1.00 0.995.\nWe use Equation 2.7.7 and solve for Nn with p 0.980 and r= 0.995:\n"]], ["block_5", ["1.\nFor any value of r, Nn is greater for larger values of p; this conclusion is the same whether the\nprOportions of A and B are balanced or not.\n2.\nThe final 0.05 increase in p has a bigger effect on Nn at r values that are closer to unity than for\nless-balanced mixtures.\n3.\nFor any value of p, Nn is greater for larger values of r; stoichiometric imbalance lowers the\naverage chain-length for the preparation.\n4.\nAn 0.05 increase in r produces a much bigger increase in NH at p 1 than in mixtures that have\nreacted to a lesser extent.\n"]], ["block_6", ["The following example illustrates some of the concepts developed in this section.\n"]], ["block_7", ["and therefore Mn 44.5 x 100 4450 g mol\u201c.\n3.\nThe effect of the lost hydroxyl groups can be offset by carrying out the polymerization to a\nhigher extent of reaction. We use Equation 2.7.7 and solve for p with Nn 50 and r = 0.995:\n"]], ["block_8", ["5-\u201d\n"]], ["block_9", ["It is desired to prepare a polyester with MH 5000 by reacting 1 mol of butane\u20141,4\u2014diol with 1 mol\nof adipic acid.\n"]], ["block_10", ["Solution\n"]], ["block_11", ["70\nStep-Growth Polymerization\n"]], ["block_12", ["Example 2.4\n"]], ["block_13", ["1\n1 + r\n1\n1.995\n=\n1 -\u2014 _\n=\n1 \u2014 \u2014 \u2014 =\n.\n2\np\n(\nN)( 2r )\n(\n50) 1.990\n098 5\n"]], ["block_14", ["Nn\n= 44.5\n"]], ["block_15", ["Nn \nl\u2014r\n(2.7.9)\n"]], ["block_16", ["An interesting special case occurs when p =1; Equation 2.7.7 then becomes\n"]], ["block_17", ["_\n1.995\n\u2014\n1.995 2 (0.995) (0.980)\n"]]], "page_86": [["block_0", [{"image_0": "86_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1.\nIn the simplest case of stoichiometric balance, that is, equal numbers of A and B reactive\ngroups, the number average degree of polymerization N,1 is given by 1/(1\u2014p), where p is the\nextent of reaction, equal to the fraction of A (or B) groups reacted. In general, therefore,\nthe reaction must be driven far toward products (p\u2014il) before appreciable molecular weights\ncan be attained.\n2.\nThe resulting distribution of molecular sizes is called the most probable distribution and the\nassociated polydispersity index approaches 2 as p\u2014il. Two important features of this distri-\nbution are that there are always more i-mers present than (i+l)-mers, for any value of i, but\nthere is an intermediate value of i for which the weight fraction w,- is maximum.\n3.\nIf the reaction is run in the presence of a catalyst (the usual situation), then N,1 should grow\nlinearly in time, whereas for the uncatalyzed case, N,1 will grow with the square root of time.\n4.\nIn reality N will almost always be lower than the theoretical value for a given p, due to a\ncombination of side reactions, including ring formation, contamination by monofunctional\nreagents, and stoichiometric imbalance.\n5.\nThe analysis of these reactions builds on the principle of equal reactivity, the assumption that\nthe reactivity of a given functional group is independent of the molecular weight of the\npolymer to which it is attached. This assumption is quite reliable in most cases of interest.\n"]], ["block_2", ["Remember from Section 2.3 that a progressively longer period of time is required to shift the\nreaction ofand \ntional reactants polymerization times, but are accepted in the\nform of lower \n"]], ["block_3", ["In this chapter we have considered step-growth or condensation polymerization, one of the two\nmain routes to synthetic polymers. Our emphasis has been on the description of the distribution of\npolymer sizes as a function of the extent of reaction and the concentration of reactants, and on the\nassociated kinetics. In addition, we have given an introduction to the two major classes of\ncommercial condensation polymers, polyesters and polyamides, and the different ways they may\nbe produced. The principal results are as follows:\n"]], ["block_4", ["Problems\n"]], ["block_5", ["*GJ. Howard, J. Polym. sci, 37, 310 (1959).\n"]], ["block_6", ["Equation 2.7.7 p usingNn 50 and r\u2019 0.990:\n"]], ["block_7", ["2.8\nChapter Summary\n"]], ["block_8", ["Problems\n71\n"]], ["block_9", ["1.\nHoward describes a model system used to test the molecular weight distribution of a conden-\nsation polymer.)r \u201cThe polymer sample was an acetic acid\u2014stabilized equilibrium nylon-\n6,6. Analysis showed it to have the following end group composition (in equivalents per\n"]], ["block_10", [{"image_1": "86_1.png", "coords": [44, 137, 267, 173], "fig_type": "molecule"}]], ["block_11", ["1\n1 + r\u2019\n1\n1.990\n=\n1\u2014\u2014\n=\n1\u2014\u2014\u2014\u2014 \u2014=0.9849\np\n ( \nNH) ( 2r\u2019 )\n ( \n50) 1.980\n"]], ["block_12", ["1.980: We use Equation 2.7.8 to define r\u2019 for this situation, assuming the number of\nhydroxyl mol:\n"]], ["block_13", ["Ir\n= 0.990\n"]], ["block_14", ["106 g): acetyl=28.9, amine=35.3, and carboxyl=96.5. The number average degree of\n"]], ["block_15", ["_\n2.00\n_\n1.980 + 2 (0.020)\n"]]], "page_87": [["block_0", [{"image_0": "87_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["72\nStep-Growth Polymerization\n"]], ["block_2", ["polymerization is, therefore, 110 and the conversion degree ( extent of reaction) 0.9909.\u201d\nVerify the self-consistency of those numbers.\n2.\nHaward et al. have reported some research in which a copolymer of styrene and hydro-\nxyethylmethacrylate was cross-linked by hexamethylene di-isocyanate.\u2019r Draw the structural\nformula for a portion of this cross-linked polymer and indicate what part of the molecule is\nthe result of a condensation reaction and what part results from addition polymerization.\nThese authors indicate that the cross-linking reaction is carried out in sufficiently dilute\nsolutions of copolymer that the cross\u2014linking is primarily intramolecular rather than intermo-\nlecular. Explain the distinction between these two terms and why concentration affects the\nrelative amounts of each.\n"]], ["block_3", ["4.\nExamination of Figure 2.5 shows that N,/N is greater for i240 at p=0.97 than at either\np=0.95 or p=0.99. This is generally true: various i\u2014mers go through a maximum in\nnumerical abundance as p increases. Show that the extent of reaction at which this maximum\noccurs varies with i as follows: pmax =(i \n1)/z'. For a catalyzed AB reaction, extend this\nexpression to give a function for the time required for an i-mer to reach its maximum\nnumerical abundance. If k0 2.47 X 10\u2018\"4 L mol\u20181 3'1 at 160.5\u00b0C for the polymerization of\n12-hydroxystearic acid,\u00a7 calculate the time at which 15-mers show their maximum abundance\nif the initial concentration of monomer is 3.0 M.\n"]], ["block_4", ["3.\nThe polymerization of B-carboxymethyl caprolactam has been observed to consist of initial\nisomerization via a second-order kinetic process followed by condensation of the isomer to\npolymer:\n"]], ["block_5", ["*RN. Haward, B.M. Parker, and E.F.T. White, Adv. Chem... 91, 49s (1969).\n111x. Reimschuessel, Adv. Chem, 91, 717 (1969).\n9 C.E.H. Bawn and MB. Huglin, Poiymer, 3, 257 (1962).\n"]], ["block_6", [{"image_1": "87_1.png", "coords": [38, 410, 287, 503], "fig_type": "figure"}]], ["block_7", [{"image_2": "87_2.png", "coords": [50, 219, 267, 290], "fig_type": "figure"}]], ["block_8", [{"image_3": "87_3.png", "coords": [51, 224, 118, 292], "fig_type": "molecule"}]], ["block_9", ["The rates of polymerization are thus of \ufb01rst order in VNH2 and in V(c0),o or second order\noverall. Since VNH2= V(C0)20, the rate=kc2, if catalyzed; third order is expected under\nuncatalyzed conditions. The indirect evaluation of c was accomplished by measuring the\namount of monomer reacted, and the average degree of polymerization of the mixture was\ndetermined by viscosity at different times. The following data were obtained at 270\u00b0C; the\nearly part of the experiment gives nonlinear results.1E Graphically test whether these data\nindicate catalyzed or uncatalyzed conditions, and evaluate the rate of constant for polymer-\nization at 270\u00b0C. Propose a name for the polymer.\n"]], ["block_10", ["20\n0.042\n90\n0.015\n30\n0.039\n1 10\n0.013\n40\n0.028\n120\n0.012\n50\n0.024\n150\n0.0096\n60\n0.021\n180\n0.0082\n80\n0.018\n"]], ["block_11", ["NH\nNH2\n0\nO\nO\n.\n.\n.\n.\nIsomenzatlon :\nO\nPolymerizatlonr:\n3\nN\nHO\nO\nO\nO\n"]], ["block_12", ["1' (min)\n6 (Mole fraction)\nt (min)\n6 (Mole fraction)\n"]], ["block_13", [{"image_4": "87_4.png", "coords": [97, 218, 383, 293], "fig_type": "figure"}]], ["block_14", [{"image_5": "87_5.png", "coords": [152, 224, 245, 282], "fig_type": "molecule"}]], ["block_15", [{"image_6": "87_6.png", "coords": [215, 226, 364, 288], "fig_type": "molecule"}]]], "page_88": [["block_0", [{"image_0": "88_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "88_1.png", "coords": [12, 77, 340, 148], "fig_type": "figure"}]], ["block_2", ["TGt). Cooper and A. Katchman, Adv. Chem, 91, 660 (1969).\n1GB. Taylor, J. Am. Chem. 806., 69, 638 (1947).\n"]], ["block_3", [{"image_2": "88_2.png", "coords": [35, 83, 308, 142], "fig_type": "molecule"}]], ["block_4", ["5.\nIn the the polymerization occurs:\n"]], ["block_5", ["Problems\n73\n"]], ["block_6", [{"image_3": "88_3.png", "coords": [36, 278, 338, 392], "fig_type": "figure"}]], ["block_7", [{"image_4": "88_4.png", "coords": [49, 177, 160, 233], "fig_type": "molecule"}]], ["block_8", ["as the x\u2014axis. Evaluate w,- by Equation 2.4.7 for 1'21, 2, 3, and 4 and 0.1 S p S 0.9 in\nincrements of 0.1. Plot these results (w,- on the y-axis) on a separate graph drawn to the same\nscale as the experimental results. Compare your calculated curves with the experimental\ncurves with respect to each of the following points: (1) coordinates used, (2) general shape\nof curves, and (3) labeling of curves.\nThe polymer described in the last problem is commercially called poly(phenylene oxide),\nwhich is not a \u201cproper\u201d name for a molecule with this structure. Propose a more correct name.\nUse the results of the last problem to criticize or defend the following proposition: The\nexperimental data for dimer polymerization can be understood if it is assumed that one\nmolecule of water and one molecule of monomer may split out in the condensation step.\nSteps involving incorporation of the monomer itself (with only water split out) also occur.\nTaylor carefully fractionated a sample of nylon-6,6 and determined the weight fraction of\ndifferent i\u2014mers in the resulting mixtureft The results obtained are given below. Evaluate NW\nfrom these data, then use Equation 2.4.9 to calculate the corresponding value of p. Calculate\nthe theoretical weight fraction of i-mers using this value of p and a suitable array of i values.\nPlot your theoretical curve and the above data points on the same graph. Criticize or defend the\n"]], ["block_9", [",\nWeight percent composition in reaction mixture\nPercent of theoretical\n02 absorbed\nMonomer\nDimer\nTrimer\nTetramer\n"]], ["block_10", ["was used as a starting material. The composition of the mixture was studied as the reaction\nprogressed and the accompanying results were obtained:1\n"]], ["block_11", ["In an investigation to examine the mechanism of this reaction, the dimer (1' 2)\n"]], ["block_12", ["2O\n3\n38 .5\n23\n9\n35\n6\n26\n21\n1 1\n60\nl l\n4\n4\n1\n80\n1\nO\n0\n0\n"]], ["block_13", ["Me\nMe\nHQOQW\n"]], ["block_14", ["n\nOH\n+\u2014g\u2014og\u2014\u2014+H\nOH+nH20\n"]], ["block_15", ["9\n1\n69\n15\n9\n12\n1 .5\n68\n24\n9\n"]], ["block_16", ["Plot a family of curves, each of different i, with composition as the y-axis and 02 absorbed\n"]], ["block_17", ["Me\nMe\n"]], ["block_18", ["Me\nMe\nn\n"]], ["block_19", ["Me\nMe\n"]], ["block_20", [{"image_5": "88_5.png", "coords": [113, 306, 325, 374], "fig_type": "figure"}]]], "page_89": [["block_0", [{"image_0": "89_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["l\u2018C.W. Ayers, J. Appl. Chem, 4, 444 (1954).\ntH.L. Reimschuessel and OJ. Dege, J. Polym. sax. 14-1. 2343 (1971).\n"]], ["block_2", ["74\nStep-Growth Polymerization\n"]], ["block_3", ["10.\nIn the study described in the last problem, caprolactam was polymerized for 24 h at 225\u00b0C in\nsealed tubes containing various amounts of water. Mn and MW were measured for the\n"]], ["block_4", ["On the basis of these observations, criticize or defend the following prOposition: the fact that\nthe separate spots fuse into a single spot of intermediate Rf value proves that block copoly\u2014\nmers form between the two species within the blend upon heating.\n9.\nReimschuessel and Dege polymerized caprolactam in sealed tubes containing about 0.0205\nmol H20 per mole caprolactam.1 In addition, acetic acid (V), sebacic acid (S), hexamethylene\ndiamine (H), and trimesic acid (T) were introduced as additives into separate runs. The\nfollowing table lists (all data per mole caprolactam) the amounts of additive present and the\nanalysis for end groups in various runs. Neglecting end group effects, calculate Mn for each of\nthese polymers from the end group data. Are the trends in molecular weight qualitatively what\nwould be expected in terms of the role of the additive in the reaction mixture? Explain brie\ufb02y.\n"]], ["block_5", ["8.\nPaper chromatograms were developed for 50:50 blends of nylon-6,6 and nylon-6,10 after the\nmixture had been heated to 290\u00b0C for various periods of time.\u2019r The following observations\ndescribe the chromatograms after the indicated times of heating:\n"]], ["block_6", [{"image_1": "89_1.png", "coords": [45, 81, 264, 276], "fig_type": "figure"}]], ["block_7", [{"image_2": "89_2.png", "coords": [53, 510, 323, 595], "fig_type": "figure"}]], ["block_8", ["following proposition: although the fit of the data points is acceptable with this value of p, it\nappears that a slightly smaller value of p would give an even better fit.\n"]], ["block_9", ["12\n6.5\n31 1\n15.2\n35\n19.6\n334\n14.1\n53\n29.4\n357\n13.0\n31\n33.0\n330\n11.5\n104\n35.4\n403\n1 1.0\n127\n36.5\n426\n9.1\n150\n33.0\n449\n7.2\n173\n27.6\n472\n6.5\n196\n25.2\n495\n4.9\n219\n22.9\n513\n4.3\n242\n19.4\n541\n3.9\n265\n13.5\n564\n3.3\n233\n16.3\n"]], ["block_10", ["S\n0.0102\n21.1\n2.3\nH\n0.0102\n1.4\n19.7\nT\n0.0067\n22.0\n2.5\n"]], ["block_11", ["0 h\u2014two spots with Rf values of individual polymers.\n1/4 h\u2014two distinct spots, but closer together than those of 0 h.\n1/2 h\u2014spots are linked together.\n3/4 h\u2014one long, diffuse spot.\n11/2 h\u2014one compact spot, intermediate Rf value.\n"]], ["block_12", ["None\n\u2014\n5.40\n4.99\nV\n0.0205\n19.8\n2.3\n"]], ["block_13", ["Additives\nMoles additive\n\u2014-\u2014COOH (mEq)\n\u2014\u2014NH2 (mEq)\n"]], ["block_14", ["i\nw, x 10\u20144\nr\nw, x 10\u20144\n"]]], "page_90": [["block_0", [{"image_0": "90_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n75\n"]], ["block_2", ["WKV. Korshak and S.V. Vinogradova, Polyester-s, Pergamon, Oxford, 1965.\n"]], ["block_3", ["11.\nAt 270\u00b0C adipic acid decomposes to the extent of 0.31 mol% after 1.5 h.Jr Suppose an initially\nequimolar mixture of adipic acid and diol achieves a value of p 0.990 after 1.5 h, compare\nthe expected and observed values of ND in this experiment. Criticize or defend the following\nproposition: the difference between the observed and expected values would be even greater\nthan calculated above if, instead of the extent of reaction being measured analytically, the\nvalue of p expected (neglecting decomposition) after 1.5 h was calculated by an appropriate\nkinetic equation.\n12.\nShow the reaction sequence and the structure of the resulting polymer from the polyconden-\nsation of these two monomers; note that the reaction (a) has two distinct steps, and that (b) it\nis base-catalyzed.\n"]], ["block_4", ["14.\nFor the most probable distribution, it is clear that there is always more i-mer present than\n(i+1)-mer, at any 0 < p < 1. However, the absolute amount of an i-mer should go through a\nmaximum with time, as the reaction progresses; there is zero to start, but at late enough stages\ni-mer will have mostly reacted to contribute to all the larger species. Use the chain rule and any\n"]], ["block_5", ["13.\nA polyester is prepared under conditions of stoichiometric balance, but no attempt is made to\nremove water. Eventually, the reaction comes to equilibrium with equilibrium constant K. If\n[COOH]0 is the initial concentration of carboxylic acid groups, show that the equilibrium\nwater concentration is\n"]], ["block_6", [{"image_1": "90_1.png", "coords": [37, 382, 317, 449], "fig_type": "figure"}]], ["block_7", [{"image_2": "90_2.png", "coords": [47, 85, 284, 174], "fig_type": "figure"}]], ["block_8", ["Use the molecular weight ratio to calculate the apparent extent of reaction of the caprolactam\nin these systems. Is the variation in p qualitatively consistent with your expectations of the\neffect of increased water content in the system? Plot p versus moisture content and estimate\nby extrapolation the equilibrium moisture content of nylon-6 at 255\u00b0C. Does the apparent\nequilibrium moisture content of this polymer seem consistent with the value given in Section\n2.6 for nylon-6,6 at 290\u00b0C?\n"]], ["block_9", ["resulting mixture by osmometry (see Chapter 7) and light scattering (see Chapter 8),\nrespectively, and the following results were obtained:\n"]], ["block_10", ["Moles H20 (x 103)/mole\nCaprolactam\nMn x 10'3\nMD x 10\"3\n"]], ["block_11", ["49.3\n13.4\n20.0\n34.0\n16.4\n25.6\n25.6\n17.9\n29.8\n20.5\n19.4\n36.6\n"]], ["block_12", ["o\no\nA\n_\n(a)\no\n|\\%\no\n+\nNH2QO\u2014QNH2\n"]], ["block_13", ["[COOH]0\n[H20] KNnavn \u2014 1)\n"]], ["block_14", [{"image_3": "90_3.png", "coords": [58, 384, 152, 445], "fig_type": "molecule"}]], ["block_15", [{"image_4": "90_4.png", "coords": [68, 455, 278, 491], "fig_type": "molecule"}]], ["block_16", [{"image_5": "90_5.png", "coords": [81, 386, 311, 445], "fig_type": "molecule"}]], ["block_17", ["o\no\n"]], ["block_18", [{"image_6": "90_6.png", "coords": [155, 395, 299, 436], "fig_type": "molecule"}]], ["block_19", ["R\n"]]], "page_91": [["block_0", [{"image_0": "91_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["2.\nSolomon, D.H. (Ed), Step Growth Polymerization, Marcel Dekker, New York, 1972.\n"]], ["block_2", ["Allcock, HR. and Lampe, F.W., Contemporary Polymer Chemistry, 2nd ed., Prentice Hall, Englewood Cliffs,\nNJ, 1990.\nOdian, G. Principles of Polymerization, 4th ed., Wiley, New York, 2004.\nRempp, P. and Merrill, E.W., Polymer Synthesis, 2nd ed., Hiithig & Wepf, Basel, 1991.\n"]], ["block_3", ["References\n"]], ["block_4", ["76\nStep-Growth Polymerization\n"]], ["block_5", ["1.\nFlory, P.J., Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\n"]], ["block_6", ["Further Readings\n"]], ["block_7", ["16.\nHydrolysis of an aromatic polyamide with M,l 24,116 gives 39.31% by weight m-amino-\naniline, 59.81% terephthalic acid, and 0.88% benzoic acid. Draw the repeat unit structure of\nthe polymer. Calculate the degree of polymerization and the extent of reaction. Calculate\nwhat the degree of polymerization would have been if the amount of benzoic acid were\ndoubled.\n17.\nCalculate the feed ratio of adipic acid and hexamethylene diamine necessary to achieve a\nmolecular weight of approximately 10,000 at 99.5% conversion. What would the identity of\nthe end groups be in the resulting polymer?\n"]], ["block_8", ["suitable simplifications (k[A]0t >> 1\u2018?) to find the degrees of conversion at which the mole\nfraction and the absolute concentration of i-mer have their maximum in time. Compare this to the\nnumber average degree of polymerization at the same conversion; does the answer make sense?\n15.\nFor the polymerization of succinic acid and 1,4\u2014butanediol under stoichiometric balance in\nxylene:\n"]], ["block_9", ["(a)\nDraw the chemical structures of the reactants, products, and important intermediates for\nboth the strong acid\u2014catalyzed and self\u2014catalyzed case.\n(b)\nGenerate a quantitative plot of NI, versus time for the self\u2014catalyzed case up to 28,000 s,\ngiven k: 6\nX\n10\u201d3 mol\u20182 L2 s\"2 and 3 mol L\u20181 starting concentration of each\nmonomer. How many hours would it take to make a polymer with NI] 300\u2018?\n(c)\nDo the same for the catalyzed case, with k 6 x 10\u20182 mol\"1 L s\u201c1 and the same starting\nconcentration. How many hours would it take to make a polymer with NI, 300\u2018?\n(d)\nQualitatively explain the origin of the different shapes of the curves in the two plots.\n"]]], "page_92": [["block_0", [{"image_0": "92_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In Chapter\n1 we indicated that the category of addition polymers is best characterized by the\nmechanism \nto be chain \nmolecules. thing chain in this discussion. \nchain continues to offer the best description of large polymer molecules. A chain reaction, on the other\nhand, describes a whole of successive events triggered by some initial occurrence. We\nsometimes encounter this description of highway accidents in which one traffic mishap on a fogbound\nhighway results in pileup colliding that can extend for miles. In nuclear reactorsa\ncascade of fission reactions occurs, which is initiated by the capture of the first neutron. In both of\n"]], ["block_2", ["Our primary focus in this section is to point out some of the similarities and differences between\nstep-growth and chain-growth polymerizations. In so doing we shall also have the opportunity to\nindicate some of the different types of chain-growth polymerization systems.\nIn Chapter 2 we saw that step-growth polymerizations occur, one step at a time, through a series of\nrelatively simple organic reactions. By treating the reactivity of the functional groups as independent\n"]], ["block_3", ["automobile collisions that result from the \ufb01rst accident, or the number of fission reactions follow\nfrom the first neutron capture. When we think about the number of monomers that react as a result of a\nsingle initiation step, we are led directly to the degree of polymerization of the resulting molecule. In\nthis way the chain mechanism and the properties of the polymer chains are directly related.\nChain reactions do not go on forever. The fog may clear and the improved visibility ends the suc-\ncession of accidents. Neutron-scavenging control rods may be inserted to shut down a nuclear reactor.\nThe chemical reactions that terminate polymer chain growth are also an important part of the poly-\nmerization mechanism. Killing off the reactive intermediate that keeps the chain going is the essence\nof a termination reaction. Some interesting polymers can be formed when this termination process\nis suppressed; these are called living polymers, and will be discussed extensively in Chapter 4.\nThe kind of reaction that produces a \u201cdead\u201d polymer from a growing chain depends on the\nnature of the reactive intermediate. These intermediates may be free radicals, anions, or cations.\nWe shall devote the rest of this chapter to a discussion of the free-radical mechanism, as it readily\nlends itself to a very general treatment. Furthermore, it is by far the most important chain-growth\nmechanism from a commercial point of view; examples include polyethylene (specifically,\nlow-density polyethylene, LDPE), polystyrene, poly(vinyl chloride), and poly(acry1ates) and\npoly(methacrylates). Anionic polymerization plays a central role in Chapter 4, where we discuss\nthe so-called living polymerizations. In this chapter we deal exclusively with homopolymers. The\nimportant case of copolymers formed by chain-growth mechanisms is taken up in Chapter 4 and\nChapter 5; block copolymers in the former, statistical or random copolymers in the latter.\n"]], ["block_4", ["these examples some initiating event is This is also true in chain-growth polymerization.\nIn the above examples the size of the chain can be measured by considering the number of\n"]], ["block_5", ["3.1\nIntroduction\n"]], ["block_6", ["3.2\nChain-Growth and Step-Growth Polymerizations: Some Comparisons\n"]], ["block_7", ["Chain-Growth Polymerization\n"]], ["block_8", ["3\n"]], ["block_9", ["77\n"]]], "page_93": [["block_0", [{"image_0": "93_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Elsewhere in this chapter we shall see that other reactions\u2014notably, chain transfer and chain\ninhibition\u2014also need to be considered to give a more fully developed picture of chain-growth\npolymerization, but we shall omit these for the time being. Most of this chapter examines the\nkinetics of these three mechanistic steps. We shall describe the rates of the three general kinds of\nreactions by the notation Ri, RP, and Rt for initiation, propagation, and termination, respectively.\nIn the last chapter we presented arguments supporting the idea that reactivity is independent of\nmolecular size. Although the chemical reactions are certainly different between this chapter and\nthe last, we shall also adopt this assumption of equal reactivity for addition polymerization. For\nstep-growth polymerization this assumption simplified the discussion tremendously and at the\nsame time needed careful qualification. We recall that the equal reactivity premise is valid only\nafter an initial size dependence for smaller molecules. The same variability applies to the\npropagation step of addition polymerizations for short-chain oligomers, although things soon\nlevel off and the assumption of equal reactivity holds. We are thus able to treat all propagation\nsteps by the single rate constant kp. Since the total polymer may be the product of hundreds or\neven thousands of such steps, no serious error is made in neglecting the variation that occurs in the\nfirst few steps.\nIn Section 2.3 we rationalized that, say, the first 50% of a step-growth reaction might be different\nfrom the second 50% because the reaction causes dramatic changes in the polarity of the reaction\n"]], ["block_2", ["of the size of the molecule carrying the group, the entire course of the polymerization is described by\nthe conversion of these groups to their condensation products. Two consequences of this are that\nboth high yield and high molecular weight require the reaction to approach completion. In contrast,\nchain-growth polymerization occurs by introducing an active growth center into a reservoir of\nmonomer, followed by the addition of many monomers to that center by a chain-type kinetic\nmechanism. The active center is ultimately killed off by a termination step. The (average) degree\nof polymerization that characterizes the system depends on the frequency of addition steps relative to\ntermination steps. Thus high-molecular\u2014weight polymer can be produced almost immediately. The\nonly thing that is accomplished by allowing the reaction to proceed further is an increased yield of\npolymer. The molecular weight of the product is relatively unaffected. (This simple argument tends\nto break down at high extents of conversion. For this reason we shall focus attention in this chapter on\nlow to moderate conversions to polymer, except where noted.)\nStep-growth polymerizations can be schematically represented by one of the individual reaction\nsteps A + B \u2014> ab, with the realization that the species so connected can be any molecules\ncontaining A and B groups. Chain-growth polymerization, by contrast, requires at least three\ndistinctly different kinds of reactions to describe the mechanism. These three types of reactions\nwill be discussed in the following sections in considerable detail; for now our purpose is just to\nintroduce some vocabulary. The principal steps in the chain-growth mechanism are the following:\n"]], ["block_3", ["78\nChain-Growth Polymerization\n"]], ["block_4", ["2.\nPropagation. The initiator fragment reacts with a monomer M to begin the conversion to\npolymer; the center of activity is retained in the adduct. Monomers continue to add in the same\nway until polymers P,- are formed with the degree of polymerization i:\n"]], ["block_5", ["If i is large enough, the initiator fragment\u2014an endgroup\u2014need not be written explicitly.\n3.\nTermination. By some reaction, generally involving two polymers containing active centers,\nthe growth center is deactivated, resulting in dead polymer:\n"]], ["block_6", ["1.\nInitiation. An active species 1* is formed by the decomposition of an initiator molecule I:\n"]], ["block_7", ["1\u2014) 1*\n(3A)\n"]], ["block_8", ["P?\" + Pj\u201c \u2014> PM (dead polymer)\n(3.C)\n"]], ["block_9", ["1* + M \u2014> IM* J\u2014> IMM* \u2014)\u2014\u00bb\u2014> 19:\"\n(3B)\n"]]], "page_94": [["block_0", [{"image_0": "94_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["mixture. We shall that, under certain circumstances, the rate of addition polymerization\naccelerates as the extent of conversion to polymer increases due to a composition-dependent effect\non termination. the assumption of equal reactivity at all extents of\nreaction, because of the simplification it allows. We will then\nseek to explain the deviations from this ideal or to \ufb01nd experimental conditions\u2014low conversions to\npolymer\u2014under which the assumptions apply. This approach is common in chemistry; for example,\nmost discussions ideal gas law and describe real gases as deviating from the\nideal at high and approaching the ideal as pressure approaches zero.\nIn the last chapter we saw that two reactive groups per molecule are the norm for the formation\nof linear step-growth polymers. A pair of monofunctional reactants might undergo essentially the\nsame reaction, but no polymer is produced because no additional functional groups remain to react.\nOn the other hand, if a molecule contains more than two reactive groups, then branched or cross-\nlinked products can result from step-growth polymerization. By comparison, a wide variety\nof unsaturated monomers undergo chain-growth polymerization. A single kind of monomer\nsuffices\u2014more than one yields a copolymer\u2014and more than one double bond per monomer\nmay result in branching or cross-linking. For example, the 1,2-addition reaction of butadiene\nresults in a chain that has a substituent vinyl group capable of branch formation. Divinyl benzene is\nan example of a bifunctional monomer, which is used as a cross-linking agent in chain-growth\npolymerizations. We shall be primarily concerned with various alkenes or olefins as the monomers\nof interest; however, the carbon\u2014oxygen double bond in aldehydes and ketones can also serve as\nthe unsaturation required for addition polymerization. The polymerization of alkenes yields a\ncarbon atom backbone, whereas the carbonyl group introduces carbon and oxygen atoms into the\nbackbone, thereby illustrating the inadequacy of backbone composition as a basis for distinguish-\ning between addition and condensation polymers.\nIt might be noted that most (butnot all) alkenes are polymerizable by the chain mechanism involving\nfree-radical intermediates, whereas the carbonyl group is generally not polymerized by the free-\nradical mechanism. Carbonyl groups and some carbon\u2014carbon double bonds are polymerized by\nionic mechanisms. Monomers display far more speci\ufb01city where the ionic mechanism is involved\nthan with the free-radical mechanism. For example, acrylamide will polymerize through an anionic\nintermediate but not a cationic one, N\u2014vinyl pyrrolidones by cationic but not anionic intermediates, and\nhalogenated ole\ufb01ns by neither ionic species. In all ofthese cases free-radical polymerization is possible.\nThe initiators used in addition polymerizations are sometimes called \u201ccatalysts,\u201d although\nstrictly speaking this is a misnomer. A true catalyst is recoverable at the end of the reaction,\nchemically unchanged. This is not true of the initiator molecules in most addition polymerizations.\nMonomer and polymer are the initial and final states of the polymerization process, and these\ngovern the thermodynamics of the reaction; the nature and concentration of the intermediates in\nthe process, on the other hand, determine the rate. This makes initiator and catalyst synonyms\nfor the same material. The former term stresses the effect of the reagent on the intermediate, and\nthe latter its effect on the rate. The term catalyst is particularly common in the language of ionic\npolymerizations, but this terminology should not obscure the importance of the initiation step in\nthe overall polymerization mechanism.\nIn the next three sections (Section 3.3 through Section 3.5) we consider initiation, termination,\nand propagation steps in the free-radical mechanism for addition polymerization. As noted above\ntwo additional steps, inhibition and chain transfer, are being ignored at this point. We shall take up\nthese latter topics in Section 3.8.\n"]], ["block_2", ["In this section we shall discuss the initiation step of free-radical polymerization. This discussion is\ncentered around initiators and their decomposition behavior. The first requirement for an initiator\nis that it be a source of free radicals. In addition, the radicals must be produced at an acceptable rate\n"]], ["block_3", ["Initiation\n79\n"]], ["block_4", ["3.3\nInitiation\n"]]], "page_95": [["block_0", [{"image_0": "95_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 3.1\nExamples of Free\u2014Radical Initiation Reactions\n"]], ["block_2", ["1.\nOrganic peroxides or hydroperoxides\n2.\nA20 compounds\n"]], ["block_3", ["Some of the most widely used initiator systems are listed below, and Table 3.1 illustrates their\nbehavior by typical reactions:\n"]], ["block_4", ["3.3.1\nInitiation Reactions\n"]], ["block_5", ["at convenient temperatures; have the required solubility behavior; transfer their activity to mono-\nmers efficiently; be amenable to analysis, preparation, purification, and so on.\n"]], ["block_6", ["80\nChain-Growth Polymerization\n"]], ["block_7", ["2. A20 compounds\n"]], ["block_8", ["3. Redox systems\n"]], ["block_9", ["4. Electromagnetic radiation\n"]], ["block_10", ["1. Organic peroxides or hydroperoxides\n"]], ["block_11", [{"image_1": "95_1.png", "coords": [42, 74, 393, 212], "fig_type": "figure"}]], ["block_12", [{"image_2": "95_2.png", "coords": [51, 93, 307, 168], "fig_type": "figure"}]], ["block_13", [{"image_3": "95_3.png", "coords": [67, 95, 281, 137], "fig_type": "molecule"}]], ["block_14", ["O/EO\u2019O\n(DE/O\n,\n2 (Zia\nBenzoyl peroxide\none?)\ncri-\n"]], ["block_15", [{"image_4": "95_4.png", "coords": [79, 241, 356, 282], "fig_type": "molecule"}]], ["block_16", [{"image_5": "95_5.png", "coords": [79, 317, 312, 378], "fig_type": "figure"}]], ["block_17", [{"image_6": "95_6.png", "coords": [79, 488, 317, 542], "fig_type": "molecule"}]], ["block_18", [{"image_7": "95_7.png", "coords": [80, 375, 314, 563], "fig_type": "figure"}]], ["block_19", [{"image_8": "95_8.png", "coords": [81, 230, 360, 292], "fig_type": "figure"}]], ["block_20", [".00\n"]], ["block_21", [{"image_9": "95_9.png", "coords": [94, 321, 321, 377], "fig_type": "molecule"}]], ["block_22", ["Me\nMe\nMe\nNC>CN \nCqMe\n"]], ["block_13", ["Me \\\nMe\n"]], ["block_14", [{"image_4": "96_4.png", "coords": [121, 586, 196, 613], "fig_type": "molecule"}]], ["block_15", [{"image_5": "96_5.png", "coords": [123, 573, 201, 625], "fig_type": "molecule"}]], ["block_16", ["Mekc\u201d\nN $111:\n"]], ["block_17", [{"image_6": "96_6.png", "coords": [132, 517, 199, 571], "fig_type": "molecule"}]], ["block_18", ["Me\n"]], ["block_19", ["Me;\nCN\n"]], ["block_20", ["\u2019l'.\n"]], ["block_21", ["CN\n"]], ["block_22", ["Me\n"]]], "page_97": [["block_0", [{"image_0": "97_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "97_1.png", "coords": [27, 520, 300, 589], "fig_type": "figure"}]], ["block_2", ["We recall some of the ideas of kinetics from the summary given in Section 2.2 and recognize that\nthe rates of initiator decomposition can be developed in terms of the reactions listed in Table 3.1.\n"]], ["block_3", ["The diphenylpicrylhydrazyl radical itself is readily followed spectrophotometrically, as it loses an\nintense purple color on reacting. Unfortunately, this reaction is not always quantitative.\n"]], ["block_4", ["The initiator efficiency is not an exclusive property of the initiator, but depends on the conditions of\nthe polymerization experiment, including the solvent. In many experimental situations,f lies in the\nrange of 0.3\u20140.8. The efficiency should be regarded as an empirical parameter whose value is\ndetermined experimentally. Several methods are used for the evaluation of initiator ef\ufb01ciency, the\nbest being the direct analysis for initiator fragments as endgroups compared to the amount of initiator\nconsumed, with proper allowances for stoichiometry. As an endgroup method, this procedure is\ndif\ufb01cult in addition polymers, where molecular weights are higher than in condensation polymers.\nResearch with isotopically labeled initiators is particularly useful in this application. Since the quantity\nis so dependent on the conditions of the experiment, it should be monitored for each system studied.\nScavengers such as diphenylpicrylhydrazyl radicals [II] react with other radicals and thus\nprovide an indirect method for analysis of the number of free radicals in a system:\n"]], ["block_5", ["While this reaction with solvent continues to provide free radicals, these may be less reactive\nspecies than the original initiator fragments. We shall have more to say about the transfer of free\u2014\nradical functionality to solvent in Section 3.8.\nThe significant thing about these, and numerous other side reactions that could be described, is\nthe fact that they lower the ef\ufb01ciency of the initiator in promoting polymerization. To quantify this\nconcept we de\ufb01ne the initiator e\ufb01i\u2018ciencyf to be the following fraction:\n"]], ["block_6", ["In both of these examples, initiator is consumed, but no polymerization is started.\nOnce the radicals diffuse out of the solvent cage, reaction with monomer is the most\nprobable reaction in bulk polymerizations, since monomers are the species most likely to be encoun-\ntered. Reaction with polymer radicals or initiator molecules cannot be ruled out, but these are less\nimportant because of the lower concentration of the latter species. In the presence of solvent, reactions\nbetween the initiator radical and the solvent may effectively compete with polymer initiation. This\ndepends very much on the speci\ufb01c chemicals involved. For example, carbon tetrachloride is quite\nreactive toward radicals because of the resonance stabilization of the solvent radical produced:\n"]], ["block_7", ["82\nChain-Growth Polymerization\n"]], ["block_8", ["3.3.3\nKinetics of Initiation\n"]], ["block_9", [{"image_2": "97_2.png", "coords": [37, 519, 258, 587], "fig_type": "molecule"}]], ["block_10", [{"image_3": "97_3.png", "coords": [41, 226, 318, 267], "fig_type": "molecule"}]], ["block_11", [{"image_4": "97_4.png", "coords": [44, 51, 229, 114], "fig_type": "molecule"}]], ["block_12", ["0\nJL\n,Me\n+\n002\no\n/\nMe\n0\n/IL\n(3B)\n2\nMe\nO \\\nMe\u2014Me +\n2(302\n"]], ["block_13", ["(3|\n(3|\n(3|\n(3|\n(3|\n(3|\n(3|\n(3|\u2018\n"]], ["block_14", ["_ \nRadicals formed by initiator\n(33.1)\n"]], ["block_15", ["(3|\n(3|\n(3|\n(3|\n"]], ["block_16", ["N\u2014N\nN02\n4-\nR\n0\nadduct\n(3 .F)\n"]], ["block_17", [{"image_5": "97_5.png", "coords": [83, 54, 227, 92], "fig_type": "molecule"}]], ["block_18", [{"image_6": "97_6.png", "coords": [103, 62, 221, 83], "fig_type": "molecule"}]], ["block_19", [{"image_7": "97_7.png", "coords": [151, 227, 296, 263], "fig_type": "molecule"}]], ["block_20", ["[H]\n"]], ["block_21", ["[1]\n"]]], "page_98": [["block_0", [{"image_0": "98_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where [abs is the intensity of the light absorbed and the constant (15\u2019 is called the quantum yield. The\nfactor 2 is again included for reasons of stoichiometry.\nSince (1/2) d[ I\u00b0]/dt\u2014d[I]/dt in the case of the azo initiators, Equation 3.3.2 can also be\nwritten as \u2014d[I]/dt= kd[l] or, by integration, 1n([I]/[I]0) \u2014kdt, where [Hg is the initiator concen\u2014\ntration at t 0. Figure 3.1 shows a test of this relationship for AIBN in xylene at 77\u00b0C. Except for a\nshort induction period, the data points fall on a straight line. The evaluation of kd from these data is\npresented in the following example.\n"]], ["block_2", ["The decomposition of AIBN in xylene at 77\u00b0C was studied by measuring the volume of N2 evolved\n"]], ["block_3", ["as a function of time. The volumes obtained at time t and r: 00 are V, and V00, respectively. Show\nthat the manner of plotting used in Figure 3.1 is consistent with the integrated \ufb01rst\u2014order rate law\nand evaluate kd.\n"]], ["block_4", ["Figure 3.1\nVolume of nitrogen evolved from the decomposition of AIBN at 77\u00b0C plotted according to the\nfirst\u2014order rate law as discussed in Example 3.1. (Reprinted from Amett, L.M., J. Am. Chem. Soc, 74, 2027,\n1952. With permission.)\n"]], ["block_5", ["2.\nFor redox systems\n"]], ["block_6", ["Using monitorthe rates, we write the\nfollowing:\n"]], ["block_7", ["3.\nFor photochemical initiation\n"]], ["block_8", ["Initiation\n33\n"]], ["block_9", ["Example 3.1\n"]], ["block_10", ["1.\nFor peroxides and compounds\n"]], ["block_11", ["A\nS\" >3\nJ-\n-\u20140.4\n\u2014\nv\nD)\n.9\n_08\n_.\n"]], ["block_12", ["9\u2014371q\u2019 abs\n(3.3.4)\n"]], ["block_13", ["d[I-] _\n\u20195\" \n(3.3.2)\n"]], ["block_14", ["d\u2014E? k[0x][Red]\n(3.3.3)\n"]], ["block_15", ["where kd is the rate constant for the homolytic decomposition of the initiator and [I] is the\nconcentration of the initiator. The factor 2 appears because of the stoichiometry in these\nparticular reactions.\n"]], ["block_16", ["where the bracketed terms describe the concentrations of oxidizing and reducing agents and k\nis the rate constant for the particular reactants.\n"]], ["block_17", ["0.0\n\u2014\n"]], ["block_18", [{"image_1": "98_1.png", "coords": [62, 465, 220, 620], "fig_type": "figure"}]], ["block_19", ["J_\nl\n_l\nl\n0\n80\n160\n240\n320\nTime (min)\n"]]], "page_99": [["block_0", [{"image_0": "99_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where we have combined the factors off and qfr\u2019 into a composite quantum yield qb, since both of\nthe separate factors are measures of efficiency.\nAny one of these expressions gives the rate of initiation R, for the particular catalytic system\nemployed. We shall focus attention on the homolytic decomposition of a single initiator as the\nmode of initiation throughout most of this chapter, since this reaction typifies the most widely used\nfree-radical initiators. Appropriate expressions for initiation that follow Equation 3.3.6 are readily\nderived.\n"]], ["block_2", ["An important application of photochemical initiation is in the determination of the rate constants\nthat appear in the overall analysis of the chain-growth mechanism. Although we outline this\n"]], ["block_3", ["1.\nV0 0 at t: 0, when no decomposition has occurred.\n2.\nV00 at t: 00, when complete decomposition has occurred.\n3.\nV, at time t, when some fraction of initiator has decomposed.\n"]], ["block_4", ["Next we assume that only a fractionf of these initiator fragments actually reacts with monomer\nto transfer the radical functionality to monomer:\n"]], ["block_5", ["As indicated in the last section, we regard the reactivity of the species IP; to be independent of\nthe value of 1'. Accordingly, all subsequent additions to IM3 in Reaction (3.G) are pr0pagation steps\nand Reaction (3.G) represents the initiation of polymerization. Although it is premature at this\npoint, we disregard endgroups and represent the polymeric radicals of whatever size by the symbol\n"]], ["block_6", ["84\nChain-Growth Polymerization\n"]], ["block_7", ["The ratio [I]/[I]0 gives the fraction of initiator remaining at time t. The volume of N2 evolved is:\n"]], ["block_8", ["3.\nBy photochemical initiation\n"]], ["block_9", ["3.3.4\nPhotochemical Initiation\n"]], ["block_10", ["2.\nBy redox systems\n"]], ["block_11", ["The fraction decomposed\nat t is given by (V,- V0)/(VOO\u2014 V0)\nand the fraction remaining\nat t is l\n\u2014 (V,\u2014V0)/(VOO\u2014V0)= (VOO\u2014V,)/(VOO\u2014-VO). Since VO=0, this becomes (VOO\u2014V,)/VOO or\n[I]/[I]O l V,/VOO. Therefore a plot of ln(l \u2014-V,/VOO) versus t is predicted to be linear with slope\n\u2014kd. (If logarithms to base 10 were used, the slope would equal -\u2014kd/2.303.)\nFrom Figure 3.1,\n"]], ["block_12", ["Po. Accordingly, we write the following for the initiation of polymer radicals:\n"]], ["block_13", ["Solution\n"]], ["block_14", ["1.\nBy peroxide and azo compounds\n"]], ["block_15", ["kd = 5.8 x 10\u20143 min\u20141\n"]], ["block_16", ["I. + M L M\n(3.G)\n"]], ["block_17", ["\u2014\u20140.4 (\u2014-0.8) _\n\u2018kd\n81\n=\u2014 _ \u20142.5\n10\u20183\n' \u20181 =\n0136\n160 320\nX\nmm\n2.303\n"]], ["block_18", ["@ 2f\u2018ie\u2019rlabs 2(iif\u2019labs\n(33\u00b0?)\ndt\n"]], ["block_19", ["5:? \ufb02c[0x][Red]\n(3.3.6)\n"]], ["block_20", ["d[P\u00b0]\nT\n= kd\ufb02]\n(3.3.5)\n"]]], "page_100": [["block_0", [{"image_0": "100_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["method in Section develop Equation 3.3.7 somewhat It\nis not feasible \npertinent desires \nanalytical or physical \n"]], ["block_2", ["2.\nAbsorbance A as measured by spectrophotometers is defined as\n"]], ["block_3", ["Initiation\n85\n"]], ["block_4", ["3.\nSince lab, equals the difference 10 It\n"]], ["block_5", ["4.\nSubstituting this result into Equation 3.3.7 gives\n"]], ["block_6", ["Note that although Equation 3.3.5 and Equation 3.3.12 are both first-order rate laws, the physical\nsigni\ufb01cance of the proportionality factors is quite different in the two cases. The rate constants\nshown in Equation 3.3.5 and Equation 3.3.6 show a temperature dependence described by the\nArrhenius equation:\n"]], ["block_7", ["where 13* is the activation energy, which is interpreted as the height of the energy barrier to a\nreaction, and A is the prefactor. Activation energies are evaluated from experiments in which rate\nconstants are measured at different temperatures. Taking logarithms on both sides of Equation\n3.3.13 gives In k In A \u2014E*/RT. Therefore 13* is obtained from the slope of a plot of In k against\n1/T. As usual, T is in kelvin and R and E* are in (the same) energy units.\nSince 8* is positive according to this picture, the form of the Arrhenius equation assures that\nk gets larger as T increases. This means that a larger proportion of molecules have sufficient energy\nto surmount the energy barrier at higher temperatures. This assumes, of course, that thermal\nenergy is the source of 13*, something that is not the case in photoinitiated reactions. The effective\n\ufb01rst-order rate constants k and Io8b\u2014fOI' thermal initiation and photoinitiation, respectively\u2014do\n"]], ["block_8", ["where [c] is the concentration of monomer or initiator for the two reactions shown in Table 3.1.\n"]], ["block_9", ["3.3.5\nTemperature Dependence of Initiation Rates\n"]], ["block_10", ["1.\nIntensity of light transmitted (subscript t) through a sample It depends on the intensity of the\nincident (subscript 0) light 10, the thickness of the sample b, and the concentration [c] of\nthe absorbing species\n"]], ["block_11", ["k Ae_E*/RT\n(3.3. 13)\n"]], ["block_12", ["1, = 106\u2014616\u201c)\n(3.3.8)\n"]], ["block_13", ["where the proportionality constants is called the absorption coefficient (or molar absorptivity\nif [c] is in moles/liter) and is a property of the absorber. The reader may recognize this\nequation as a form of the famous Beer\u2019s law.\n"]], ["block_14", ["10\nA loglo (X)\n(3.3.9)\n"]], ["block_15", ["labs Io(8[clb)\n(3.3.11)\n"]], ["block_16", ["% 2\u00a2106[c]b\n(3.3.12)\n"]], ["block_17", ["labs M1 e\u2018elcl\u201d)\n(3.3.10)\n"]], ["block_18", ["The variation in absorbance with wavelength re\ufb02ects the wavelength dependence of .9.\n"]], ["block_19", ["If the exponent in Equation 3.3.10 is small\u2014which in practice means dilute solutions, since\nmost absorption experiments are done where .9 is large\u2014then the exponential can be expanded\n(see Appendix), e)r g 1 +x + o, with only the leading terms retained to give\n"]]], "page_101": [["block_0", [{"image_0": "101_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "101_1.png", "coords": [30, 583, 315, 628], "fig_type": "molecule"}]], ["block_2", ["This mode of termination produces a negligible effect on the molecular weight of the reacting\nspecies, but it does produce a terminal unsaturation in one of the dead polymer molecules.\nEach polymer molecule contains one initiator fragment when termination occurs by disproportionation.\n"]], ["block_3", ["The formation of initiator radicals is not the only process that determines the concentration of free\nradicals in a polymerization system. Polymer propagation itself does not change the radical\nconcentration; it merely converts one radical to another. Termination steps also occur, however,\nand these remove radicals from the system. We shall discuss combination and disproportionation\nreactions as the two principal modes of termination.\n"]], ["block_4", ["The degrees of polymerization i andj in the two combining radicals can have any values, and the\nmolecular weight of the product molecule will be considerably higher on the average than the\nradicals so terminated. The polymeric product molecule contains two initiator fragments per\nmolecule by this mode of termination. Note also that for a vinyl monomer, such as styrene or\nmethyl methacrylate, the combination reaction produces a single head-to-head linkage, with the\nside groups attached to adjacent backbone carbons instead of every other carbon.\nTermination by disproportionation comes about when an atom, usually hydrogen, is transferred\nfrom one polymer radical to another:\n"]], ["block_5", ["Table 3.2\nRate Constants (at the Indicated Temperature) and Activation Energies for Some Initiator\nDecomposition Reactions\n"]], ["block_6", ["Source: Data from Masson, J.C. in Polymer Handbook, 3rd ed., Brandrup, J. and Immergut, E.H. (Eds), Wiley, New York,\n1989.\n"]], ["block_7", ["Termination by combination results in the simultaneous destruction of two radicals by direct\ncoupling:\n"]], ["block_8", ["not show the same temperature dependence. The former follows the Arrhenius equation, whereas\nthe latter cluster of terms in Equation 3.3.12 is essentially independent of T.\nThe activation energies for the decomposition (subscript (1) reaction of several different\ninitiators in various solvents are shown in Table 3.2. Also listed are values of kd for these systems\nat the temperature shown. The Arrhenius equation can be used in the form In (kd,1/kd,2)= \u2014(E*/R)\n(l/Tl HR) to evaluate kd values for these systems at temperatures different from those given in\nTable 3.2.\n"]], ["block_9", ["3.4.1\nCombination and Disproportionation\n"]], ["block_10", ["2,2\u2019-Azobisisobutyronitrile\nBenzene\n70\n3.17 x 10\u20145\n123.4\nccn\n40\n2.15 x 10-7\n128.4\nToluene\n100\n1.60 x 10-3\n121.3\nt-Buty] peroxide\nBenzene\n100\n8.8 x 10\u20147\n146.9\nBenzoyl peroxide\nBenzene\n70\n1.48 x 10\u20145\n123.8\nCumene\n60\n1.45 x 10'6\n120.5\nt\u2014Butyl hydroperoxide\nBenzene\n169\n2.0 x 10\u20145\n170.7\n"]], ["block_11", ["Initiator\nSolvent\nr (\u201d(3)\nk, (3\")\n53 (kJ moi\u201c)\n"]], ["block_12", ["86\nChain-Growth Polymerization\n"]], ["block_13", ["3.4\nTermination\n"]], ["block_14", [{"image_2": "101_2.png", "coords": [37, 74, 438, 203], "fig_type": "figure"}]], ["block_15", [{"image_3": "101_3.png", "coords": [43, 589, 211, 625], "fig_type": "molecule"}]], ["block_16", ["Pf. + 'Pj --> Pg+j\n(3.H)\n"]], ["block_17", ["X\nX\nX\nX\nPom\n+\nrpm Pi\u20141/\\/\n+\n \\fpj \ufb01 \n(3.1)\nH\nH\n"]], ["block_18", [{"image_4": "101_4.png", "coords": [140, 588, 302, 619], "fig_type": "molecule"}]]], "page_102": [["block_0", [{"image_0": "102_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Since the disproportionation reaction requires bond breaking, which is not required for combin-\nation, Ei'fd is expected to be greater than E\ufb01fc. This causes the exponential to be large at low\ntemperatures, making combination the preferred mode of termination under these circumstances.\nNote that at higher temperatures this bias in favor of one mode of termination over another\ndecreases as the difference in activation energies becomes smaller relative to the thermal energy\nRT. Experimental results on modes of termination show that this qualitative argument must be\napplied cautiously. The actual determination of the partitioning between the two modes of\ntermination is best accomplished by analysis of endgroups, using the difference in endgroup\ndistribution noted above.\nTable 3.3 lists the activation energies for termination (these are overall values, not identified as\nto mode) of several different radicals. The rate constants for termination at 60\u00b0C are also given. We\nshall see in Section 3.6 how these constants are determined.\n"]], ["block_2", ["Combination and disproportionation are competitive processes and do not occur to the same\nextent for all polymers, although in general combination is more prevalent. For poly(methyl\nmethacrylate), both reactions contribute to termination, with disproportionation favored. Both\nrate constants for termination individually follow the Arrhenius equation, so the relative amounts\nof termination by the two modes are given by\n"]], ["block_3", ["where the subscript c specifically indicates termination by combination.\n4.\nFor disproportionation,\n"]], ["block_4", ["Termination\n87\n"]], ["block_5", ["3.\nFor combination,\n"]], ["block_6", ["termination \nfollowing:\n"]], ["block_7", ["where RI and kt are the rate and rate constant for termination (subscript t) and the factor 2\nenters (by convention) because two radicals are lost for each termination step.\n2.\nThe polymer radical concentration in Equation 3.4.1 represents the total concentration of all\nsuch species, regardless of their degree of polymerization; that is,\n"]], ["block_8", ["where the subscript d specifically indicates termination by disproportionation.\n5.\nIn the event that the two modes of termination are not distinguished, Equation 3.4.1 represents\nthe sum of Equation 3.4.3 and Equation 3.4.4, or\n"]], ["block_9", ["1.\nFor general termination,\n"]], ["block_10", ["Termination by combination\n_ _ _ \n\u2014(E\u00a7':c\u2014 Eifd)\n3 4 6\nTermination by disproportionation \nkm\n\u2014\nAt,de*5?,d/RT\n\u2014\nAnd\nCXP(\nRT\n)\n(\n.\n\u00b0 )\n"]], ["block_11", ["Kinetic analysis of the two modes of termination is quite straightforward, since each mode of\n"]], ["block_12", ["d P-\nR. \u2014Q 24,413.12\n(3.4.3)\ndz\n"]], ["block_13", ["d P0\nRt \u2014\u00a5 2kt,d[P~]2\n(3.4.4)\ndz\n"]], ["block_14", ["Rt \u201c(E5 2kt[P']2\n(3.4.1)\ndr\n"]], ["block_15", ["kt kt,c + kt,d\n(3.4.5)\n"]], ["block_16", ["[P] Z [Pr]\n(3.4.2)\n"]], ["block_17", [{"image_1": "102_1.png", "coords": [63, 198, 115, 232], "fig_type": "molecule"}]], ["block_18", ["all i\n"]]], "page_103": [["block_0", [{"image_0": "103_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 3.3\nRate Constants at 60\u00b0C and Activation Energies for Some\nTermination Reactions\n"]], ["block_2", ["Source: Data from Korus, R. and O\u2019Driscoll, K.F. in Polymer Handbook, 3rd ed.,\nBrandrup, J. and Irnmergut, E.H. (Eds), Wiley, New York, 1989.\n"]], ["block_3", ["88\nChain-Growth Polymerization\n"]], ["block_4", ["Acrylonitrile\n1.5\n78.2\nMethyl acrylate\n22.2\n0.95\nMethyl methacrylate\n11.9\n2.55\nStyrene\n8.0\n6.0\nVinyl acetate\n21.9\n2.9\n2-Vinyl pyridine\n21.0\n3.3\n"]], ["block_5", ["Monomer\nE; (kJ 1110]\u201c)\nk, x 10\u201c7 (L mor\u2018 s\")\n"]], ["block_6", ["The assumption that k values are constant over the entire duration of the reaction breaks down\nfor termination reactions in bulk polymerizations. Here, as in Section 2.2, we can consider the\ntermination process\u2014\u2014\u2014whether by combination or disproportionation\u2014to depend on the rates at\nwhich polymer molecules can diffuse into (characterized by ki) or out of (characterized by k0) the\nsame solvent cage and the rate at which chemical reaction between them (characterized by k,)\noccurs in that cage. In Chapter 2, we saw that two limiting cases of Equation 2.2.8 could be readily\nidentified:\n"]], ["block_7", ["4.\nThis situation is expected to apply to radical termination, especially by combination, because\nof the high reactivity of the trapped radicals. Only one constant appears that depends on the\ndiffusion of the polymer radicals, so it cannot cancel out and may contribute to a dependence\nof kt on the extent of reaction or the degree of polymerization.\n"]], ["block_8", ["Figure 3.2 shows how the percent conversion of methyl methacrylate to polymer varies with\ntime. These experiments were carried out in benzene at 50\u00b0C. The different curves correspond to\ndifferent concentrations of monomer. Up to about 40% monomer, the conversion varies smoothly\nwith time, gradually slowing down at higher conversions owing to the depletion of monomer. At\nhigh concentrations, however, the polymerization starts to show an acceleration between 20% and\n40% conversion. This behavior, known as the Trommsdorjf e\ufb01\u2018ect [2], is attributed to a decrease in\nthe rate of termination with increasing conversion. This, in turn, is due to the increase in viscosity\nthat has an adverse effect on kt through Equation 3.4.8. Considerations of this sort are important in\nbulk polymerizations where high conversion is the objective, but this complication is something\nwe will avoid. Hence we shall be mainly concerned with solution polymerization and/or low\ndegrees of conversion where Itt may be justifiably treated as a true constant. We shall see in Section\n3.8 that the introduction of solvent is accompanied by some complications of its own, but we shall\nignore this for now.\n"]], ["block_9", ["3.4.2\nEffect of Termination on Conversion to Polymer\n"]], ["block_10", ["2.\nThis situation seems highly probable for step-growth polymerization because of the high\nactivation energy of many condensation reactions. The constants for the diffusion-dependent\nsteps, which might be functions of molecular size or the extent of the reaction, cancel out.\n3.\nRate of reaction > rate of diffusion (Equation 2.2.10):\n"]], ["block_11", ["1.\nRate of diffusion > rate of reaction (Equation 2.2.9):\n"]], ["block_12", [{"image_1": "103_1.png", "coords": [37, 79, 317, 179], "fig_type": "figure"}]], ["block_13", ["kt \u2014kr\n(3.4.7)\n"]], ["block_14", ["lct ki\n(3.4.8)\n"]]], "page_104": [["block_0", [{"image_0": "104_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "104_1.png", "coords": [28, 17, 286, 194], "fig_type": "figure"}]], ["block_2", ["Termination\n89\n"]], ["block_3", ["benzene \npermission.)\n"]], ["block_4", ["Polymer propagation \ntwo opposing processes, willeventually reach point of balance. This\ncondition is called the and characterized by constant total concentration of free\nradicals. Under s) the net rate of initiation must equal the net\nrate of termination. Using Equation 3.3.5 for the rate of initiation (i.e., two radicals per initiator\nmolecule) and Equation 3.4.1 for termination, we write\n"]], ["block_5", ["Calculate [Po]S and the time required for the free-radical concentration to reach 99% of this value\nusing the following as typical values for constants and concentrations: kd =1.0 X 10\"4 s\", kt:\n3 x 107 L mol\u2014l s\u201c,f: 1/2, and [I]O 10\u20183 mol L\u201c. Comment on the assumption [I] [I]0 that\nwas made in deriving this nonstationary-state equation.\n"]], ["block_6", ["3.4.3\nStationary-State Radical Concentration\n"]], ["block_7", ["01'\n"]], ["block_8", ["This important equation shows that the stationary-state free-radical concentration increases with\n[I]\u201d2 and varies directly with kit/2 and inversely with ktm. The concentration of free radicals\ndetermines the rate at which polymer forms and the eventual molecular weight of the polymer,\nsince each radical is a growth site. We shall examine these aspects of Equation 3.4.10 in the next\nsection. We conclude this section with a numerical example illustrating the stationary-state radical\nconcentration for a typical system.\n"]], ["block_9", ["For an initiator concentration that is constant at [I]0, the nonstationary-state radical concentration\nvaries with time according to the following expression:\n"]], ["block_10", ["Figure 3.2\nAcceleration of the polymerization rate for methyl methacrylate at the concentrations shown in\n"]], ["block_11", ["Example 3.2\n"]], ["block_12", [".5\nE\n10%\ne 60\ng\nr\nC\n8 40 \n"]], ["block_13", ["a?\nr\n20 \n"]], ["block_14", ["2mm 2mm:\n(3.4.9)\n"]], ["block_15", ["\ufb02cd \n[13.15: (7;)\n[111/2\n(3.4.10)\n"]], ["block_16", [{"image_2": "104_2.png", "coords": [47, 395, 140, 441], "fig_type": "molecule"}]], ["block_17", ["[12.] = eXP[(4fkdkt[I]0)l/2 t] 1\n[P].\nexp [(4\ufb02cdk,[l]0)'/2 t] + 1\n"]], ["block_18", ["30 \n100%\n80%\n60%\n40%\n"]], ["block_19", ["l\nI\nI\nI\nI\nI\nI\nI\nI\n_L\nI\nI\nI\nI\n_I_\nI\nI\n0\n500\n1000\n1500\nTime (min)\n"]], ["block_20", ["L\n"]]], "page_105": [["block_0", [{"image_0": "105_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This low concentration is typical of free-radical polymerizations. Next we inquire how long it will\ntake the free-radical concentration to reach 0.99 [P\u00b0]s, or 4.04 x\n10\u20148 mol L\u20181 in this case. Let\na =(4fkdkt[l]0)1/2 and rearrange the expression given to solve for t when [P\u00b0]/[P\u00b0]s 0.99: 0.99\n(ear + 1) e\u2018\u201d\u2014 1, or 1 + 0.99 e\u2018\u201d(1 0.99). Therefore the product at = ln(1.99/0.01) In 199 5.29,\nand a [4(1/2)(1.0 x 10\u20144)(3 x 10500\u2014511\u201d: 2.45 3*. Hence t= 529/245 2.16 s.\nThis short period is also typical of the time required to reach the stationary state. The\nassumption that [I] [He may be assessed by examining the integrated form of Equation 3.3.2\nfor this system and calculating the ratio [I]/[I]0 after 2.16 s:\n"]], ["block_2", ["The propagation of polymer chains is easy to consider under stationary-state conditions. As the\npreceding example illustrates, the stationary state is reached very rapidly, so we lose only a brief\nperiod at the start of the reaction by restricting ourselves to the stationary state. Of course, the\nstationary-state approximation breaks down at the end of the reaction also, when the radical\nconcentration drops toward zero. We shall restrict our attention to relatively low conversion to\npolymer, however, to avoid the complications of the Trommsdorff effect. Therefore deviations\nfrom the stationary state at long times need not concern us.\nIt is worth taking a moment to examine the propagation step more explicitly in terms of the\nreaction mechanism itself. As an example, consider the case of styrene as a representative vinyl\nmonomer. The polystyryl radical is stabilized on the terminal-substituted carbon by resonance\ndelocalization:\nsea~a>s\n"]], ["block_3", ["Use Equation 3.4.10 to evaluate [Po]S for the system under consideration:\n"]], ["block_4", ["Over the time required to reach the stationary state, the initiator concentration is essentially\nunchanged. As a matter of fact, it would take about 100 s for [I] to reach 0.99 [HO and about 8.5\nmin to reach 0.95 me, so the assumption that [I] [Hg is entirely justi\ufb01ed over the short times\ninvolved.\n"]], ["block_5", ["Consequently, the addition of the next monomer is virtually exclusively in a \u201chead-to-tail\u201d arrange-\nment, leading to an all-carbon backbone with substituents (X) on alternating backbone atoms:\n"]], ["block_6", ["90\nChain-Growth Polymerization\n"]], ["block_7", ["3.5\nPropagation\n"]], ["block_8", ["Solution\n"]], ["block_9", [{"image_1": "105_1.png", "coords": [44, 548, 306, 611], "fig_type": "molecule"}]], ["block_10", [{"image_2": "105_2.png", "coords": [45, 92, 360, 138], "fig_type": "molecule"}]], ["block_11", ["ln(\ufb02) \u2014kdt ~(1.0 >< 10\u20144)(2.16)= \u20142.16 X 10*4\n[110\n[I]\n\u2014\u2014 = 0.9998\n[Ila\n"]], ["block_12", ["1/2\n\u20144\n\u20143\nl/2\n[P 1, (@1110)\n\u2014(\n3 X107\n)\n_ (1.67 x 10\n)\n"]], ["block_13", ["= 4.08 x 10\u20148 mol L-1\n"]], ["block_14", [{"image_3": "105_3.png", "coords": [108, 549, 219, 611], "fig_type": "molecule"}]], ["block_15", [{"image_4": "105_4.png", "coords": [179, 552, 291, 603], "fig_type": "molecule"}]]], "page_106": [["block_0", [{"image_0": "106_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["in which the second form reminds us that an experimental study of the rate of polymerization\nyields a single apparent rate constant (subscript app), which the mechanism reveals to be a\ncomposite of three different rate constants. Equation 3.5.3 shows that the rate of polymerization\nis first order in monomer and half order in initiator and depends on the rate constants for each of\nthe three types of steps\u2014initiation, propagation, and termination\u2014\u201cthat make up the chain mech-\nanism. Since the concentrations change with time, it is important to realize that Equation 3.5.3\ngives an instantaneous rate of polymerization at the concentrations considered. The equation can\nbe applied to the initial concentrations of monomer and initiator in a reaction mixture only to\ndescribe the initial rate of polymerization. Unless stated otherwise, we shall assume the initial\nconditions apply when we use this result.\nThe initial rate of polymerization is a measurable quantity. The amount of polymer formed after\nvarious times in the early stages of the reaction can be determined directly by precipitating the\npolymer and weighing. Alternatively, some property such as the volume of the system (or\nthe density, the refractive index, or the viscosity) can be measured. Using an analysis similar to\nthat followed in Example 3.1, we can relate the values of the property measured at r, r: 0 and\nt: 00 to the fraction of monomer converted to polymer. If the rate of polymerization is measured\nunder known and essentially constant concentrations of monomer and initiator, then the cluster\nof constants (fkgkd/ktX/Z can be evaluated from the experiment. As noted earlier, f is best\ninvestigated by endgroup analysis. Even with the factor f excluded, experiments on the rate of\npolymerization still leave us with three unknowns. Two other measurable relationships among\nthese unknowns must be found if the individual constants are to be resolved. In anticipation of this\ndeve10pment, we list values of kp and the corresponding activation energies for several common\nmonomers in Table 3.4.\n"]], ["block_2", ["1.\nThe radical concentration has the stationary\u2014state value given by Equation 3.4.10.\n2.\nkp is a constant independent of the size of the growing chain and the extent of conversion\nto polymer.\n3.\nThe rate at which monomer is consumed is equal to the rate of polymer formation RP:\n"]], ["block_3", ["This should linkage that results from termination by\nrecombination \n"]], ["block_4", ["as the expression converted to polymer. In writing this expression,\nwe assume \n"]], ["block_5", ["Consideration (3.B) leads to\n"]], ["block_6", ["3.5.1\nRate Laws for Propagation\n"]], ["block_7", ["Propagation\n91\nW'\n+ AX\n\u2014\" WYW\n(3.1)\nX\nX\nX\nX\nX\nX\nX\n"]], ["block_8", [{"image_1": "106_1.png", "coords": [38, 331, 232, 367], "fig_type": "molecule"}]], ["block_9", [{"image_2": "106_2.png", "coords": [38, 48, 354, 90], "fig_type": "molecule"}]], ["block_10", [{"image_3": "106_3.png", "coords": [38, 52, 123, 88], "fig_type": "molecule"}]], ["block_11", ["d M\n\ufb01._[_].:kp[M][p.]\n(3.5.1)\ndt\n"]], ["block_12", ["k\u2014t)\n[111/2 kapp[M][I]1/2\n(3-5-3)\nRD kp[M]<\n"]], ["block_13", ["Combining Equation 3.4.10 and Equation 3.5.1 yields\n"]], ["block_14", ["d[M] _ d[polymer] _\ndt\nT\ndr\n_ Rp\n(3.5.2)\n"]], ["block_15", ["d\n1/2\n"]], ["block_16", [{"image_4": "106_4.png", "coords": [190, 51, 334, 86], "fig_type": "molecule"}]]], "page_107": [["block_0", [{"image_0": "107_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 3.4\nRate Constants at 60\u00b0C and Activation Energies for Some\nPropagation Reactions\n"]], ["block_2", ["92\nChain-Growth Polymerization\n"]], ["block_3", ["Monomer\nEff (kl moi\u2014l)\nk1, x 10\u20183 (L mol\u2018l s\u2018l)\n"]], ["block_4", ["Source: Data from Korus, R. and O\u2019Driscoll, K.F. in Polymer Handbook, \nBrandrup, J. and Immergut, E.H. (Eds), Wiley, New York, 1989.\n"]], ["block_5", ["Acrylonitrile\n16.2\n1.96\nMethyl acrylate\n29.7\n2.09\nMethyl methacrylate\n26.4\n0.515\nStyrene\n26.0\n0. 165\nVinyl acetate\n18.0\n2.30\n2-Vinyl pyridine\n33.0\n0.186\n"]], ["block_6", ["Figure 3.3 shows some data that constitute a test of Equation 3.5.3. In Figure 3.33, RP and [M] are\nplotted on a log\u2014log scale for a constant level of redox initiator. The slope of this line, which\nindicates the order of the polymerization with respect to monomer, is unity, showing that the\npolymerization of methyl methacrylate is first order in monomer. Figure 3.3b is a similar plot of\nthe initial rate of polymerization\u2014which essentially maintains the monomer at constant concentra\u2014\ntion\u2014versus initiator concentration for two different monomer\u2014initiator combinations. Each of the\nlines has a slope of 1/2, indicating a half-order dependence on [I] as predicted by Equation 3.5.3.\n"]], ["block_7", ["where [M] = [M]0 at t: 0.\n2.\nInstead of using 2\ufb02 \u2014(\nkt\n[HO\nt\n(3.54)\n"]], ["block_19", ["R.\n1/2\n[Pr]. (2\u20141;)\n(3.5.5)\n"]], ["block_20", [{"image_3": "107_3.png", "coords": [64, 399, 147, 449], "fig_type": "molecule"}]]], "page_108": [["block_0", [{"image_0": "108_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This enables us to identify the apparent activation energy in Equation 3.5.7 with the difference in\nE* values for the various steps:\n"]], ["block_2", ["Equation 3.5.9 allows us to conveniently assess the effect of temperature variation on the rate of\npolymerization. This effect is considered in the following example.\n"]], ["block_3", ["Using typical activation energies from Table 3.2 through Table 3.4, estimate the percent change in\nthe rate of polymerization with a 1\u00b0C change in temperature at 50\u00b0C, for both thermally initiated\nand photoinitiated polymerization.\n"]], ["block_4", ["Figure 3.3\nLog\u2014log plots of RP versus concentration that confirm the kinetic order with reSpect to the\nconstituent varied. (a) Monomer (methyl methacrylate) concentration varied at constant initiator concentra\u2014\ntion. (Data from Sugirnura, T. and Minoura, Y., J. Polym. Sci, A-l, 2735, 1966.) (b) Initiator concentration\nvaried: AIBN in methyl methacrylate (data from Amett, L.M., J. Am. Chem. Soc., 74, 2027, 1952) and\nbenzoyl peroxide in styrene (data from Mayo, F.R., Gregg, R.A., and Matheson, M.S., J. Am. Chem. 806., 73,\n1691, 1951).\n"]], ["block_5", ["The mechanistic analysis of the rate of polymerization and the fact that the separate constants\nindividually follow the Arrhenius equations means that\n"]], ["block_6", ["Propagation\n93\n"]], ["block_7", ["Example 3.3\n"]], ["block_8", [{"image_1": "108_1.png", "coords": [39, 64, 223, 289], "fig_type": "figure"}]], ["block_9", ["T\n__\n03\n\u20180\n\u2018T\n"]], ["block_10", ["1|\u2014\n_\n_l\n\"\u2018\na\nE\ng\ns.\nv\nmn-\nTO.\n10 3\n10-5:\nx\nZ\n:\na:\n\u2018\nStyrene/benzoyl peroxide at 60\u00b0C\n"]], ["block_11", ["E\n10\u20144.\u2014\n,.\nI\n\u20197\"\n'\n"]], ["block_12", ["(a)\n[M] (mol L4)\n(b)\n[He (mol L\"\u2018)\n"]], ["block_13", ["E3;\n5*\nEgg, =E\u00a7+7\u20147\u2018\n(3.5.9)\n"]], ["block_14", [{"image_2": "108_2.png", "coords": [46, 425, 250, 503], "fig_type": "figure"}]], ["block_15", ["1/2\nln km, =ln k, (73,9)\n"]], ["block_16", [{"image_3": "108_3.png", "coords": [54, 426, 243, 498], "fig_type": "molecule"}]], ["block_17", ["100 \n"]], ["block_18", [{"image_4": "108_4.png", "coords": [60, 122, 200, 266], "fig_type": "figure"}]], ["block_19", ["\u20141\nfAd \nEg\u2018+E,\u00a7\u201c/2\u2014E;\"/2\n_ \u201cAP _\n_\nRT\n"]], ["block_20", ["llllll\nI\nIIIIIIII\nI\n10_6\n....|\n........|\n.......|\n........|\n"]], ["block_21", ["1-0\n10\n1041\n10-3\n10-2\n10*1\n"]], ["block_22", [{"image_5": "108_5.png", "coords": [168, 47, 452, 277], "fig_type": "figure"}]], ["block_23", ["10\u20143:\"\"'|\n'\nIIIII\"I\n'\nI\"\"\"l\n\u2018\n\"_'_'_'\"|\n'2'\n"]], ["block_24", [{"image_6": "108_6.png", "coords": [226, 78, 438, 226], "fig_type": "figure"}]], ["block_25", ["Methyl methacrylate/AIBN at 50\u00b0C\n"]], ["block_26", ["(3.5.8)\n"]]], "page_109": [["block_0", [{"image_0": "109_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["under conditions where an initiator yields one radical, where f1, and where the final polymer\ncontains one initiator fragment per molecule. For this set of conditions the ratio gives the number\nof monomers polymerized per chain initiated, which is the average degree of polymerization.\nA more general development of this idea is based on a quantity called the kinetic chain length 17.\nThe kinetic chain length is defined as the ratio of the number of propagation steps to the number of\ninitiation steps, regardless of the mode of termination:\n"]], ["block_2", ["or 1.90% per \u00b0C. Note that the initiator decomposition makes the largest contribution to E*;\ntherefore photoinitiated processes display a considerably lower temperature dependence for the\nrate of polymerization.\n"]], ["block_3", ["3.5.3\nKinetic Chain Length\n"]], ["block_4", ["Suppose we consider the ratio\n"]], ["block_5", ["Finally, we recognize that a 1\u00b0C temperature variation can be approximated as (17' and that\n(dRp/Rp) x 100 gives the approximate percent change in the rate of polymerization. Taking average\nvalues of Efrom the appropriate tables, we obtain E3; 145, E? 16.8, and E3 24.9 k] mol\u2018l.\nFor thermally initiated polymerization\n"]], ["block_6", ["or 10.3% per \u00b0C.\nFor photoinitiation there is no activation energy for the initiator decomposition; hence\n"]], ["block_7", ["Expand d In kapp by means of the Arrhenius equation via Equation 3.5.8:\n"]], ["block_8", ["Write Equation 3.5.3 in the form\n"]], ["block_9", ["Taking the derivative, treating [M] and [I] as constants with respect to T while k is a function of T:\n"]], ["block_10", ["Substitute Equation 3.5.9 for \u00a331,131\n"]], ["block_11", ["94\nChain-Growth Polymerization\n"]], ["block_12", ["Solution\n"]], ["block_13", [{"image_1": "109_1.png", "coords": [42, 631, 109, 678], "fig_type": "molecule"}]], ["block_14", ["Rp/Ri \nd[M]/dt\n\u2014d[I]/dt\n"]], ["block_15", ["d\ufb02 _ (24.9 + 145/2 16.8/2)(103)(1) = 0.103\nRP\n(8.314)(323)2\n"]], ["block_16", ["dRp (24.9 16.8/2)(103)(1)\nRp \n(8.314)(323)2\n= 1.90 x 10-2\n"]], ["block_17", ["dRp\nE?\nE:\n\u2014=d1Aa\n\u2014\u2014d _P =__PP\nRp\n\"\nPP\n(R75\nRTZdT\n"]], ["block_18", ["dRp E; + Iii/2 133/s\nRp \nRT2\n"]], ["block_19", ["dR\n(:11a =R\u2014P= drnk...pp\n"]], ["block_20", ["1\n1n RP = 1n kapp + In [M] + 2 In [I]\n"]], ["block_21", ["= _\n(3.5.10)\n"]], ["block_22", ["P\n"]]], "page_110": [["block_0", [{"image_0": "110_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["As with the rate of polymerization, we see from Equation 3.5.14 that the kinetic chain length\ndepends on the monomer and initiator concentrations and on the constants for the three different\nkinds of kinetic processes that constitute the mechanism. When the initial monomer and initiator\nconcentrations are used, Equation 3.5.14 describes the initial polymer formed. The initial degree of\npolymerization is a measurable quantity, so Equation 3.5.14 provides a second functional rela-\ntionship, distinct from Equation 3.5.3, among experimentally available quantities\u2014ND, [M], [I]\u2014\nand theoretically important parameters\u2014-\u2014kp, kt, and kd. Note that the mode of termination, which\nestablishes the connection between\n17 and Nn, and the value off are both accessible through\nendgroup characterization. Thus we have a second equation with three unknowns; one more\nindependent equation and the evaluation of the individual kinetic constants from experimental\nresults will be feasible.\nThere are several additional points about Equation 3.5.14 that are worthy of comment. First it\nmust be recalled that we have intentionally ignored any kinetic factors other than initiation,\npropagation, and termination. We shall see in Section 3.8 that another process, chain transfer,\nhas significant effects on the molecular weight of a polymer. The result we have obtained,\ntherefore, is properly designated as the kinetic chain length without transfer. A second observation\nis that 1'\u00bb depends not only on the nature and concentration of the monomer, but also on the nature\nand concentration of the initiator. The latter determines the number of different sites competing for\nthe addition of monomer, so it is not surprising that L7 is decreased by increases in either kd or [I].\nFinally, we observe that both kp and kt are properties of a particular monomer. The relative\nmolecular weight that a speci\ufb01c monomer tends toward\u2014all other things being equal\u2014is charac\u2014\nterized by the ratio ftp/kt\u201d2 for a monomer. Using the values in Table 3.3 and Table 3.4, we see that\nkp/kt\u201d2 equals 0.678 for methyl acrylate and 0.0213 for styrene at 60\u00b0C. The kinetic chain length for\npoly(methyl acrylate) is thus expected to be about 32 times greater than for polystyrene if the two\nare prepared with the same initiator (kd) and the same concentrations [M] and [1]. Extension of this\ntype of comparison to the degree of polymerization requires that the two polymers compared show\nthe same proportion of the modes of termination. Thus for vinyl acetate (subscript V) relative to\nacrylonitrile\n(subscript A)\nat 60\u00b0C, with\nthe\nsame\nprovisos\nas\nabove,\n{xv/17A 6 while\nv/NHA = 3 because of the differences in the mode of termination for the two.\n"]], ["block_2", ["where the uses the stationary-state condition R, =Rt. The \ncance of\n"]], ["block_3", ["2.\nFor termination by combination\n"]], ["block_4", ["3.\n17 is an average quantity\u2014indicated by the overbar\u2014since not all kinetic chains are identical\nany more than all molecular chains are.\n"]], ["block_5", ["Propagation\n95\n"]], ["block_6", ["1,\nFor termination by disproportionation\n"]], ["block_7", ["This may be combined with Equation 3.4.10 to give the stationary-state value for 17:\n"]], ["block_8", ["Using Equation 3.5.3 and Equation 3.4.4 for RP and Rt, respectively, we write\n"]], ["block_9", ["where N\" is the number average degree of polymerization.\n"]], ["block_10", ["17 =Nn\n(3.5.11)\n"]], ["block_11", ["Nn\nv 3\n(3.5.12)\n"]], ["block_12", ["k\nM\nk\nM\n17 :\np[\n]\np[\n]\n(3.5.14)\nmums/M\u201d\nZ awash\u201d:\n"]], ["block_13", ["\u2014 = kp[P\u00b0][M] _ kp[M]\n_\n3.5.13\nV\n2k.[Po 12\n2mm\n(\n)\n"]]], "page_111": [["block_0", [{"image_0": "111_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In the preceding section we observed that both the rate of polymerization and the degree of\npolymerization under stationary-state conditions can be interpreted to yield some cluster of the\nconstants kp, k,, and kd. The situation is summarized diagrammatically in Figure 3.4. The circles at\nthe two bottom corners of the triangle indicate the particular grouping of constants obtainable from\nthe measurement of RF, or N\u201c, as shown. By combining these two sources of data in the manner\nsuggested in the boxes situated along the lines connecting these circles kd can be evaluated, as well\nas the ratio lag/kt. Using this stationary\u2014state data, however, it is not possible to further resolve the\npropagation and termination constants. Another relationship is needed to do this. A quantity called\nthe radical li\ufb01ztime 1\" supplies the additional relationship and enables us to move off the base of\nFigure 3.4.\nTo arrive at an expression for the radical lifetime, we return to Equation 3.5.1, which may be\ninterpreted as follows:\n"]], ["block_2", ["It is interesting to compare the application of this result to thermally initiated and photoinitiated\npolymerizations as we did in Example 3.3. Again using the average values of the constants from\nTable 3.2 through Table 3.4 and taking T= 50\u00b0C, we calculate that 17 decreases by about 6.5% per\n\u00b0C for thermal initiation and increases by about 2% per \u00b0C for photoinitiation. It is clearly the large\nactivation energy for initiator dissociation that makes the difference. This term is omitted in the\ncase of photoinitiation, where the temperature increase produces a bigger effect on propagation\nthan on termination. On the other hand, for thermal initiation an increase in temperature produces a\nlarge increase in the number of growth centers, with the attendant reduction of the average kinetic\nchain length.\nPhotoinitiation is not as important as thermal initiation in the overall picture of free-radical\nchain-growth polymerization. The foregoing discussion reveals, however, that the contrast\nbetween the two modes of initiation does provide insight into, and confirmation of, various aspects\nof addition polymerization. The most important application of photoinitiated polymerization is in\nproviding a third experimental relationship among the kinetic parameters of the chain mechanism.\nWe shall consider this in the next section.\n"]], ["block_3", ["3.6\nRadical Lifetime\n"]], ["block_4", ["The proviso \u201call other things being equal\u201d in discussing the last point clearly applies to\ntemperature as well, since the kinetic constants can be highly sensitive to temperature. To evaluate\nthe effect of temperature variation on the molecular weight of an addition polymer, we follow the\nsame sort of logic as was used in Example 3.3:\n"]], ["block_5", ["2.\nDifferentiate with respect to T, assuming the temperature dependence of the concentrations is\nnegligible compared to that of the rate constants:\n"]], ["block_6", ["96\nChain-Growth Polymerization\n"]], ["block_7", ["3.\nBy the Arrhenius equation (1 ln k = \u2014d(E*/RT) = (E*/ R72) dT; therefore\n"]], ["block_8", ["1.\n\u2014d[M]/dt gives the rate at which monomers enter polymer molecules. This, in turn, is given by\nthe product of number of growth sites, [Po], and the rate at which monomers add to each\n"]], ["block_9", ["1.\nTake logarithms of Equation 3.5.14:\n"]], ["block_10", ["cw __ 13;; Eat/2 Eat/2d],\n"]], ["block_11", ["9; dln k, 1 /2d 1h (ktkd)\n(3.5.16)\n"]], ["block_12", ["1n :7 = 1n kp(ktkd)_1/2 + 1n(%)\n(3.5.15)\n"]], ["block_13", ["I7 \nRT2\n(3.5.17)\n"]]], "page_112": [["block_0", [{"image_0": "112_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 3.4\nSchematic relationship among the various experimental quantities (RP, NH, and 1\") and the rate\nconstants kd, kp, and kt derived therefrom.\n"]], ["block_2", ["Radical Lifetime\n97\n"]], ["block_3", ["Hp\n"]], ["block_4", ["\ufb02) = kd\nN\nkd\n%\n\"\nkp\n\u2014-\u2014>\u2014\nk\n\u2014\n1\n4\u2014 N\n\"\nkt\n(mo/2\n\u201d\n"]], ["block_5", ["The degree of polymerization in Equation 3.6.1 can be replaced with the kinetic chain length,\nand the resulting expression simplified. To proceed, however, we must choose between the\npossibilities described in Equation 3.5.11 and Equation 3.5.12. Assuming termination by\ndisproportionation, we replace N,n by 17, using Equation 3.5.14:\n"]], ["block_6", ["growth site. On the basis of Equation 3.5.1, the rate at which monomers add to a radical is\ngiven by kP[M].\nIf kp[M] gives the number of monomers added per unit time, then 1/kp[M] equals the time\nelapsed per monomer addition.\nIf we multiply the time elapsed per monomer added to a radical by the number of monomers in\nthe average chain, then we obtain the time during which the radical exists. This is the\ndefinition of the radical lifetime. The number of monomers in a polymer chain is, of course,\nthe degree of polymerization. Therefore we write\n"]], ["block_7", ["The radical lifetime is an average quantity, as indicated by the overbar.\n"]], ["block_8", ["kp [M]\n(3.6.1)\n7\u2014:\n"]], ["block_9", ["2(1\u20187ctt\u2019alll)\u201d2 \n_\n2(\ufb02ctkduD1/2\n(3.6.2)\nIr:\n"]], ["block_10", [{"image_1": "112_1.png", "coords": [75, 42, 313, 369], "fig_type": "figure"}]], ["block_11", ["Nn\n"]], ["block_12", ["kpiM]\n1\n_\n1\n"]], ["block_13", [{"image_2": "112_2.png", "coords": [89, 152, 292, 309], "fig_type": "figure"}]], ["block_14", ["+57 M\n1\nes,\n\u201c\\3,\n1rt\n\u00ab3\u2018\n"]], ["block_15", ["[91%\nkt\u2018/&\n"]], ["block_16", ["kn\np\n"]], ["block_17", ["RD Nn kpe/kt\n"]], ["block_18", ["1\u2014\n(ktkdiyz\n"]], ["block_19", ["?\nl\n"]]], "page_113": [["block_0", [{"image_0": "113_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\" = 0.5[(1.0)(2.9 x 107)(0.85 x 10\u20185)(10\u20193)]\u20181/2 = 1.01s. This figure contrasts sharply with the\ntime required to obtain high-molecular-weight molecules in step-growth polymerizations.\nSince the radical lifetime provides the final piece of information needed to independently\nevaluate the three primary kinetic constants\u2014remember, we are still neglecting chain transfer\u2014\nthe next order of business is a consideration of the measurement of f. A widely used technique for\nmeasuring radical lifetime is based on photoinitiated polymerization using a light source,\nwhich blinks on and off at regular intervals. In practice, a rotating opaque disk with a wedge\nsliced out of it is interposed between the light and the reaction vessel. Thus the system is in\ndarkness when the solid part of the disk is in the light path and is illuminated when the notch\npasses. With this device, called a rotating sector or chopper, the relative lengths of light and dark\nperiods can be controlled by the area of the notch, and the frequency of the \ufb02ickering by the\nvelocity of rotation of the disk. We will not describe the rotating sector experiments in detail. It is\nsufficient to note that, with this method, the rate of photoinitiated polymerization is studied as a\nfunction of the time of illumination with the rapidly blinking light. The results show the rate of\npolymerization dropping from one plateau value at slow blink rates (\u201clong\u201d bursts of illumination)\nto a lower plateau at fast blink rates (\u201cshort\u201d periods of illumination). A plot of the rate of\npolymerization versus the duration of an illuminated interval resembles an acid\u2014base titration\ncurve with a step between the two plateau regions. Just as the \u201cstep\u201d marks the end point of a\ntitration, the \u201cstep\u201d in rotating sector data identifies the transition between relatively long and\nshort periods of illumination. Here is the payoff: \u201clong\u201d and \u201cshort\u201d times are defined relative to\nthe average radical lifetime. Thus f may be read from the time axis at the midpoint of the transition\nbetween the two plateaus.\nThis qualitative description enables us to see that the radical lifetime described by Equation\n3.6.2 is an experimentally accessible quantity. More precise values of i may be obtained by curve\nfitting since the nonstationary-state kinetics of the transition between plateaus have been analyzed\nin detail. To gain some additional familiarity with the concept of radical lifetime and to see how\nthis quantity can be used to determine the absolute value of a kinetic constant, consider the\nfollowing example:\n"]], ["block_2", ["We can use the constants tabulated elsewhere in the chapter to get an idea of a typical\nradical lifetime. Choosing 10\u201c3 mol L\u20181 AIBN as the initiator (kd=0.85 x 10\u20145 s\u201c at 60\u00b0C)\nand vinyl acetate as the monomer (terminates entirely by disproportionation, k,=2.9 x 107 L\nmol\u2019l\ns\u20181\nat\n60\u00b0C),\nand\ntaking f=1\nfor\nthe\npurpose\nof\ncalculation,\nwe\nfind\n"]], ["block_3", ["The polymerization of ethylene at 130\u00b0C and 1500 atm was studied using different concentrations\nof the initiator, 1-t-butylazo-1-phenoxycyclohexane. The rate of initiation was measured directly\n"]], ["block_4", ["1.\nIn going from the experimental quantities RP, N\u201c, and f to the associated clusters of kinetic\nconstants, it has been assumed that the monomer and initiator concentrations are known and\nessentially constant. In addition, the efficiency factorf has been left out, the assumption being\nthat still another type of experiment has established its value.\n2.\nBy following the lines connecting two sources of circled information, the boxed result in the\nperimeter of the triangle may be established. Thus kl, is evaluated from f and N\u201c.\n3.\nHere kp can be combined with one of the various kP/kt ratios to permit the evaluation of k,.\n"]], ["block_5", ["We shall see presently that the lifetime of a radical can be measured. When such an experiment is\nconducted with a known concentration of initiator, then the cluster of constants (ktkd)_u2 can be\nevaluated. This is indicated at the apex of the triangle in Figure 3.4.\nThere are several things about Figure 3.4 that should be pointed out:\n"]], ["block_6", ["98\nChain-Growth Polymerization\n"]], ["block_7", ["Example 3.4\n"]]], "page_114": [["block_0", [{"image_0": "114_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "114_1.png", "coords": [27, 335, 177, 428], "fig_type": "figure"}]], ["block_2", [{"image_2": "114_2.png", "coords": [32, 83, 194, 170], "fig_type": "figure"}]], ["block_3", ["We begin by writing a kinetic expression for the concentration of radicals of the degree of\npolymerization i, which we designate [Pp]. This rate law will be the sum of three contributions:\n"]], ["block_4", ["Until this point in the chapter we have intentionally avoided making any differentiation among\nradicals on the basis of the degree of polymerization of the radical. Now we seek a description of\nthe molecular weight distribution of addition polymer molecules. Toward this end it becomes\nnecessary to consider radicals with different i values.\n"]], ["block_5", ["Even though the rates of initiation span almost a 10-fold range, the values of kt show a standard\ndeviation of only 4%. which is excellent in view of the inevitable experimental errors. Note that the\nrotating sector method can be used in high-pressure experiments and other unusual situations, a\nhighly desirable characteristic it shares with many optical methods in chemistry.\n"]], ["block_6", ["If the data follow the kinetic scheme presented here, the values of kt calculated for the different\nruns should be constant:\n"]], ["block_7", ["and radical lifetimes using the rotating sector method. The following results were\n.\nobtained.1\u2018\n1\nRun\n1\"(s)\nR,x109(molL\u2014ls\u20141)\n"]], ["block_8", ["3.7.1\nDistribution of i-mers: Termination by Disproportionation\n"]], ["block_9", ["3.7\nDistribution of Molecular Weights\n"]], ["block_10", ["tData from T. Takahashi and P. Ehrlich, Polym. Prepr. Am. Chem. Soc. Polym. Chem. Div., 22, 203 (1981).\n"]], ["block_11", ["Since the rate of initiation is measured, we can substitute R, for the terms (kdm)\n3.6.2 to give\n"]], ["block_12", ["Run\n1ct X 10\u20143 (L mol\u20141 3'1)\n"]], ["block_13", ["13\n3.98\nAverage\n3.89\n"]], ["block_14", ["Demonstrate that the variations in the rate of initiation and\nF? are consistent with free-radical\nkinetics, \n"]], ["block_15", ["Solution\n"]], ["block_16", ["Distribution of Molecular Weights\n99\n"]], ["block_17", ["12\n4.00\n"]], ["block_18", ["12\n0.50\n5.00\n"]], ["block_19", ["13\n0.29\n14.95\n"]], ["block_20", ["5\n3.99\n"]], ["block_21", ["6\n3.64\n"]], ["block_22", ["8\n3.83\n"]], ["block_23", ["5\n0.73\n2.35\n6\n0.93\n1.59\n"]], ["block_24", ["8\n0.32\n12.75\n"]], ["block_25", ["\u2019 \u2014\n1\nor\nk \n1\nT (zany/2\n\u2018 2am,\n"]], ["block_26", ["1\u20192 in Equation\n"]]], "page_115": [["block_0", [{"image_0": "115_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "115_1.png", "coords": [28, 554, 199, 588], "fig_type": "molecule"}]], ["block_2", ["The change in [P3] under stationary\u2014state conditions equals zero for all values of 1'; hence we\n"]], ["block_3", ["The ratio [P1-]/[P-] in Equation 3.7.8 can be eliminated by applying Equation 3.7.1 explicitly to the\n"]], ["block_4", ["Dividing both sides of Equation 3.7.7 by [P-], the total radical concentration, gives the number\n(or mole) fraction of i-rner radicals in the total radical population. This ratio is the same as the\nnumber of i-mers n,- in the sample containing a total of n (no subscript) polymer molecules:\n"]], ["block_5", ["1.\nAn increase that occurs by addition of monomer to the radical P,-_1-\n2.\nA decrease that occurs by addition of a monomer to the radical P).\n3.\nA decrease that occurs by the termination of P,\u00bb with any other radical P-\n"]], ["block_6", ["or\n"]], ["block_7", ["This shows that the number of i-mer radicals relative to the number of the smallest radicals is given\nby multiplying the ratio [P,~]/ [P,-_1\u00b0] by i 2 analogous ratios. Since each of the individual ratios\nis given by 17/(1 + 17), we can rewrite Equation 3.7.4 as\n"]], ["block_8", ["P1 radical:\n"]], ["block_9", ["Since it is more convenient to focus attention on i-mers than (i\u20141)\u2014mers, the corresponding\nexpression for the i\u2014mer is written by analogy:\n"]], ["block_10", ["can write\n"]], ["block_11", ["Dividing the numerator and denominator of Equation 3.7.2 by 2kt [Po] and recalling the de\ufb01nition\nof 17 provided by Equation 3.5.13 enables us to express this result more succinctly as\n"]], ["block_12", ["Next let us consider the following sequence of multiplications:\n"]], ["block_13", ["100\nChain-Growth Polymerization\n"]], ["block_14", ["which can be rearranged to\n"]], ["block_15", ["1.\nWrite Equation 3.7.1 for Pp, remembering in this case that the leading term describes initiation:\n"]], ["block_16", [{"image_2": "115_2.png", "coords": [37, 358, 154, 396], "fig_type": "molecule"}]], ["block_17", [{"image_3": "115_3.png", "coords": [42, 409, 170, 446], "fig_type": "molecule"}]], ["block_18", ["Mn.-_[P.--1_[P1-1\n:7\n\u201c1\nx\u2018 \u2018Z\" [13-] [P-] (1+ :7)\n(3'7\u20188)\n"]], ["block_19", ["d[P.-'l\n \u201821? \n(311)\n"]], ["block_20", ["[Pi\u20141\"]\nh mm + 2kt[P-]\n(3.7.2)\n"]], ["block_21", ["Pi'\nPi\u2014'\nPi\u2014'\nPi\u2014i\u2014\n\u00b0\nPr'\n[ ll\n1][\n2]_\u201d[\n(2)]_[\nl\n(314)\n[Pi\u20141'] [Pi\u20142'] [Pi\u20143\u00b0]\n[Pi\u2014(i\u2014l)'] _[P1']\n"]], ["block_22", ["_\ni\u2014l\n[Pr] [Pi-1 (\ufb01g)\n(3.7.7)\n"]], ["block_23", ["[Pg\u20141'] \n(3.7.6)\n"]], ["block_24", ["[[151]] [15:] (1:12)\n(3.7.5)\n"]], ["block_25", ["[Pr]\nI7\n:\n3.7.3\n[Par]\n1 + 17\n(\n)\n"]], ["block_26", ["[Pi-1 _\nkp[M]\n"]], ["block_27", ["93;? = Ri kp[M][P1-] \u2014 2kt[P1-][P-] : 0\n(3.7.9)\n"]], ["block_28", ["17\n0\u20140\u20141\n"]], ["block_29", ["_\ni\u20142\n"]]], "page_116": [["block_0", [{"image_0": "116_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This change of notation now expresses Equation 3.7.14 in exactly the same form as its equivalent\nin Section 2.4. In other words, the distribution of chain lengths is the most probable distribution,\njust as was the case for step-growth polymerization! Several similarities and differences should be\nnoted in order to take full advantage of the parallel between this result and the corresponding\nmaterial for condensation polymers in Chapter 2:\n"]], ["block_2", ["This expression gives the number fraction or mole fraction, x,, of i-mers in the polymer and is thus\nequivalent to Equation 2.4.2 for step-growth polymerization.\nThe kinetic chain length 13 may also be viewed as merely a cluster of kinetic constants and\nconcentrations, which was introduced into Equation 3.7.13 to simplify the notation. As an\nalternative, suppose we define for the purposes of this chapter a fraction p such that\n"]], ["block_3", ["4,\nRearrange under stationary\u2014state conditions:\n"]], ["block_4", ["1.\nIn Chapter 2, p was defined as the fraction (or probability) of functional groups that had\nreacted at a certain point in the polymerization. According to the current definition provided\nby Equation 3.7.15, p is the fraction (or probability) of propagation steps among the com-\nbined total of propagation and termination steps. The quantity 1 p is therefore the fraction\n(or probability) of termination steps. An addition polymer with the degree of polymerization\ni has undergone i 1 propagation steps and one termination step. Therefore it makes sense to\ndescribe its probability in the form of Equation 3.7.16.\n2.\nIt is apparent from Equation 3.7.15 that p \u20149\n1 as\n17 \u2014> oo; hence those same conditions\nthat favor the formation of a high-molecular-weight polymer also indicate p values close\nto unlty.\n"]], ["block_5", ["It follows from this definition that 1/(1 + 17) 1 [9, so Equation 3.7.14 can be rewritten as\n"]], ["block_6", ["2,\nRearrange under stationary\u2014state conditions:\n"]], ["block_7", ["3,\nWrite Equation 3.4.9 using the same notation for initiation as in Equation 3.7.9:\n"]], ["block_8", ["5.\nTake the ratio of Equation 3.7.10 to Equation 3.7.12:\n"]], ["block_9", ["Distribution of \n101\n"]], ["block_10", ["\u201di\nr\u2014r\nx,- 2;:\n(1 \u2014p)p\n(3.7.16)\n"]], ["block_11", [":7\n_\nkp[M]\n1+ :7 \nkp1M1+ 2k11P-1\np 2\n(3.7.15)\n"]], ["block_12", [{"image_1": "116_1.png", "coords": [52, 294, 230, 326], "fig_type": "molecule"}]], ["block_13", ["n,-\n1\nr\u00bb\nH\n1\n:7\n\"\n;=\u2014=\n_\n=\u2014\n.z\nx\nn\n1+5<1+v>\na<1+v>\n(3\n14)\n"]], ["block_14", ["Egj=Ri \u20142kt[P-]2:0\n(3.7.11)\n"]], ["block_15", ["P \n2k P-\n1\n\u2014\u2014[\n1 ]\u2014\n\u2018[\n1\n(3.7.13)\n[P-1 \nkp1M1+ \n1 + 17\n"]], ["block_16", ["Combining Equation 3.7.13 with Equation 3.7.8 gives\n"]], ["block_17", ["The total radical concentration under stationary-state conditions can be similarly obtained.\n"]], ["block_18", ["R1\n' =\n3.7.12\n[P 1\nM1,]\n(\n)\n"]], ["block_19", ["R1\n[Pl'] kptM1+ 2k11P-1\n(3.7.10)\n"]]], "page_117": [["block_0", [{"image_0": "117_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "117_1.png", "coords": [26, 464, 178, 506], "fig_type": "molecule"}]], ["block_2", ["3.\nIn Chapter 2 all molecules\u2014whether monomer or i-mers of any i\u2014carry functional groups;\nhence the fraction described by Equation 2.4.] applies to the entire reaction mixture. Equation\n3.7.16, by contrast, applies only to the radical population. Since the radicals eventually end up\nas polymers, the equation also describes the polymer produced. Unreacted monomers are\nspecifically excluded, however.\n4.\nOnly one additional stipulation needs to be made before adapting the results that follow from\nEquation 2.4.1 to addition polymers. The mode of termination must be specified to occur by\ndisproportionation to use the results of Section 2.4 in this chapter, since termination by\ncombination obviously changes the molecular weight distribution. We shall return to the\ncase of termination by combination presently.\n5.\nFor termination by disproportionation (subscript d), we note thatp kp[M]/(kP[M] + 2k d[ P D,\nand therefore by analogy with Equation 2.4.5, Equation 2.4.9, and Equation 2.4.10,\n"]], ["block_3", ["To deal with the case of termination by combination, it is convenient to write some reactions by\nwhich an i-mer might be formed. Table 3.5 lists several specific chemical reactions and the\ncorresponding rate expressions as well as the general form for the combination of an (i j)-mer\nand a j-mer. On the assumption that all km, values are the same, we can write the total rate of\nchange of [Pi]:\n"]], ["block_4", ["The\nfraction\nof\ni-mers\nformed\nby\ncombination\nmay\nbe\nevaluated\nby\ndividing\nd[P,-]/dt by Z, d[P,-]/dt. Assuming that termination occurs exclusively by combination, then\n"]], ["block_5", ["3.7.2\nDistribution of i-mers: Termination by Combination\n"]], ["block_6", ["and the number or mole fraction of i\u2014mers formed by combination (subscript c) is\n"]], ["block_7", ["Equation 3.7.16 can be used to relate [Pi\u20141'] and [Pl-o] to the total radical concentration:\n"]], ["block_8", ["102\nChain-Growth Polymerization\n"]], ["block_9", ["[Pr\u2014f] (1 \u2014p)p\"\u201c\u201d-\u2018[Po]\n(3.7.23)\n"]], ["block_10", ["(3.7.22)\n (iii): \nd[P;]/dt\nzktacz,;:[Pt-j'][Prl\n"]], ["block_11", ["d P,-\ni\u201c\n(El\u2014t1) = kt, Z; [Pr\u20141\"][PJ\u2018]\n(3720)\n[0!\nj:\n"]], ["block_12", ["By virtue of Equation 3.7.15, (Nn)a can also be written as\n1 + 17 \u201c=-\u2019 i! for large 17, which is\nthe result already obtained in Equation 3.5.11. Figure 2.5 and Figure 2.6 also describe the\ndistribution by number and weight of addition polymers, if the provisos enumerated above\nare applied.\n"]], ["block_13", ["1\n(Nn)d\n- to\n(3.7.17)\n"]], ["block_14", ["1\n(Nw)d = _1__'_\"_P\n(3.7.18)\n\u2018P\n"]], ["block_15", ["1'\n"]], ["block_16", ["c\nEdna-yd:\nkart\u00bb?\nl1\n"]], ["block_17", ["Nw\n(\u2014)=1+p\u2014>2\nas\np\u2014>1\n(3.7.19)\nNn\nd\n"]], ["block_18", ["(1 P,-\n[dt] \n(3.7.21)\n"]], ["block_19", [{"image_2": "117_2.png", "coords": [72, 587, 239, 620], "fig_type": "molecule"}]]], "page_118": [["block_0", [{"image_0": "118_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "118_1.png", "coords": [11, 87, 278, 204], "fig_type": "figure"}]], ["block_2", [{"image_2": "118_2.png", "coords": [28, 384, 180, 415], "fig_type": "molecule"}]], ["block_3", ["Table 3.5\nSome Free-Radical Combination Reactions\n"]], ["block_4", ["Which \n"]], ["block_5", ["Therefore\n"]], ["block_6", ["The indexj drops out of the last summation; we compensate for this by multiplying the final result\nby i\u2014l in recognition of the fact that the summation adds up i\u2014l identical terms. Accordingly, the\ndesired result is obtained:\n"]], ["block_7", ["This expression is plotted in Figure 3.5 for several values ofp near unity. Although it shows the\nnumber distribution of polymers terminated by combination, the distribution looks quite different\nfrom Figure 2.5, which describes the number distribution for termination by disproportionation. In\nthe latter x,- decreases monotonically with increasing 1'. With combination, however, the curves go\nthrough a maximum, which re\ufb02ects the fact that the combination of two very small or two very\nlarge radicals is a less probable event than a more random combination.\nExpressions for the various averages are readily derived from Equation 3.7.26 by procedures\nidentical to those used in Section 2.4 (see Problem 6). We only quote the final results for the case\nwhere termination occurs exclusively by combination:\n"]], ["block_8", ["These various expressions differ from their analogs in the case of termination by disproportiona-\ntion by the appearance of occasional 2\u2019s. These terms arise precisely because two chains are\ncombined\nin\nthis\nmode\nof\ntermination.\nAgain\nusing\nEquation\n3.7.15,\nwe\nnote\nthat\n(Nu)c 2(1\n\u2014|\u2014 17) 9\u2014: 217 for large 17, a result that was already given as Equation 3.5.12.\n"]], ["block_9", ["Reaction\nRate law\n"]], ["block_10", [",\n.\no\n\u2014>\n.\nCl Pf\nPM \n1:\"\n%=k.,c[P.-_1.][Pr-i\n"]], ["block_11", ["d P,\nP;_2\u00b0 + Pz\u2018 Pr\n% kt,c[Pi\u20142\u00b0][P2\u00b0]\n"]], ["block_12", ["d P,-\nP,-_3\u00b0 + P3\u00b0 Pr\n% kt,c[Pi\u20143\u00b0][P3\u00b0]\n"]], ["block_13", [":\nd P:\n:\nP54 + P,\u00bb P.-\n% kr,c[P.-_j\u00b0][Py]\n"]], ["block_14", ["and\n"]], ["block_15", [{"image_3": "118_3.png", "coords": [35, 264, 271, 315], "fig_type": "molecule"}]], ["block_16", ["Distribution \n103\n"]], ["block_17", ["2\n(Nw)c 1\u2014H\n(3.7.28)\n\u2014 P\n"]], ["block_18", ["2\n(Nn)c If\n(3.7.27)\n"]], ["block_19", ["[Pj'] (1 p)p0\u2018\u201d[P'1\n(3.7.24)\n"]], ["block_20", ["(3) =x,- (r 1)(1 10)?\u201d\n(3.7.26)\nn\nc\n"]], ["block_21", ["Nw\n_ \n(N\u2014nl\u2014\n2\n(3.7.29)\n"]], ["block_22", ["(m)\nk... 2:1 (1 mp\u2018 1[P\u00b0](1 mp! l[Pu\n"]], ["block_23", ["n c\n\u2014\n1c.,.,[i?\u00ab]2\n"]], ["block_24", ["f\u2014l\n(3.7.25)\n= Z (1 mid\u20142\nj=1\n"]], ["block_25", ["i\u2014l\n-_-_\n._\n"]]], "page_119": [["block_0", [{"image_0": "119_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "119_1.png", "coords": [31, 55, 273, 247], "fig_type": "figure"}]], ["block_2", [{"image_2": "119_2.png", "coords": [32, 357, 185, 387], "fig_type": "molecule"}]], ["block_3", ["Figure 3.5\nMole fraction of i-mers as a function of i for termination by combination, according to Equation\n3.7.26, for various values of p.\n"]], ["block_4", ["One rather different result that arises from the case of termination by combination is seen by\nexamining the limit of Equation 3.7.29 for large values of p:\n"]], ["block_5", ["This contrasts with a limiting ratio of 2 for the case of termination by disproportionation. Since Mn\nand MW can be measured, the difference is potentially a method for determining the mode of\ntermination in a polymer system. In most instances, however, termination occurs by some\nproportion of both modes. Furthermore, other factors in the polymerization such as transfer,\nautoacceleration, etc., will also contribute to the experimental molecular weight distribution, so\nin general it is risky to draw too many conclusions about mechanisms from the measured\ndistributions. Also, we have used p and 17 to describe the distribution of molecular weights, but\nit must be remembered that these quantities are de\ufb01ned in terms of various concentrations and\ntherefore change as the reactions proceed. Accordingly, the results presented here are most simply\napplied at the start of the polymerization reaction when the initial concentrations of monomer and\ninitiator can be used to evaluate p or 17.\n"]], ["block_6", ["The three-step mechanism for free-radical polymerization represented by Reaction (3.A) through\nReaction (3.C) does not tell the whole story. Another type of free\u2014radical reaction, called chain\ntransfer, may also occur. This is unfortunate in the sense that it further complicates the picture\npresented so far. On the other hand, this additional reaction can be turned into an asset in\nactual polymer practice. One consequence of chain transfer reactions is a lowering of the kinetic\nchain length and hence the molecular weight of the polymer, without necessarily affecting the\nrate of polymerization. A certain minimum average molecular weight is often needed to achieve\na desired physical property, but further increases in chain length simply make processing\nmore dif\ufb01cult.\n"]], ["block_7", ["3.8\nChain Transfer\n"]], ["block_8", ["104\nChain-Growth Polymerization\n"]], ["block_9", ["0,004llll\u2018l\u2018l\u2014I\u2014IIIIIIIIIIIIIIIIIIIIII\n"]], ["block_10", ["IIIIIIIIII'IIIIIIIIIIIIIIIIIII\n0.001\n"]], ["block_11", ["_,><\n"]], ["block_12", ["N\u201c.\n2+1\n_)_....\u2014\nNn\n2\n21.5\nas\np\u2014>l\n(3.7.30)\n"]], ["block_13", ["0\n200\n.\n400\n600\nt\n"]], ["block_14", ["p 0.995\n"]]], "page_120": [["block_0", [{"image_0": "120_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["involves Chain transfer \nmolecule in system. reactions specifically describe transfer to initiator, mono-\nmer, solvent, and respectively:\n"]], ["block_2", ["If the rate constant is comparable to kp, the substitution of a polymer radical with a new radical\nhas little or no effect on the rate of polymerization. If kR << kp, the rate of polymerization will be\ndecreased, or even effectively suppressed by chain transfer.\nThe kinetic chain length acquires a slightly different definition in the presence of chain transfer.\nInstead of being simply the ratio Rp/Rt, it is redefined to be the rate of propagation relative to the\nrates of all other steps that compete with propagation; specifically, termination and transfer\n(subscript tr):\n"]], ["block_3", ["4.\nTransfer to polymer, PJ-X:\n"]], ["block_4", ["designated Thus if\n"]], ["block_5", ["The transfer reactions follow second\u2014order kinetics, the general rate law being\n"]], ["block_6", ["3.8.1\nChain Transfer Reactions\n"]], ["block_7", ["Chain transfer arises orsome other atom X is transferred from a molecule in the\nsystem to the growth the original radical but it \n"]], ["block_8", ["a new one: the fragment species from which X was extracted. These latter molecules will be\n"]], ["block_9", ["It is apparent from these reactions how chain transfer lowers the molecular weight of a chain-\ngrowth polymer. The effect of chain transfer on the rate of polymerization depends on the rate at\nwhich the new radicals reinitiate polymerization:\n"]], ["block_10", ["chain Transfer\n105\n"]], ["block_11", ["2.\nTransfer to monomer, MX:\n"]], ["block_12", ["3.\nTransfer to solvent, SX:\n"]], ["block_13", ["where ktr is the rate constant for chain transfer to a specific compound RX. Since chain transfer can\noccur with several different molecules in the reaction mixture, Equation 3.8.1 becomes\n"]], ["block_14", ["5.\nGeneral transfer to RX:\n"]], ["block_15", ["1.\nTransfer to initiator, IX:\n"]], ["block_16", ["Ru kH[P-][RX]\n(3.8.2)\n"]], ["block_17", ["kP\nR- + M 3 RM- \u2014+\u2014>\u2014> RP,--\n(3.P)\n"]], ["block_18", ["RP\n\u2018\nr \n3.8.1\nVt\nRt +RH\n(\n)\n"]], ["block_19", ["a. kplP'HMl/l 2am? + ktr,1[P-][IX] + ku,M[P\u00b0][MX] + ku,s[P'][SX] + km. [Para-X1}\n2\narm\n2d-1 + Z kmalRX]\n"]], ["block_20", ["Pi- + IX \u2014> + I-\n(3.K)\n"]], ["block_21", ["P;- + SX \u2014> PgX \u2014|\u2014 8-\n(3M)\n"]], ["block_22", ["P; + MX \u2014> Pix \u2014|\u2014 M-\n(3L)\n"]], ["block_23", ["P,- + PjX \u2014> PIX + Pj'\n(3N)\n"]], ["block_24", ["Pr + RX Pix + R-\n(3.0)\n"]], ["block_25", ["(3.8.3)\n"]]], "page_121": [["block_0", [{"image_0": "121_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "121_1.png", "coords": [25, 258, 158, 303], "fig_type": "molecule"}]], ["block_2", [{"image_2": "121_2.png", "coords": [32, 82, 168, 123], "fig_type": "molecule"}]], ["block_3", ["Since the first term on the right-hand side is the reciprocal of the kinetic chain length in the absence\nof transfer, this becomes\n"]], ["block_4", ["This notation is simplified still further by defining the ratio of constants\n"]], ["block_5", ["It is apparent from this expression that the larger the sum of chain transfer terms becomes, the\nsmaller will be it\u201d.\nThe magnitude of the individual terms in the summation depends on both the speci\ufb01c chain\ntransfer constants and the concentrations of the reactants under consideration. The former are\ncharacteristics of the system and hence quantities over which we have little control; the latter can\noften be adjusted to study a particular effect. For example, chain transfer constants are generally\nobtained under conditions of low conversion to polymer where the concentration of polymer is low\nenough to ignore the transfer to polymer. We shall return below to the case of high conversions\nwhere this is not true.\n"]], ["block_6", ["If an experimental system is investigated in which only one molecule is significantly involved in\ntransfer, then the chain transfer constant to that material is particularly straightforward to obtain. If\nwe assume that species SX is the only molecule to which transfer occurs, Equation 3.8.7 becomes\n"]], ["block_7", ["This suggests that polymerizations should be conducted at different ratios of [SX]/[M] and the\nresulting molecular weight measured for each. Equation 3.8.8 indicates that a plot of 1 \ufb02y, versus\n[SX]/[M] should be a straight line with s10pe CSX. Figure 3.6 shows this type of plot for the\npolymerization of styrene at 100\u00b0C in the presence of four different solvents. The fact that all show\n"]], ["block_8", ["a common intercept as required by Equation 3.8.8 shows that the rate of initiation is unaffected by\nthe nature of the solvent. The following example examines chain transfer constants evaluated in\nthis situation.\n"]], ["block_9", ["Estimate the chain transfer constants for styrene to isopropylbenzene, ethylbenzene, toluene, and\nbenzene from the data presented in Figure 3.6. Comment on the relative magnitude of these\nconstants in terms of the structure of the solvent molecules.\n"]], ["block_10", ["where the summation is over all pertinent RX species. It is instructive to examine the reciprocal of\nthis quantity:\n"]], ["block_11", ["which is called the chain transfer constant for the monomer in question to molecule RX:\n"]], ["block_12", ["3.8.2\nEvaluation of Chain Transfer Constants\n"]], ["block_13", ["106\nChain-Growth Polymerization\n"]], ["block_14", ["Example 3.5\n"]], ["block_15", [{"image_3": "121_3.png", "coords": [46, 155, 148, 192], "fig_type": "molecule"}]], ["block_16", ["1\n1\n[RX]\n\u2014 = \u2014\nC\n\u2014~\u2014\u2014\n3.8.7\na.\nr\u00bb Z\nRX [M]\n(\n)\nall RX\n"]], ["block_17", ["klt_,R: CRx\n(3.8.6)\nk1:\n"]], ["block_18", ["17\u201c\nf) [M]\n(3.8.8)\n"]], ["block_19", ["1 _1+Zktr,R[RXI\n(3 5)\n17,,\na\nkp[M]\n'\n\u2018\n"]], ["block_20", [".\nk,\nRX\n_i=2kI[P]+Z\n,R[\n]\n(3.8.4)\nVtr\nkl]\nkl]\n"]], ["block_21", ["1\n1\n[SX]\n"]]], "page_122": [["block_0", [{"image_0": "122_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "122_1.png", "coords": [34, 42, 259, 247], "fig_type": "figure"}]], ["block_2", ["In certain commercial processes it is essential to regulate the molecular weight of the polymer either\nfor ease of processing or because low molecular weight products are desirable for particular applica-\ntions such as lubricants or plasticizers. In such cases the solvent or chain transfer agent is chosen and its\nconcentration selected to produce the desired value of \ufb01n. Certain mercaptans have particularly large\nchain transfer constants for many common monomers and are especially useful for molecular weight\nregulation. For example, styrene has a chain transfer constant for n-butyl mercaptan equal to 21 at 60\u00b0C.\nThis is about 107 times larger than the chain transfer constant to benzene at the same temperature.\nChain transfer to initiator or monomer cannot always be ignored. It may be possible, however,\nto evaluate these transfer constants by conducting a similar analysis on polymerizations without\nadded solvent or in the presence of a solvent for which Csx is known to be negligibly small. Fairly\n"]], ["block_3", ["cSX x 104\n2.08\n1.38\n0.55\n0.16\n"]], ["block_4", ["The relative magnitudes of these constants are consistent with the general rule that benzylic\nhydrogens are more readily abstracted than those attached directly to the ring. The reactivity of\nthe benzylic hydrogens themselves follows the order tertiary > secondary > primary, which is a\nwell\u2014established order in organic chemistry. The benzylic radical resulting from hydrogen abstrac-\ntion is resonance stabilized. For toluene, as an example,\n"]], ["block_5", ["SX\nl\"(:31'I'KCsHs)\nC2H5(C6Hs)\nCH3(C6H5)\nH(C6Hs)\n"]], ["block_6", ["The chain transfer constants are given by Equation 3.8.8 the slopes of the lines in Figure 3.6.\nThese are estimated to be as follows (note that X H in this case):\n"]], ["block_7", ["Figure 3.6\nEffect of chain transfer to solvent according to Equation 3.8.8 for polystyrene at 100\u00b0C.\nSolvents used were ethylbenzene isopropylbenzene (O), (A), and benzene (D). (Data from\nGregg, R.A. and Mayo, F.R., Faraday Soc., 2, 328, 1947. permission).\n"]], ["block_8", ["Chain Transfer\n107\n"]], ["block_9", ["Solution\n"]], ["block_10", [{"image_2": "122_2.png", "coords": [37, 42, 148, 235], "fig_type": "figure"}]], ["block_11", ["X105\n"]], ["block_12", ["Nn\n"]], ["block_13", ["H\nH\nH\nH\nH\nH\nH\nH\n"]], ["block_14", ["200\n"]], ["block_15", ["150\n"]], ["block_16", ["100\n"]], ["block_17", ["50\n"]], ["block_18", [{"image_3": "122_3.png", "coords": [98, 481, 296, 527], "fig_type": "molecule"}]], ["block_19", ["[SX]I[M]\n"]]], "page_123": [["block_0", [{"image_0": "123_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This procedure is used commercially to produce rubber-modi\ufb01ed or high impact polystyrene (HIPS).\nThe polybutadiene begins to segregate from the styrene as it polymerizes (see Chapter 7 to learn\nwhy), but is prevented from undergoing macroscopic phase separation due to the covalent linkages to\npolystyrene chains. Consequently, small (micron\u2014sized) domains of polybutadiene rubber are\ndistributed throughout the glassy polystyrene matrix. These \u201crubber balls\u201d are able to dissipate\nenergy effectively (see Chapter 10 and Chapter 12), and counteract the brittleness of polystyrene.\nA second example of chain transfer to polymer is provided by the case of polyethylene. In this\ncase the polymer product contains mainly ethyl and butyl side chains. At high conversions such\nside chains may occur as often as once every 15 backbone repeat units on the average. These short\nside chains are thought to arise from transfer reactions with methylene hydrogens along the same\npolymer chain. This process is called backbiting and reminds us of the stability of rings of certain\nsizes and the freedom of rotation around unsubstituted bonds:\n"]], ["block_2", ["As noted above, chain transfer to polymer does not interfere with the determination of other transfer\nconstants, since the latter are evaluated at low conversions. In polymer synthesis, however, high\nconversions are desirable and extensive chain transfer can have a dramatic effect on the properties of\nthe product. This comes about since chain transfer to polymer introduces branching into the product:\n"]], ["block_3", ["extensive tables of chain transfer constants have been assembled on the basis of investigations of\nthis sort. For example, the values of CM}; for acrylamide at 60\u00b0C is 6 X 10\u2018s, and that for vinyl\nchloride at 30\u00b0C is 6.3 X 1041. Likewise, for methyl methacrylate at 60\u00b0C, CD; is 0.02 to benzoyl\nperoxide and 1.27 to t-butyl hydroperoxide.\n"]], ["block_4", [{"image_1": "123_1.png", "coords": [34, 179, 254, 212], "fig_type": "molecule"}]], ["block_5", ["A moment\u2019s re\ufb02ection reveals that the effect on 17 of transfer to polymer is different from the\neffects discussed above inasmuch as the overall degree of polymerization is not decreased by such\ntransfers. Investigation of chain transfer to polymer is best handled by examining the extent of\nbranching in the product. We shall not pursue the matter of evaluating the transfer constants, but\nshall consider describing two important specific examples of transfer to polymer.\nRemember from Section 1.3 that graft copolymers have polymeric side chains that differ in the\nnature of the repeat unit from the backbone. These can be prepared by introducing a prepolymer-\nized sample of the backbone polymer into a reactive mixture\u2014i.e., one containing a source of free\nradicals\u2014of the side-chain monomer. As an example, consider introducing 1,4-polybutadiene into\na reactive mixture of styrene:\n"]], ["block_6", ["3.8.3\nChain Transfer to Polymer\n"]], ["block_7", ["108\nChain-Growth Polymerization\n"]], ["block_8", ["\"\"w\nH\n\u00b0\nH\n\u201dI\u201c\nMe\nU\n"]], ["block_9", ["N+P.-\u2014+W+PX\u2014-+\u2014M\u2014F\u2014\u2014+\n"]], ["block_10", [{"image_2": "123_2.png", "coords": [47, 344, 213, 387], "fig_type": "molecule"}]], ["block_11", ["Y\nY\nx\n(3Q)\nYPJ-\n"]], ["block_12", [{"image_3": "123_3.png", "coords": [81, 534, 347, 686], "fig_type": "figure"}]], ["block_13", ["+\n.\n__;...\n+ M __... .S\ufb02\u2019\ufb01ne\u2014\ufb01\n(3.R)\n"]], ["block_14", ["\u2014\u2014\u2014)--\n+\nH20: CH2\u2014\u2014\u2014-I-\n"]], ["block_15", [{"image_4": "123_4.png", "coords": [118, 333, 397, 387], "fig_type": "molecule"}]], ["block_16", ["n\nMe\nO\n"]], ["block_17", [{"image_5": "123_5.png", "coords": [132, 343, 341, 378], "fig_type": "molecule"}]], ["block_18", ["l\nnHQC= CH2\n"]], ["block_19", [{"image_6": "123_6.png", "coords": [210, 182, 370, 211], "fig_type": "molecule"}]], ["block_20", [{"image_7": "123_7.png", "coords": [212, 539, 331, 592], "fig_type": "molecule"}]], ["block_21", ["1\n"]], ["block_22", ["etc.\n"]], ["block_23", [{"image_8": "123_8.png", "coords": [295, 330, 373, 389], "fig_type": "molecule"}]], ["block_24", ["/n\n"]], ["block_25", ["(3.8)\n"]]], "page_124": [["block_0", [{"image_0": "124_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "124_1.png", "coords": [31, 324, 318, 396], "fig_type": "figure"}]], ["block_2", ["However, transfer long-chain branches. The \nknown as low density is formed by mechanism in process \n"]], ["block_3", ["p\u2014benzoquinone is 518, and that for 02 is 1.5 x 104. The Polymer Handbook [1] is an excellent\nsource for these and most other rate constants discussed in this chapter.\n"]], ["block_4", ["at high \ntherefore have \nproperties of\n"]], ["block_5", ["We conclude this an extreme case of chain transfer, a reaction that produces\nradicals suppressed. \naccomplish \nduring inhibitors, depending \nof protection afford. be removed from monomers before use, and failure\n"]], ["block_6", ["and retarders differ in whichthey interfere with polymerization, but not in their\nessential activity. substance blocks polymerization \nuntil it is either removed failure to totally eliminate an inhibitor from puri\ufb01ed\nmonomer is first converted \nform before polymerization A retarder is and merely slows down the\npolymerization process by competing for radicals.\nBenzoquinone [III] is widely used as an inhibitor:\n"]], ["block_7", ["3.8.4\nSuppressing Polymerization\n"]], ["block_8", ["The resulting radical is stabilized by electron delocalization and eventually reacts with either\nanother inhibitor radical by combination (dimerization) or disproportionation or with an inhibitor\nor other radical. Another commonly used inhibitor is 2,6-di-tert\u2014butyl\u20144-methylphenol (butylated\nhydroxytoluene, or BHT):\n"]], ["block_9", ["Chain \n109\n"]], ["block_10", ["to achieve \n"]], ["block_11", ["Inhibitors are characterized by inhibition constants, which are defined as the ratio of the rate\nconstant for transfer to inhibitor to the propagation constants for the monomer, by analogy\nwith Equation 3.8.6 for chain transfer constants. For styrene at 50\u00b0C the inhibition constant of\n"]], ["block_12", ["which is also known as an antioxidant. Such free-radical scavengers often act as antioxidants, in\nthat the first stage of oxidative attack generates a free radical.\nMolecular oxygen contains two unpaired electrons and has the distinction of being capable of\nboth initiating and inhibiting polymerization. Molecular oxygen functions in the latter capacity by\nforming the relatively unreactive peroxy radical:\n(h+M~+M\u2014O\u2014O-\n8U)\n"]], ["block_13", [{"image_2": "124_2.png", "coords": [38, 462, 124, 531], "fig_type": "molecule"}]], ["block_14", ["Me \nOH\nMeMe\nMe\nMe\n"]], ["block_15", [{"image_3": "124_3.png", "coords": [48, 333, 338, 380], "fig_type": "molecule"}]], ["block_16", ["0\nnm\n"]], ["block_17", ["O\n"]], ["block_18", [".\n.\n\u2014.\u2014.\u2014\nO\nO \n\u2014> Inert products\n0-\u201c\n+\n.4o\n"]], ["block_19", ["Me\n"]]], "page_125": [["block_0", [{"image_0": "125_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1.\nIn comparison with step-growth polymerization, free-radical polymerization can lead to much\nhigher molecular weights and in much shorter times, although the resulting distributions of\nmolecular weight are comparably broad.\n2.\nThere are three essential reaction steps in a chain-growth polymerization: initiation, propagation,\nand termination. A wide variety of free-radical initiators are available; the most common act by\nthermally induced cleavage of a peroxide or azo linkage. Propagation occurs by head-to-tail\naddition of a monomer to a growing polymer radical, and is typically very rapid. Termination\noccurs by reaction between two radicals, either by direct combination or by disproportionation.\n3.\nA fourth class of reactions, termed transfer reactions, is almost always important in practice.\nThe primary effect of transfer of a radical from a growing chain to another molecule is to\nreduce the average degree of polymerization of the resulting polymer chains, but in some cases\nit can also lead to interesting architectural consequences in the \ufb01nal polymer.\n4.\nKinetic analysis of the distribution of chain lengths is made tractable by three key assump-\ntions. The steady-state approximation requires that the net rates of initiation and termination\nbe equal; thus the total concentration of radicals is constant. The same approximation extends\nto the concentration of each radical species individually. The principle of equal reactivity\nasserts that a single rate constant describes each propagation step and each termination step,\nindependent of the degree of polymerization of the radicals involved. Thirdly, transfer\nreactions are assumed to be absent.\n5.\nThe aforementioned assumptions are most successful in describing the early stages of poly-\nmerization, before a host of competing factors become significant, such as depletion of\nreactants, loss of mobility of chain radicals, etc. Under these assumptions explicit expressions\nfor the number and weight distribution of polymer chains can be developed. In the case that\ntermination occurs exclusively by disproportionation, the result is a most probable distribution\nof molecular weights, just as with step-growth polymerization. Termination by recombination,\non the other hand, leads to a somewhat narrower distribution, with MW/Mn m 1.5 rather than 2.\n"]], ["block_2", ["In this chapter we have explored chain growth or addition polymerization, as exemplified by the\nfree-radical mechanism. This particular polymerization route is the most prevalent from a commer-\ncial perspective, and is broadly applicable to a wide range of monomers, especially those containing\ncarbon\u2014carbon double bonds. The main points of the discussion may be summarized as follows:\n"]], ["block_3", ["I\u2018J.C. Bevington, J.I-l. Bradbury, and GM. Burnett, J. Polym. Sal, 12, 469 (1954).\n"]], ["block_4", ["3.9\nChapter Summary\n"]], ["block_5", ["Problems\n"]], ["block_6", ["110\nChain-Growth Polymerization\n"]], ["block_7", ["1.\nThe efficiency of AIBN in initiating polymerization at 60\u00b0C was determinedr by the following\nstrategy. They measured Rp and I7 and calculated R1 RP/17. The constant kd was measured\ndirectly in the system, and from this quantity and the measured ratio RP/ 17 the fractionf could\nbe determined. The following results were obtained for different concentrations of initiator:\n"]], ["block_8", [{"image_1": "125_1.png", "coords": [40, 523, 225, 633], "fig_type": "figure"}]], ["block_9", ["0.0556\n0.377\n0.250\n1.57\n0.250\n1.72\n1.00\n6.77\n1.50\n10.9\n2.50\n17.1\n"]], ["block_10", ["Using kd 0.0388 h\u2018l, evaluate ffl\u2018om these data.\n"]], ["block_11", ["[I] (g L\u201c)\nrep/:7 x 108 (mol L\u201c s\u201c)\n"]]], "page_126": [["block_0", [{"image_0": "126_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["2,\nAIBN was synthesized using 14C\u2014labeled reagents and the tagged compound was used to\ninitiate \nto methyl methacrylate\nand\nstyrene.\nSamples of initiator\nand\npolymers containing initiator fragments were burned to C02. The radioactivity of uniform\n(in sample size and treatment) C02 samples was measured in counts per minute (cpm) by a\nsuitable Geiger counter. A formula for poly(methy1 methacrylate) with its initiator\nfragments is (C5H802),,(C4H6N)m, where n is the degree of polymerization for the polymer\nand m is either or 2, depending on the mode of termination. The specific activity measured\nin the C02, resulting combustion of the polymer relative to that produced by the\ninitiator is\n"]], ["block_2", ["1to. Bevington, H.w. Melville, and RP. Taylor, J. Polym. 5a., 12, 449 (1954).\n"]], ["block_3", ["3.\nIn the research described in Example 3.4, the authors measured the following rates of\npolymerization:\n"]], ["block_4", ["problems\n111\n"]], ["block_5", [{"image_1": "126_1.png", "coords": [40, 466, 215, 569], "fig_type": "figure"}]], ["block_6", [{"image_2": "126_2.png", "coords": [51, 293, 317, 439], "fig_type": "figure"}]], ["block_7", ["Methyl methacrylate\nStyrene\nMn\nCounts per minute\nMn\nCounts per minute\n"]], ["block_8", ["Activity of C in polymer _\n4m\nN \nActivity of C in initiator \n5n + 4m\n\u2014\n5n.\n"]], ["block_9", ["From the ratio of activities and measured values of n, the average number of initiator\nfragments per polymer can be determined.\nCarry out a similar argument for the ratio of activities for polystyrene and evaluate the\naverage number of initiator fragments per molecule for each polymer from the following data.Jr\nFor both sets of data, the radioactivity from the labeled initiator gives 96,500 cpm when\nconverted to C02.\n"]], ["block_10", ["Run number\n1\u2018?p X 104 (H101 1:1 3\u20141)\n"]], ["block_11", ["They also reported a kp value of 1.2 x 104 L mol_l s\ufb02l, but the concentrations of monomer in\neach run were not given. Use these values of RP and kp and the values of 17 and k, given in\nExample 3.4 to evaluate [M] for each run. As a double check, evaluate [M] from these values\nof RI, (and kp) and the values of R, and k, given in the example.\n"]], ["block_12", ["444,000\n20.6\n383,000\n25.5\n312,000\n30.1\n117,000\n86.5\n298,000\n29.0\n114,000\n89.5\n147,000\n60.5\n104,000\n96.4\n124,000\n76.5\n101,000\n113.5\n91,300\n103.4\n89,400\n104.6\n"]], ["block_13", ["13\n7.59\n"]], ["block_14", ["12\n5.48\n"]], ["block_15", ["5\n3.40\n"]], ["block_16", ["6\n2.24\n"]], ["block_17", ["8\n6.50\n"]]], "page_127": [["block_0", [{"image_0": "127_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "127_1.png", "coords": [33, 88, 334, 243], "fig_type": "figure"}]], ["block_2", ["1\u2018 M. Amett, J. Am. Chem. 306., 74, 2027 (1952).\n1MS. Matheson, E.E. Auer, E.B. Bevilacqua, and J.E. Hart, J. Am. Chem. 306., 73, 1700 (1951).\n\u201c\u2018 M. May Jr. and WE. Smith, J. Phys. Chem, 72, 216 (1968).\n"]], ["block_3", ["Propose an explanation for the variation observed.\n6.\nDerive Equation 3.7.27 and Equation 3.7.28.\n7.\nThe equations derived in Section 3.7 are based on the assumption that termination occurs\nexclusively by either disproportionation or combination. This is usually not the case; some\npr0portion of each is more common. If as equals the fraction of chains for which termination\noccurs by disproportionation, it can be shown that\n"]], ["block_4", [{"image_2": "127_2.png", "coords": [35, 482, 193, 525], "fig_type": "molecule"}]], ["block_5", ["Use these data to evaluate the cluster of constants (fkd/kt)\u201d2kp at this temperature. Evaluate\nftp/kt\u201d2 using Arnett\u2019s finding thatf2 1.0 and assuming the kd value determined in Example 3.1\nfor AIBN at 77\u00b0C in xylene also applies in benzene.\n5.\nThe lifetime of polystyrene radicals at 50\u00b0C was measuredI as a function of the extent of\nconversion to polymer. The following results were obtained:\n"]], ["block_6", ["4.\nArnettJr initiated the polymerization of methyl methacrylate in benzene at 77\u00b0C with AIBN and\nmeasured the initial rates of polymerization for the concentrations listed:\n"]], ["block_7", ["1 12\nChain-Growth Polymerization\n"]], ["block_8", [{"image_3": "127_3.png", "coords": [44, 313, 169, 393], "fig_type": "figure"}]], ["block_9", ["1\u2014\n2\n2*\u2014\na+(\na)_\na\nN:\n_\nn\nl\u2014p\nl\u2014p\nl\u2014p\n"]], ["block_10", ["Nw_4\u20143a\u2014ap+2p\n"]], ["block_11", ["From measurements of NJ\u201c and NW/Ml it is possible in principle to evaluate a and p. May\nand Smith)k have done this for a number of polystyrene samples. A selection of their data for\nwhich this approach seems feasible is presented below. Since p is very close to unity, it is\n"]], ["block_12", ["0\n2.29\n32.7\n1.80\n36.3\n9.1\n39.5\n13.1\n43.8\n18.8\n"]], ["block_13", ["9.04\n2.35\n1 1.61\n8.63\n2.06\n10.20\n7.19\n2.55\n9.92\n6.13\n2.28\n7.75\n4.96\n3.13\n7.31\n4.75\n1.92\n5.62\n4.22\n2.30\n5.20\n4.17\n5.81\n7.81\n3.26\n2.45\n4.29\n2.07\n2.1 l\n2.49\n"]], ["block_14", ["and\n"]], ["block_15", ["N11\n(2 a?\n"]], ["block_16", ["Percent conversion\n1\u201c- (s)\n"]], ["block_17", ["[M] (mol L\u201c)\n[I]0 x 104 (mol L\u201c)\nRP x 103 (mol L-1 min\u2014l)\n"]]], "page_128": [["block_0", [{"image_0": "128_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1S.R. Palit and S.K. Das, Proc. Roy. Soc. London, 226A, 82 (1954).\n"]], ["block_2", ["Problems\n1 1 3\n"]], ["block_3", ["10.\n"]], ["block_4", ["11.\n"]], ["block_5", [{"image_1": "128_1.png", "coords": [46, 226, 246, 335], "fig_type": "figure"}]], ["block_6", [". In the research described in Problem 7, the authors determined the following distribution of\nmolecular weights by a chromatographic procedure (w,- is the weight fraction of i-mer):\n"]], ["block_7", [". Derive the two equations given in the previous problem. It may be helpful to recognize that\nfor any distribution taken as a whole, w. = ixi/Nn.\n"]], ["block_8", [{"image_2": "128_2.png", "coords": [53, 84, 120, 164], "fig_type": "figure"}]], ["block_9", ["Na\nNw/Nn\n"]], ["block_10", ["with a 0.65 and p 0.99754. Calculate some representative points for this function and\nplot the theoretical and experimental points on the same graph. From the expression given\nextract the weight fraction i-mer resulting from termination by combination.\nIn fact, the expression in the previous problem is slightly incorrect. Derive the correct\nexpression, and see if the implied values of a and p are signi\ufb01cantly different. The solution\nto Problem 8 provides part of the answer.\nPalit and Das\u2019r measured 17\u201c at 60\u00b0C for different values of the ratio [SX]/[M] and evaluated\nCsx and 17 for vinyl acetate undergoing chain transfer with various solvents. Some of their\nmeasured and derived results are tabulated below (the same concentrations of AIBN and\nmonomer were used in each run). Assuming that no other transfer reactions occur, calculate\nthe values missing from the table. Criticize or defend the following proposition: The 17 values\nobtained from the limit [SX]/[M] \u2014> 0 show that the AIBN initiates polymerization identi-\ncally in all solvents.\n"]], ["block_11", ["100\n3.25\n800\n6.88\n200\n5.50\n900\n6.10\n300\n6.80\n1200\n4.20\n400\n7.45\n1500\n2.90\n500\n7.91\n2000\n1.20\n600\n7.82\n2500\n0.50\n700\n7.18\n3000\n0.20\n"]], ["block_12", ["t-Butyl alcohol\n6580\n3709\n\u2014\u2014\n0.46\nMethyl isobutyl ketone\n6670\n510\n0.492\n\u2014\u2014\nDiethyl ketone\n6670\n\u2014\n0.583\n114.4\nChloroform\n\u2014\n93\n0.772\n125.2\n"]], ["block_13", ["i\nw. x 104\ns\nw; x 104\n"]], ["block_14", ["adequate to assume this value and evaluate a from NW/Nn and then use the value of at so\nobtained to evaluate a better value of p from N\u201c.\n"]], ["block_15", ["They asserted that the points are described by the expression\n"]], ["block_16", ["Solvent\n:7\n17..\n[SX]/[M]\nCSX x 104\n"]], ["block_17", ["w,- = ai(1 p)2pi_1 + 0.5(1 a)i(i 1)(1 p)3p*'*2\n"]], ["block_18", ["1 129\n1.60\n924\n1.67\n674\n1.73\n609\n1 .74\n"]]], "page_129": [["block_0", [{"image_0": "129_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "129_1.png", "coords": [33, 119, 319, 211], "fig_type": "figure"}]], ["block_2", ["1.\nBrandrup, J. and Immergut, E.H., Eds., Polymer Handbook, 3rd ed., Wiley, New York, 1989.\n2.\nTrommsdorff, E., Kohle, H., and Lagally, P., Makromol. Chem, 1, 169 (1948).\n"]], ["block_3", [{"image_2": "129_2.png", "coords": [34, 78, 376, 233], "fig_type": "figure"}]], ["block_4", ["*R.A. Gregg and ER. Mayo, J. Am. Chem. Soc., 70, 2373 (1943).\n"]], ["block_5", ["1 14\nChain-Growth Polymerization\n"]], ["block_6", ["References\n"]], ["block_7", ["12.\nGregg and Mayor studied the chain transfer between styrene and carbon tetrachloride at 60\u00b0C\nand 100\u00b0C. A sample of their data is given below for each of the temperatures.\n"]], ["block_8", ["Evaluate the chain transfer constant (assuming that no other transfer reactions occur) at each\ntemperature. By means of an Arrhenius analysis, estimate Et\"; Eff for this reaction. Are the\nvalues of T2 in the limit of no transfer in the order expected for thermal polymerization?\nExplain.\n13.\nMany olefins can be readily polymerized by a free-radical route. On the other hand,\nisobutylene is usually polymerized by a cationic mechanism. Explain.\n14.\nDraw the mechanisms for the following processes in the radical polymerization of styrene in\ntoluene: (a) initiation by cumyl peroxide; (b) propagation; (c) termination by disproportiona-\ntion; and (d) transfer to solvent.\n15.\nShow the mechanisms of addition of a butadiene monomer to a poly(butadienyl) radical, to\ngive each of the three possible geometric isomers.\n16.\nConsider the polymerization of styrene in toluene at 60\u00b0C initiated by di-t\u2014butylperoxide for\na solution containing 0.04 mol of initiator and 2 mol of monomer per liter. The initial rates of\ninitiation, Ri, and propagation, RP, are found to be 1.6 x 10'10 M s\u20181 and 6.4 x 10\u20187 M 3\u20181,\nrespectively, at 60\u00b0C.\n"]], ["block_9", ["0.00614\n16.1\n0.00582\n36.3\n0.0267\n35.9\n0.0222\n68.4\n0.0393\n49.8\n0.0416\n109\n0.0704\n74.8\n0.0496\n124\n0.1000\n106\n0.0892\n217\n0.1643\n156\n0.2595\n242\n0.3045\n289\n"]], ["block_10", ["((1)\nThe molecular weight of this polymer is too high. The desired molecular weight of this\npolymer is 40,000 g mol\u201d. How much CCl4 (in g L\u2018I) should be added to the reaction\nmedium to attain the desired molecular weight? CT of CC14 is 9 x 10\u20143.\n(e)\nUnder the conditions stated, the polymerization is too slow. What is the initial rate of\npolymerization if the temperature is raised to 100\u00b0C?\n(f)\nCalculate the conversion attained after the reaction has gone for 5 h at 100\u00b0C. Assume\nvolume expansion does not change concentration significantly and that the initiator\nconcentration is constant throughout the entire reaction.\n"]], ["block_11", ["(a)\nCalculate fkd and kp/ktm.\n(b)\nAssuming no chain transfer, calculate the initial kinetic chain length.\n(c)\nAssuming only disproportionation and under the conditions stated, the transfer constant\nof styrene, CM, is 0.85 x 104. How much does this transfer affect the molecular weight\nof the polymer?\n"]], ["block_12", ["[CC14]/[Styrene]\n17;] x 105\n[CCI4]/[Styrene]\n17;! X 105\n"]], ["block_13", ["At 60\u00b0C\nAt 100\u00b0C\n"]]], "page_130": [["block_0", [{"image_0": "130_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["North, \n"]], ["block_2", ["Odian, \nRempp, P. and Merrill, E.W., Polymer Synthesis, 2nd ed., Hiithig & Wepf, Basel, 1991.\n"]], ["block_3", ["Allcock, \n"]], ["block_4", ["Further \n115\n"]], ["block_5", ["Further \n"]], ["block_6", ["NJ, 1990.\n"]]], "page_131": [["block_0", [{"image_0": "131_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In the preceding chapters we have examined the two main classes of polymerization, namely step\u2014\ngrowth and chain\u2014growth polymerizations, with the latter exemplified by the free-radical mech-\nanism. These are the workhorses of the polymer industry, permitting rapid and facile production of\nlarge quantities of useful materials. One common feature that emerged from the discussion of these\nmechanisms is the statistical nature of the polymerization process, which led directly to rather broad\ndistributions of molecular weight. In particular, even in the simplest case (assuming the principle of\nequal reactivity, no transfer steps or side reactions, etc.), the product polymers of either a poly-\ncondensation or of a free\u2014radical polymerization with termination by disproportionation would\nfollow the most probable distribution, which has a polydispersity index (MW/Mn) approaching 2.\nIn commercial practice, the inevitable violation of most of the simplifying assumptions leads to even\nbroader distributions, with polydispersity indices often falling between 2 and 10. In many cases the\npolymers have further degrees of heterogeneity, such as distributions of composition (e.g., copoly-\nmers), branching, tacticity, or microstructure (e.g., cis 1,4\u2014, trans 1,4\u2014, and 1,2-configurations in\npolybutadiene).\nThis state of affairs is rather unsatisfying, especially from the chemist\u2019s point of view. Chemists\nare used to the idea that every molecule of, say, ascorbic acid (vitamin C) is the same as every\nother one. Now we are confronted with the fact that a tank car full of the material called\npolybutadiene is unlikely to contain any two molecules with exactly the same chemical structure\n(recall Example 1.4). As polymers have found such widespread applications, we have obviously\nlearned to live with this situation. However, if we could exert more control over the distribution of\nproducts, perhaps many more applications would be realized. In this chapter we describe several\napproaches designed to exert more control over the products of a polymerization. The major one is\ntermed living polymerization, and can lead to much narrower molecular weight distributions.\nFurthermore, in addition to molecular weight control, living polymerization also enables the\nlarge-scale production of block copolymers, branched polymers of controlled architecture, and\nend-functionalized polymers.\nA comparison between synthetic and biological macromolecules may be helpful at this stage. If\ncondensation and free-radical polymerization represent the nadir of structural control, proteins and\nDNA represent the zenith. Proteins are \u201ccopolymers\u201d that draw on 20 different amino acid\nmonomers, yet each particular protein is synthesized within a cell with the identical degree of\npolymerization, composition, sequence, and stereochemistry. Similarly, DNAs with degrees\nof polymerization far in excess of those realized in commercial polymers can be faithfully\nreplicated, with precise sequences of the four monomer units. One long-standing goal of polymer\nchemistry is to imitate nature\u2019s ability to exert complete control over polymerization. There are\ntwo ways to approach this. One is to begin with nature, and try to adapt its machinery to our\npurpose. This is exemplified by \u201ctraining\u201d cells into growing polymers that we want, for example,\nvia recombinant DNA technology. The other approach, and the one described in this chapter, is to\nstart with the polymerizations we already have, and try to improve them. Both approaches have\nmerit, and we select the latter because it is currently much more established, and plays a central\n"]], ["block_2", ["4.1\nIntroduction\n"]], ["block_3", ["Controlled Polymerization\n"]], ["block_4", ["4\n"]], ["block_5", ["117\n"]]], "page_132": [["block_0", [{"image_0": "132_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["be done in a chain-growth mechanism, through the concentration of initiators. The number of\ninitiators will be equal to the number of polymers, assuming 100% initiation efficiency and\nassuming no transfer reactions that lead to new polymers. The second requirement is to distribute\nthe total number of monomers as uniformly as possible among the fixed number of growing chains.\nIf the polymerization then proceeds to completion, we could predict Nn precisely: it would simply\nbe the ratio of the number of monomers to the number of initiators. To allow the reaction to\nproceed to completion, we would need to prevent termination steps, or at least defer them until we\nwere ready. Now, suppose further that the reaction proceeds statistically, meaning that any\nmonomer is equally likely to add to any growing chain at any time. If N,1| was reasonably large,\nwe could expect a rather narrow distribution of the number of monomers in each chain, just by\nprobability. (This argument also assumes no transfer reactions, so that growing polymers are not\nterminated prematurely.) As an illustration, imagine placing an array of empty cups out in a steady\nrain; an empty cup is an \u201cinitiator\u201d and a raindrop is a \u201cmonomer.\u201d As time goes on, the raindrops\nare distributed statistically among the cups, but after a lot of drops have fallen, the water level will\nbe pretty much equal among the various cups. If a cup fell over, or a leaf fell and covered its top,\nthat \u201cpolymer\u201d would be \u201cterminated,\u201d and its volume of water would not keep up with the others.\nSimilarly, if you placed a cup outside a few minutes after the others, the delayed initiation would mean\nthat it would never catch up with its neighbors. What we have just described is, in fact, the essence of a\ncontrolled polymerization: start with a fixed number of initiators, choose chemistry and conditions to\neliminate transfer and termination reactions, and let the reaction start at a certain time and then go to\ncompletion. In order to control the local structural details, such as microstructure and stereochemistry,\nwe have to in\ufb02uence the relative rates of various propagation steps. This can be achieved to some\nextent by manipulating the conditions at the active site at the growing end of the chain.\nThe remainder of this chapter is organized as follows. First we demonstrate how the kinetics of\nan ideal living polymerization leads to a narrow, Poisson distribution of chain lengths. Then, we\nconsider chain\u2014growth polymerization via an anionic propagating center; this has historically\nbeen the most commonly used controlled polymerization mechanism, and it can be conducted in\nsuch a way as to approach the ideal case very closely. In Section 4.4 we explore how the anionic\nmechanism can be extended to the preparation of block copolymers, end-functional polymers, and\nregular branched polymers of various architectures. We then turn our attention to other mechan-\nisms that are capable of controlled polymerization, including cationic (Section 4.5), ring-opening\n(Section 4.8), and, especially, controlled radical polymerizations (Section 4.6). The concluding\nsections also address the concept of equilibrium polymerization, and a special class of controlled\npolymers called dendrimers.\n"]], ["block_2", ["4.2\nPoisson Distribution for an Ideal Living Polymerization\n"]], ["block_3", ["In this section we lay out the kinetic scheme that describes a living polymerization, and thereby\nderive the resulting distribution of chain lengths. This scenario is most closely approached in the\nanionic case, but because it is not limited to anionic polymerizations, we will designate an active\n"]], ["block_4", ["role in much of polymer research. It is worth noting that nature also makes use of many other\nmacromolecular materials that are not so well\u2014controlled as proteins and DNA; examples include\npolysaccharides such as cellulose, chitin, and starch. So in nature, as with commercial polymers,\nuseful properties can still result from materials that are very heterogeneous at the molecular level.\nThe lack of control over molecular weight in polymerization arises directly from the random\ncharacter of each step in the reaction. In a polycondensation, any molecule can react with any other\nat any time; the number of molecules is steadily decreasing, but the mole fraction of monomer is\nalways larger than the mole fraction of any other species. In a free-radical polymerization, chains\nmay be initiated at any time. Growing chains may also add monomer, or undergo a transfer or\ntermination reaction at any time. The first requirement in controlling molecular weight is to fix the\ntotal number of polymers. This cannot be done in an unconstrained step-growth process, but it can\n"]], ["block_5", ["118\nControlled Polymerization\n"]]], "page_133": [["block_0", [{"image_0": "133_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["When the reaction has gone to completion, [M] will be 0, and the kinetic chain length will be\nthe number average degree of polymerization of the resulting polymer. It will also be helpful in the\nfollowing development to differentiate Equation 4.2.5 with respect to time, and then incorporate\nEquation 4.2.3:\n 9?- \n1 d[M]\n"]], ["block_2", ["This is a linear, first-order differential equation for [M], which has the solution\n"]], ["block_3", ["Therefore the concentration of monomer decreases exponentially to zero as time progresses. (Note\nthat we are also assuming that propagation is irreversible, that is, there is no \u201cback arrow\u201d in\nEquation 4.2.2. The possibility of depolymerization reaction steps will be taken up in Section 4.7.)\nAt this stage it is very helpful to introduce a kinetic chain length, a, analogous to the one we\ndefined in Equation 3.5.10, as the ratio of the number of monomers incorporated into polymers to\nthe number of polymers. The former is given by [M]0 [M], and the latter by [I]0, so we write\n"]], ["block_4", ["Note that in using a single propagation rate constant, kp, we are once again invoking the principle\nof equal reactivity.\nWe will now assume that initiation is effectively instantaneous relative to propagation (ki >> kp),\nso that at time t= 0, [Pf] [I]0, and we will not worry about Equation 4.2.1 any further. Note that\nthis criterion is not necessary to have a living polymerization, but it is necessary to achieve a\nnarrow distribution of molecular weights. The concentration of unreacted monomer, [M], will\ndecrease in time as propagation takes over. The overall rate of polymerization, RP, is the sum of the\nrates of consumption of monomer by all growing chains Pf. However, we know that, in the absence\nof termination or transfer reactions, the total concentration of P? is always [I]0: we have fixed the\nnumber of polymers. Therefore we can write\n"]], ["block_5", ["The concentration ofmonomer at time t will be denoted [M]. The initial concentrations\nof monomer and initiator are [M]0 and [I]0, respectively. The reaction steps can be represented as\nfollows:\n"]], ["block_6", ["4.2.1\nKinetic Scheme\n"]], ["block_7", ["polymer of degree of polymerization i by P?\u2018, and its concentration by [Pf], where represents the\nreactive end. is defined as a chain-growth processfor which there are no\nirreversible \nabout but \nwill not \n"]], ["block_8", ["Poisson Distribution Polymerization\n119\n"]], ["block_9", ["[M10 [M]\n_ = \u2014\u2014\u2014-\u2014-\u2014\n4.2.5\nV\n[Ho\n(\n)\n"]], ["block_10", ["(1 M\nRp \u2014\u2014[d7\u2014] MM]2 [Pi] kl][ I 10\n(4.2.3)\n"]], ["block_11", ["k\nPropagation:\nPi\u201c + M \u2014\">\n193\u2018\n"]], ["block_12", ["Initiation:\nI+M ii Pi\"\n(42.1)\n"]], ["block_13", ["[M] [Mloe\u2018kpllb\u2018\n(4.2.4)\n"]], ["block_14", ["k\nPf-\u201c+ M \u2014\"i\nP511\n(4.22)\n"]]], "page_134": [["block_0", [{"image_0": "134_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "134_1.png", "coords": [26, 340, 267, 411], "fig_type": "figure"}]], ["block_2", [{"image_2": "134_2.png", "coords": [31, 415, 152, 454], "fig_type": "molecule"}]], ["block_3", [{"image_3": "134_3.png", "coords": [31, 159, 194, 195], "fig_type": "molecule"}]], ["block_4", ["We could insert Equation 4.2.4 into Equation 4.2.7 to replace [M], and thereby obtain an equation\nthat can be solved. However, a simpler approach turns out to be to invoke the chain rule, as\nfollows:\n"]], ["block_5", ["Now we repeat this process for [PER], beginning with the rate law. This is slightly more compli-\ncated, because [Pfrf] grows by the reaction of Pi\" with monomer, but decreases by the reaction of Pi\"\nwith monomer:\n"]], ["block_6", ["In order to obtain the distribution of chain lengths, we need to do a bit more work. We begin by\nwriting an explicit equation for the rate of consumption of [PT]:\n"]], ["block_7", ["and comparing with Equation 4.2.11 we obtain\n"]], ["block_8", ["This equation has the solution\n"]], ["block_9", ["We can go through this sequence of steps once more, considering the concentration of trimer [133\u2018]:\n"]], ["block_10", ["By invoking the chain rule once more\n"]], ["block_11", ["If we now compare Equation 4.2.7 and Equation 4.2.8 we can see that\n"]], ["block_12", ["leading to\n"]], ["block_13", ["and this equation is readily solved:\n"]], ["block_14", ["120\nControlled Polymerization\n"]], ["block_15", [{"image_4": "134_4.png", "coords": [36, 563, 263, 631], "fig_type": "molecule"}]], ["block_16", ["d 13*\n_\nL114. [133\u2018] : [Pg\u2018] : \ufb02uoe-V\n(4.2.16)\ndii\n"]], ["block_17", ["d 13*\n_\n[1;] + [133\u2018] [133\u2018] [Hoe\u2014V\n(4.2.13)\n"]], ["block_18", ["d[133\u2018]\nd[133\u2018] an\n"]], ["block_19", ["d[131*\u2018] : d[P3\u2018] (1'17 __ d[P1\"]\ndt\ndv E-\nd}? \n(4.2.8)\n"]], ["block_20", ["d P*\n*\n_ _%__\ufb01ll : [131]\n(4.2.9)\n"]], ["block_21", ["[133\u2018] : E[I]0e\u201d'_\u2019\n(4.2.14)\n"]], ["block_22", ["(1 [P3]\ndt \n"]], ["block_23", ["d P*\n{.31 41341114 4 1341 [M]\n"]], ["block_24", ["[133\u2018] [13\u20181\"]04rE [nae\u20145\n(4.2.10)\n"]], ["block_25", ["dr \nd?\nd;\n(4.2.12)\n"]], ["block_26", ["d [PT]\ndt\n2 kp [133\u2018] [M]\n(4.2.7)\n"]], ["block_27", ["d?\n:\np[M]([P\u00a7] [Path 51112:] \n[13311)\n(42.15)\n"]], ["block_28", ["d?\n=41M1<1P\ufb02\u20141P41>=~a;<1311311341)\n(4.2.11)\n"]]], "page_135": [["block_0", [{"image_0": "135_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "135_1.png", "coords": [20, 384, 266, 526], "fig_type": "figure"}]], ["block_2", ["a small number of boxes or polymers). Once the polymerization reaction has gone to\ncompletion, and the polymers by introduction of some appropriate reagent, the resulting\nmolecular weight distribution should obey Equation 4.2.19, with i equal to [M]0/[I]0.\nThe following example illustrates some aspects of the kinetics of a living polymerization.\n"]], ["block_3", ["From this the mole fraction \npolymers, x,, by dividing 4.2.18 by the total number of polymers, [I]0:\nvine\u2014a\nx, \n(i 1)!\n(4.2.19)\n"]], ["block_4", ["The following data were reported for the living anionic polymerization of styrene:f The initial\nmonomer concentration was 0.29 mol LTI, and the initiator concentration was 0.00048 mol L\u2018l.\nThe reactor was sampled at the indicated times, and the resulting polymer was terminated and\nanalyzed for molecular weight and polydispersity. Use these data and Equation 4.2.4 and Equation\n4.2.5 to answer the following questions: Does conversion of monomer to polymer follow the\nexpected time dependence? What is the propagation rate constant under these conditions?\n"]], ["block_5", ["This particular is called Poisson distribution. \nobtained itfrom consideringfact it will describe \nwhenever larger objects or are distributed randomly \n"]], ["block_6", ["This pattern population of i-mer is\n"]], ["block_7", ["Using Equation 4.2.4 we see how p should evolve in time:\n"]], ["block_8", ["Therefore a plot of 1n(1 \u2014p) versus t should give a straight line with slope equal to \u2014kp[I]0. The data\nprovided do not include [M] explicitly, but we can infer [M] and p from Mn. From Equation 4.2.5,\n"]], ["block_9", ["which has \n"]], ["block_10", ["t (s)\nMn (g/mol)\nNn\nPDI\nl\u2014p\n"]], ["block_11", ["We can equate the conversion of monomer to polymer with the familiar extent of reaction, p, as in\nChapter 2 and Chapter 3:\n:[Mlo\u2014[M] ____1_ [M1\n[M10\n[M10\n"]], ["block_12", ["Poisson \n121\n"]], ["block_13", ["Solution\n"]], ["block_14", ["TW. Lee, H. Lee, 1. Cha, T. Chang, K]. Hanley, and T.P. Lodge, Macromolecules, 33, 5111 (2000).\n"]], ["block_15", ["Example 4.1\n"]], ["block_16", [{"image_2": "135_2.png", "coords": [36, 161, 111, 206], "fig_type": "molecule"}]], ["block_17", ["238\n3,770\n36.3\n1.06\n0.940\n888\n20,600\n198\n1.02\n0.672\n1,626\n33,700\n324\n1.02\n0.463\n2,296\n43,000\n413\n1.01\n0.316\n3,098\n49,800\n479\n1.008\n0.207\n4,220\n54,900\n528\n1.006\n0.127\n14,345\n61,700\n593\n1.005\n0.018\n"]], ["block_18", ["p \n1 \n"]], ["block_19", ["1\n_\n[Pi] gazilloe\u2018\u201d\n(4.2.17)\n"]], ["block_20", ["1\n.\n_\n:1:\n_\n-\u20141-1\n\u2014V\n[P,-] \n(F1)!\n:2\n[Hoe\n(4.2.18)\n"]]], "page_136": [["block_0", [{"image_0": "136_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The suggested plot is shown below and the resulting slope from linear regression implies that\nkl, a:\n1 mol L\u20181 5\u20181. (This is actually a rather low value, and in fact only an apparent value, due to\n"]], ["block_2", ["4.2.2\nBreadth of the Poisson Distribution\n"]], ["block_3", ["Figure 4.1 illustrates the Poisson distribution for values of ? equal to 100, 500, and 1000. For\npolystyrene with M0 104, these would correspond to polymers with number average molecular\nweights of about 104, 5 x 104, and 105, respectively, which are moderate. The width of the distribu-\ntions, although narrow, increases with r, but as we shall see in a moment, the relative width (i.e., the\nwidth divided by 17) decreases steadily. It should be clear that these distributions are very narrow\ncompared to the step-growth or free\u2014radical polymerizations shown in Figure 2.5 and Figure 3.5,\nrespectively. To underscore this, Figure 4.2 compares the theoretical distributions for free-radical\npolymerization with termination by combination (Equation 3.7.26) and for living polymerization, both\nwith E 100. The difference is dramatic, and is made even more so when we recall that termination by\ncombination leads to a relatively narrow distribution with Mw/Mn approaching 1.5 rather than 2.\nFor the Poisson distribution, the polydispersity index, Mw/Mn, in fact approaches unity as 3\nincreases indefinitely. The explicit relation for the Poisson distribution is\n"]], ["block_4", ["the table was obtained as\n"]], ["block_5", ["a phenomenon to be described in Section 4.3 [also see Problem 3]. Also note that the last data point\nhas been omitted from the fit, as it corresponds to essentially complete conversion, and thus is\nindependent of t once the reaction is finished.)\n"]], ["block_6", ["the kinetic chain length is equal to p[I]0/[M]O, and it is also equal to NH \n"]], ["block_7", ["where the approximation applies for large ?. For 17: 1000 Equation 4.2.20 indicates that the\npolydispersity index will be 1.001, which is a far cry from 2.\n"]], ["block_8", ["122\nControlled Polymerization\n"]], ["block_9", [{"image_1": "136_1.png", "coords": [38, 595, 220, 628], "fig_type": "molecule"}]], ["block_10", [{"image_2": "136_2.png", "coords": [44, 190, 284, 416], "fig_type": "figure"}]], ["block_11", ["MW\nNW\n1\nE\n__:_____\n____~1\n4.2.20\nMn\nN.\n+(1+r)2\n(\n)\nElli-d\n"]], ["block_12", ["1_ #1_ N \np \n[M]0 \n(0.29 molL\u201c)\n\u201c\n"]], ["block_13", ["__2-5_I_II_I_III_III_I_IIIIIIII|\nIll\u2014\n0\n1 000\n2000\n3000\n4000\n5000\n"]], ["block_14", ["\u20140.5 \nSlope = \u20140.00051 3\u20181\n-\n"]], ["block_15", ["O\nI\nI\nr'\nI\nI\nI\nI\nI_l\u2014I\nI\nl\u2014l\u2014W\nI\u2014l\u2014I\nITI\nI\nl\u2014\n"]], ["block_16", ["\u2018l\u2014I'\n"]], ["block_17", ["I\n"]], ["block_18", ["lI\ufb01\u2014II\ufb01\u2014Ilj\u2014I\n"]], ["block_19", ["{,5\n"]], ["block_20", ["lrr\n"]]], "page_137": [["block_0", [{"image_0": "137_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "137_1.png", "coords": [19, 400, 298, 655], "fig_type": "figure"}]], ["block_2", [{"image_2": "137_2.png", "coords": [21, 25, 297, 303], "fig_type": "figure"}]], ["block_3", [{"image_3": "137_3.png", "coords": [25, 364, 173, 410], "fig_type": "molecule"}]], ["block_4", ["Figure 4.2\nComparison of Poisson distribution and distribution for free-radical polymerization with\ntennination by combination.\n"]], ["block_5", ["The derivation of Equation 4.2.20 is not too complicated, but it involves a couple of useful\nnicks, as we will show now. From Equation 1.7.2, we recall the definition of N\u201c, and insert\nEquation 4.2.19 to obtain\n"]], ["block_6", ["Poisson \n123\n"]], ["block_7", ["Figure 4.1\nMole fraction of i\u2014mer for the Poisson distribution with the indicated kinetic chain lengths.\n"]], ["block_8", ["s:\n-\n'.\n\u2018\n_\n-.\n-l\n0.02 \n-.\n_\n'\ng\n9.\nV = 500\n-\n"]], ["block_9", ["s;- 0.02 i\n-\nJ_\ni\n.\n\u00b0\n_;\n"]], ["block_10", ["0'0\ni\u2014i\u2014 \nNn:;ixi=:L\n(1\u2014 1)!\n(4.2.21)\n"]], ["block_11", ["0'04 .\u2014\n5 17: 100\ni\n"]], ["block_12", ["g\n' \nIdeal living polymerization, v = 100\n:\n0.03 i:\n. \n1\n"]], ["block_13", ["0.01 E\n:\nZ\n.\n_\n'\n:\n'\n_\n.\nIdeal free-radical polymerization,\n:\n"]], ["block_14", ["0,04\nL'I_l_\nI\nI\nI T I\nI\nI ; I\nl\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI \"l_l\nI\nI\nIT I\nI\nI\nl\n"]], ["block_15", ["0.01 :2\nf\n"]], ["block_16", ["0.03\n\u2014\n-'\n\u2014\n_\n_.\ni\n"]], ["block_17", ["0.05\n"]], ["block_18", ["0\n:\nl\n1\nI 1\u20191\nl _I_Ll \\4\ni\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nl\nI\nI\nI\nI\nI\nl\nI\nI\nI\nI\n"]], ["block_19", ["0\nI}! i\nL\nl_L\n] l_l\nLL_I\nI\nI\nI\nI\n\u2019\n0\n200\n400\n600\n800\n1 000\n1 200\n"]], ["block_20", [{"image_4": "137_4.png", "coords": [62, 132, 293, 284], "fig_type": "figure"}]], ["block_21", ["0\n100\n200\n300\n400\n"]], ["block_22", ["E\n,\n.\ntermination by recombination,\n-\n"]], ["block_23", ["__\n.\n'_\ni7\n= 100\n.I\n"]], ["block_24", ["t\n0 \n1\n"]], ["block_25", ["i\n- j\ni\n"]], ["block_26", ["k\u2014I\u2014\n"]], ["block_27", ["_\nd:\n_{\n"]], ["block_28", ["\u2014\n. _\n17: 1000\n\u2014\n"]], ["block_29", ["E\n'5\n5\u2018\n"]], ["block_30", ["l\nI \u2018l\u2014I\u2014 I _I\u2014|'\nI \u2014I\u2014I\u2014\ni\nI\nI\nI\ni\nl\nI\u2014I\u2014\nI\nI\nI\nI\n"]], ["block_31", ["-\n"]]], "page_138": [["block_0", [{"image_0": "138_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "138_1.png", "coords": [26, 81, 161, 122], "fig_type": "molecule"}]], ["block_2", ["This relationship establishes that Nn = 1 + 3. (You may be wondering where the \u201c1\u201d came from.\nA glance at Equation 4.2.5 reveals the answer: before the reaction begins, when [M] [M]0, then\n3 0 when the degree of polymerization is actually 1. Of course, for any reasonable value of NH,\nthe difference between Nn and M, + 1 is inconsequential.)\nThe development to obtain an expression for NW follows a similar approach, beginning with the\ndefinition from Equation 1.7.4:\n"]], ["block_3", ["We already know that the denominator on the right-hand side of Equation 4.2.25 is equal to 1 + a,\nso we just need to sort out the numerator\n"]], ["block_4", ["This differentiation is straightforward, recalling the rule for differentiating the product of two\nfunctions, and that d(e\")/dx\u2014\n- ex:\n"]], ["block_5", ["which leaves us with some more derivatives to take:\n"]], ["block_6", ["To progress further with this, it is helpful to recall the infinite series expansion of ex (see the\nAppendix if this is unfamiliar):\n"]], ["block_7", ["We will use this expansion to get rid of the factorials. Returning to Equation 4.2.21, we perform a\nseries of manipulations, recognizing that e\u2018\u201d does not depend on i and can be factored out of the\nsum, and that 1'74 can be written as d(v')/dv:\n"]], ["block_8", ["124\nControlled Polymerization\n"]], ["block_9", [{"image_2": "138_2.png", "coords": [35, 298, 226, 329], "fig_type": "molecule"}]], ["block_10", [{"image_3": "138_3.png", "coords": [36, 163, 216, 202], "fig_type": "molecule"}]], ["block_11", [{"image_4": "138_4.png", "coords": [41, 400, 147, 458], "fig_type": "molecule"}]], ["block_12", [{"image_5": "138_5.png", "coords": [45, 486, 246, 517], "fig_type": "molecule"}]], ["block_13", ["00\n00\n2 \ufb01x,\nN, = Ziw, = '3\n(4.2.25)\ni=1\n2 ix,-\ni=1\n"]], ["block_14", ["..;7\nd\n_ d\n__\n3\n_\n_;\nd\n_\n3\n_\n'17\ne\nE(va;{ve })\n\u2014 e\nd? (v{e +ve })\n"]], ["block_15", ["00\nx\u2018\n00\n=\nex; g)\u201c\n(i 1)!\n(4.222)\n"]], ["block_16", ["ivl\u2018\n1 6\u2014D\n_\n00\n1311\nN\n::\n\u2014\u2014\u2014-\u2014\u2014\u2014-\u2014\u2014\u2014\u2014-:\nFV\nn\nIElla\u20141)!\ne\n;(_i\u2014\u2014l)!\n"]], ["block_17", ["oo\n00\n-i-\u2014 1\n00\n\u2014i\n2\n2\u201d\ne\n_\n\u2014?\n.2\nV\nx\u2019= 2\u2019 (5\u20141)!\u2018e\n1230\u20141)!\ni=1\ni=1\n"]], ["block_18", ["\u2019Vd\u2014\ufb01\n\u2014{ve }\u2014e \u201d{e + ve\"\u2014\u2014} 1 +1)\n(4.2.24)\n"]], ["block_19", ["_;d\n00\n_\n31\u20141\n\u201417\nd\n__\n00\n\ufb01h-l\n--l_/ d\n_ 3\n_ e\nE;V(i\u20141)!_e\n3-;{1/1);???\n\u2014\u2014-\u2014-\u2014)\u20141!}\u2014e\nd\u2014?{ve}\n(4.2.23)\n"]], ["block_20", ["_ e_,\n0\u00b0\nd\nv\"\n_e_,\u2014,d\n\u00b0\u00b0\ni7\"\n\u201d\ni=1 \ndam (xx\u20141)!\n"]], ["block_21", [{"image_6": "138_6.png", "coords": [74, 505, 318, 585], "fig_type": "molecule"}]], ["block_22", [{"image_7": "138_7.png", "coords": [79, 519, 308, 553], "fig_type": "molecule"}]], ["block_23", ["\u2014 e\u2018F\u2014d\u2014vfl\u2014 i\n1,:-\n\u2014 e\u2014\ufb01i\ufb01fl\u2014 {reB\n\u2018\ndv\ncw\n.\n(i .\u2014 1)!\n_\nd\ufb01 dr\n(4.2.26)-\n"]], ["block_24", ["= e\u20187 {\ufb01eF + e3 + 2%; + 32 e;}\n= 1 + 33 + v?\n(4.2.27)\n"]]], "page_139": [["block_0", [{"image_0": "139_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "139_1.png", "coords": [30, 426, 267, 610], "fig_type": "figure"}]], ["block_2", ["using \n"]], ["block_3", ["The polydispersity \n(Equation consistently \ntion, but \n"]], ["block_4", ["seem to be approaching result; the implications of this observation are considered in\nProblem \nonly recently a polydispersity index \nwould require an accuracy much in the determination of M and Mn, and this is not\nyet possible using standard techniques in Chapter 7, Chapter 8, and\nChapter 9. In one sample obtained \nlaser desorption/ionization (MALDI)spectrometry shown \nwith the Poisson distribution mean; the the \ntal distribution being only slightly broader than the theoretical one.\nWe conclude this section with a summary of the requirements to achieve a narrow molecular\nweight distribution, and thereby draw an important distinction between \u201clivingness\u201d and the\nPoisson distribution. To recall the basic definition, a living polymerization is one that proceeds\nin the absence of transfer and termination reactions. Satisfying these two criteria is not sufficient to\n"]], ["block_5", ["It is \n"]], ["block_6", ["Finally, \n"]], ["block_7", ["Poisson \n125\n"]], ["block_8", ["Figure 4.3\n(a) Experimental polydispersities versus molecular weight for anionically polymerized polysty-\nrenes, from the data in Example 4.1. (b) The distribution obtained by MALDI mass spectrometry for one\nparticular sample. The smooth curves represent the results for the Poisson distribution, Equation 4.2.29 in (a)\nand Equation 4.2.19 in (b).\n"]], ["block_9", ["Mn,g/mol\nM\n(a)\n(b)\n"]], ["block_10", ["1.06 2-0\n-.\n_\n:\n:\n.5 0'6?\n1.05:-\n\u2014j\n<3\n,\n.\n.\n3\n:\n2\nQ-\n'\ng 1.04 ;\n__\nDC: 0.4_\n_\n\\3\nE\ng\n-\nE\n1.03\n_\u2014\n\u2014_\n.\n"]], ["block_11", [{"image_2": "139_2.png", "coords": [44, 400, 411, 578], "fig_type": "figure"}]], ["block_12", ["1 + 3? + 32\nNW _ _1\u2014+\u00a7\u2014\n(4.2.28)\n"]], ["block_13", ["_N_.,,_:1_\u2014I\u20143>Z;I\u2014TT\u00bb2=_(1\u2014I\u2014P\u2014)i-l2\u20143: +L_2\n(4.2.29)\nN\u201c\n(1 +12)\n(1+V)\n(1 +12)\n"]], ["block_14", ["1.07\n:I\n"]], ["block_15", ["r\n3\n'\n1.01 \no\n,\n\u2014j\n-\n"]], ["block_16", ["102:-\n.\nO\n-:\n02-\n_\n"]], ["block_17", ["Q\ng\nz\n-\n1IIIIIIIllllIIIIIIIIIIIIIIIII_I\u2014l\u2014IIII.\nOllgrllrrrllrllll\n.\n"]], ["block_18", ["0\n1x104 2x104 3x104 4x104 5x104 6x104 7x104\n3.5x104 4x104 4.5>-<1o4 5x104 5.5><1o\u2018l sxro4 6.5x104\n"]], ["block_19", ["l\nI\nI\n\u2018I\nl_l_l_l\nI\nI\nI\nl\nI\nl\nI\nI\nI\nI\nI\nl\nI\nI\nI\nII\nI\nl\nI\nI\nI\nl\nl\nl_\n:\n"]], ["block_20", [{"image_3": "139_3.png", "coords": [199, 390, 441, 591], "fig_type": "figure"}]], ["block_21", ["0.3-\n_\n"]], ["block_22", ["1\nI\nI\n-\n"]]], "page_140": [["block_0", [{"image_0": "140_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Anionic polymerization has been the most important mechanism for living polymerization, since\nits first realization in the 1950s [2]. Both modes of ionic polymerization (i.e., anionic and cationic)\nare described by the same vocabulary as the corresponding steps in the free\u2014radical mechanism for\nchain-growth polymerization. However, initiation, propagation, transfer, and termination are quite\ndifferent in ionic polymerization than in the free-radical case and, in fact, different in many ways\nbetween anionic and cationic mechanisms. In particular, termination by recombination is clearly\nnot an option in ionic polymerization, a simple fact that underpins the development of living\npolymerization. In this section we will discuss some of the factors that contribute to a successful\nliving anionic polymerization, and in the following section we will illustrate the extension of these\ntechniques to block copolymers and controlled architecture branched polymers.\nMonomers that are amenable to anionic polymerization include those with double bonds (vinyl,\ndiene, and carbonyl functionality), and heterocyclic rings (see also Table 4.3). In the case of vinyl\nmonomers CH2 CHX, the X group needs to have some electron withdrawing character, in order\nto stabilize the resulting carbanion. Examples include styrenes and substituted styrenes, vinyl\naromatics, vinyl pyridines, alkyl methacrylates and acrylates, and conjugated dienes. The relative\nstabilities of these carbanions can be assessed by considering the pKa of the corresponding\nconjugate acid. For example, the polystyryl carbanion is roughly equivalent to the conjugate\nbase of toluene. The smaller the pKa of the corresponding acid, the more stable the resulting\ncarbanion. The more stable the carbanion, the more reactive the monomer in anionic polymeriza-\ntion. In the case of anionic ring-opening polymerization (ROP), the ring must be amenable to\nnucleophilic attack, as well as present a stable anion. Examples include epoxides, cyclic siloxanes,\nlactones, and carbonates. At the same time, there are many functionalities that will interfere with\nan anionic mechanism, especially those with an acidic proton (e.g., \u2014OH, \u2014NH3, \u2014COOH) or an\nelectrophilic functional group (e.g., 02, \u2014C(O)\u2014, C02). Anionic polymerization of monomers that\ninclude such functionalities can generally only be achieved if the functional group can be\nprotected. As a corollary, the polymerization medium must be rigorously free of protic impurities\nsuch as water, as well as oxygen and carbon dioxide.\nA wide variety of initiating systems have been developed for anionic polymerization. The \ufb01rst\nconsideration is to choose an initiator that has a comparable or slightly higher reactivity than the\nintended carbanion. If the initiator is less reactive, the reaction will not proceed. If, on the other\nhand, it is too reactive, unwanted side reactions may result. As the pKas of the conjugate acids for\nthe many possible monomers span a wide range, so too must the pKas of the conjugate acids of the\ninitiators. Second, the initiator must be soluble in the same solvent as the monomer and resulting\npolymer. Common classes of initiators include radical anions, alkali metals, and especially\nalkyllithium compounds. We will illustrate two particular initiator systems: sodium naphthalenide,\nas an example of a radical anion, for the polymerization of styrene, and sec\u2014butyllithium, as an\nalkyllithium, in the polymerization of isoprene.\n"]], ["block_2", ["4.3\nAnionic Polymerization\n"]], ["block_3", ["guarantee a narrow distribution, however. The additional requirements for approaching the Poisson\ndistribution are:\n"]], ["block_4", ["1.\nAll active chain ends must be equally likely to react with a monomer throughout the\npolymerization. This requires both the principle of equal reactivity, and good mixing of\nreagents at all times.\n2.\nAll active chain ends must be introduced at the same time. In practice, this means that the rate\nof initiation needs to be much more rapid than the rate of propagation, if all the monomer is\nadded to the reaction mixture at the outset.\n3.\nPropagation must be essentially irreversible, that is, the reverse \u201cdepolymerization\u201d reaction\ndoes not occur to a significant extent. There are, in fact, cases where the propagation step is\nreversible, leading to the concept of an equilibrium polymerization, which we will take up in\nSection 4.8.\n"]], ["block_5", ["126\nControlled Polymerization\n"]]], "page_141": [["block_0", [{"image_0": "141_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "141_1.png", "coords": [31, 383, 272, 437], "fig_type": "figure"}]], ["block_2", ["4.\n"]], ["block_3", ["Now we consider the polymerization of isoprene by sec\u2014butyllithium, in benzene at room\ntemperature. In the first step, one monomer is added, but immediately there are many possibilities,\n"]], ["block_4", [{"image_2": "141_2.png", "coords": [35, 558, 336, 601], "fig_type": "molecule"}]], ["block_5", ["Anionic Polymerization\n127\n"]], ["block_6", ["as indicated:\n"]], ["block_7", ["3.\n"]], ["block_8", ["tetrahydrofuran \n"]], ["block_9", ["1.\n"]], ["block_10", [{"image_3": "141_3.png", "coords": [36, 379, 247, 447], "fig_type": "molecule"}]], ["block_11", ["The first living polymer studied in detail was polystyrene initiated with sodium naphthalenide in\n"]], ["block_12", ["Me\nMe\n1 4\nMe\n/\nMe\n/0\\\n"]], ["block_13", ["Me+Li\nMe\n1 2 3 4\nMe\n+Li\n"]], ["block_14", [{"image_4": "141_4.png", "coords": [49, 284, 128, 342], "fig_type": "figure"}]], ["block_15", [{"image_5": "141_5.png", "coords": [50, 274, 241, 351], "fig_type": "figure"}]], ["block_16", [{"image_6": "141_6.png", "coords": [51, 381, 150, 442], "fig_type": "molecule"}]], ["block_17", ["The latter undergoes radical combination to form the dianion, which subsequently initiates the\npolymerization:\n"]], ["block_18", ["In this case, the degree of polymerization is 2? because the initiator is difunctional; further-\nmore, there will be a single tail-to\u2014tail linkage somewhere near the middle of each chain.\nThe propagation step at either end of the chain can be written as follows:\n"]], ["block_19", ["Of course the structure of the radical anion shown is just one of the several possible resonance\nforms.\nThese green radical ions react with styrene to produce the red styryl radical ion:\n"]], ["block_20", ["The precursor to the initiator is prepared by the reaction of sodium metal with naphthalene and\nresults in the formation of a radical ion:\n"]], ["block_21", ["H\nCH2\nH..c.:H2 \ni\u201d +6\n\u201dJ\u201d\n(4.3)\n"]], ["block_22", ["The carbanion attacks the more electropositive and less sterically hindered carbon to regen\u2014\nerate the more stable benzylic carbanion. Thus, the addition is essentially all head\u2014to-tail in\nthis case. Note also that the sodium counterions have not been written explicitly in Reactions\n(4.B)\ufb02(4.D), although of course they are present. As we will see below, the counterion can\nactually play a crucial role in the polymerization itself.\n"]], ["block_23", ["Na+HNa*+[i_\n(4A)\n"]], ["block_24", ["1,44\n_H_\nH\nCH\n"]], ["block_25", [{"image_7": "141_7.png", "coords": [61, 116, 216, 150], "fig_type": "molecule"}]], ["block_26", ["H .. o\n"]], ["block_27", [{"image_8": "141_8.png", "coords": [81, 192, 311, 258], "fig_type": "molecule"}]], ["block_28", [{"image_9": "141_9.png", "coords": [110, 387, 247, 437], "fig_type": "molecule"}]], ["block_29", [{"image_10": "141_10.png", "coords": [110, 279, 222, 339], "fig_type": "molecule"}]], ["block_30", [{"image_11": "141_11.png", "coords": [124, 551, 386, 616], "fig_type": "molecule"}]], ["block_31", [{"image_12": "141_12.png", "coords": [153, 626, 360, 687], "fig_type": "molecule"}]], ["block_32", [{"image_13": "141_13.png", "coords": [157, 619, 374, 686], "fig_type": "figure"}]], ["block_33", ["+LI\n1,2\nMe\n\u2018Me\n3,4\n\u201c\" JAE\nMel/I:\nMe\n\\CHQ\nMe\nMe\n(413)\n"]], ["block_34", [{"image_14": "141_14.png", "coords": [218, 635, 343, 680], "fig_type": "molecule"}]], ["block_35", [{"image_15": "141_15.png", "coords": [271, 549, 378, 605], "fig_type": "molecule"}]], ["block_36", ["(4D)\n"]], ["block_37", ["(4.C)\n"]]], "page_142": [["block_0", [{"image_0": "142_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "142_1.png", "coords": [24, 527, 454, 667], "fig_type": "figure"}]], ["block_2", [{"image_2": "142_2.png", "coords": [33, 306, 432, 447], "fig_type": "figure"}]], ["block_3", ["THF\nLi\n30\n12 combined\n29\n59\nDioxane\nLi\n15\n3\n1 1\n18\n68\nHeptanea\nLi\n\u2014\u2014 10\n74\n18\n\u2014\n8\nHeptaneb\nLi\n\u201410\n97\n\u2014\n\u2014\n\u2014-\n"]], ["block_4", ["Table 4.1\nPolymerization of Polyisoprene under Various Conditions,\nand the Resulting Microstructure in %\n"]], ["block_5", ["aInitiator concentration 6 x 10\u20193 M.\nbInitiator concentration 8 x 10\"6 M.\n"]], ["block_6", ["Source: From Hsieh, H.L. and Quirk, R.P., Anionic Polymerization, Principles and Practical Applications, Marcel Dekker,\nInc., New York, NY, 1996.\n"]], ["block_7", ["Addition steps occur primarily when the living chain end is not associated. This leads to an\ninteresting dependence of the rate of polymerization, RP, on the living chain concentration, as can\nreadily be understood as follows (recall Equation 4.2.3):\n"]], ["block_8", ["None\nLi\n25\n94\n\u2014\n\u2014\n6\nNone\nNa\n25\n\u2014\n45\n7\n48\nNone\nCs\n25\n4\n51\n8\n37\n"]], ["block_9", ["Which happens, and why? What happens when the next monomer adds? Is it the same configuration,\nor not? What does it all depend on? There is no simple answer to these questions, but we can gain a\nlittle insight into how to control the microstructure of a polydiene by looking at some data.\nTable 4.1 gives the results of chemical analysis of the microstructure of polyisoprene after\npolymerization under the stated conditions. In the first two cases, there is a strong preference for\n3,4 addition, with significant amounts of 1,2; relatively little 1,4 addition is found. The key feature\nhere turns out to be the solvent polarity, as will be discussed below. When switching to heptane, a\nnonpolar solvent, the situation is reversed; now 1,4 cis is heavily favored. Interestingly, decreasing\nthe initiator concentration by a factor of a thousand exerts a significant in\ufb02uence on the 1,4 cis/\ntrans ratio. At first glance this seems strange; the details of an addition step should not depend on\nthe number of initiators. However, the answer lies in kinetics, as the propagation step is not as\nsimple as one might naively expect. Finally, the last three entries show isoprene polymerized in\nbulk, which also corresponds to a nonpolar medium. In this case, we see that changing the\ncounterion has a huge effect. Simply replacing lithium with sodium switches the product from\nalmost all cis 1,4 to a mixture of trans 1,4 and 3,4.\nThe key factor that comes into play in nonpolar solvents is ion pairing or clustering of the living\nends. Ionic species tend to be sparingly soluble in hydrocarbons, as the dielectric constant of the\nmedium is too low. Consequently, the counterion is rather tightly associated with the carbanion,\nforming a dipole; these dipoles have a strong tendency to associate into a small cluster, with\nperhaps it 2, 4, or 6 chains effectively connected as a star molecule. This equilibrium is\nillustrated in the cartoon below for the case n 4:\nair-\n41\u20144\n"]], ["block_10", ["Solvent\nCounterion\nT, \u00b0C\n1,4 cis\n1,4 trans\n1,2\n3,4\n"]], ["block_11", ["128\nControlled Polymerization\n"]], ["block_12", [{"image_3": "142_3.png", "coords": [38, 309, 200, 446], "fig_type": "figure"}]], ["block_13", [{"image_4": "142_4.png", "coords": [189, 355, 374, 407], "fig_type": "molecule"}]], ["block_14", [{"image_5": "142_5.png", "coords": [195, 336, 364, 415], "fig_type": "figure"}]], ["block_15", [{"image_6": "142_6.png", "coords": [210, 529, 439, 646], "fig_type": "figure"}]]], "page_143": [["block_0", [{"image_0": "143_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a suitable acidic proton source, such as methanol, will cap the growing chain and produce the\ncorresponding salt, for example, Li +OCH3\u2014 . Care must be taken that the termination is conducted\nunder the same conditions of purity as the reaction itself, however. For example, introduction of\noxygen along with the terminating agent can induce coupling of two living chains. However, in\nmany cases it is desirable to introduce a particular chemical functionality at the end of the growing\nChain. One prime example is to switch to a second monomer, which is capable of continued\npolymerization to form a block copolymer. The second example is to use particular multifunctional\nterminating agents to prepare star\u2014branched polymers. These cases, and other uses of end\u2014functional\nchains, are the next subject we take up.\n"]], ["block_2", ["where we recognize [(P*)4] m (1/4)[P*], as most of the chains are in aggregates, and that the\napparent rate constant kappzkp(KdiS/4)U4. The rate of polymerization is therefore first order in\nmonomer concentration, as one should expect, but has a (l/n) fractional dependence on initiator\nconcentration, is the average size. Accurate experimental determination of n is\ntricky, but a large body of data exists. It should also be noted that there is in all likelihood a\ndistribution of of association or ion clustering, so that the actual situation is considerably\nmore complicated than implied by Reaction (4F).\nIncreasing the size of the counterion increases the separation between charges at the end of the\ngrowing chain, thereby facilitating the insertion of the next monomer. The concentration of\ninitiator can also in\ufb02uence:2, presumably by the law of mass action. The dependence of the cis\nisomer concentration in heptane indicated in Table 4.1 is actually thought to be the result of a more\nsubtle effect than however. Itis generally accepted that the cis configuration is preferred\nimmediately after addition of a monomer, but that isomerization to trans is possible, within an\naggregate, given time. The rate of isomerization is proportional to the concentration of chains in\naggregates and therefore proportional to [P*], whereas the rate of addition is proportional to a\nfractional power of [P*]. Increasing the initiator concentration increases both rates, but favors\nisomerization relative to propagation.\nTermination of an anionic polymerization is a relatively straightforward process; introduction of\n"]], ["block_3", ["Before addressing the preparation of block copolymers by anionic polymerization, it is appropriate to\nconsider some of the reasons why block copolymers are such an interesting class of macromolecules.\n"]], ["block_4", ["4.4\nBlock Copolymers, End-Functional Polymers, and Branched\nPolymers by Anionic Polymerization\n"]], ["block_5", ["The central importance of living anionic polymerization to current understanding of polymer\nbehavior cannot be overstated. For example, throughout Chapter 6 through Chapter 13 we will\nderive a host of relationships between observable physical properties of polymers and their\nmolecular weight. These relationships have been largely confirmed or established experimentally\nby measurements on narrow molecular weight distribution polymers, which were prepared by\nliving anionic methods. However, it can be argued that even more important and interesting\napplications of living polymerization arise in the production of elaborate, controlled architectures;\nthis section touches on some of these possibilities.\n"]], ["block_6", ["4.4.1\nBlock Copolymers\n"]], ["block_7", ["Inserting \nR, :kp\u2014\nBr\n2\nn\nm\n+ \ufb01Br --\u2014-* (styrene)n-\u2014\u2014\u2014(methyl methacrylate)2m\u2014(styrene)n\nO\nO O\nO\n|\n|\n9\nMe\n+ 2 LiBr\nM\n(4.G)\n"]], ["block_8", [{"image_4": "146_4.png", "coords": [48, 620, 266, 684], "fig_type": "figure"}]], ["block_9", ["H\nOH\n' _\n+\nO\nH+\nn+1\n[.1 \n3]\nSwo-\nn\n\u201c+Oj\u2014\u2014F\u2014)-\nn+1\n"]], ["block_10", [{"image_5": "146_5.png", "coords": [173, 617, 435, 681], "fig_type": "molecule"}]], ["block_11", [{"image_6": "146_6.png", "coords": [284, 622, 410, 679], "fig_type": "molecule"}]], ["block_12", ["(4.H)\n"]], ["block_13", ["+\n"]]], "page_147": [["block_0", [{"image_0": "147_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["end\u2014functional polymer with\nfunctionality at one end, or at both ends. Such polymers are also referred to as telechelic. It should\nbe apparent that most condensation polymers have reactive groups at each end, and thus fall in\nthis class. here with polymers that have narrow molecular weight\ndistributions In essence, an end-functional is\nmacromolecular characterized and then stored on the shelf until needed\nfor particular application. is a list of a few of the many examples of possible uses\nfor end-functional polymers:\n"]], ["block_2", ["The previous illustration of the macroinitiator approach is an excellent example of the utility of an\n"]], ["block_3", ["4.4.2\nEnd-Functional Polymers\n"]], ["block_4", ["1.\nMacroinitiators. As illustrated in the previous section, a macroinitiator is an end-functional\npolymer in which the functional group can be used to initiate polymerization of a second\nmonomer. In this way, block copolymers can be prepared that are not readily accessible by\nsequential monomer addition. Indeed, the second block could be polymerized by an entirely\ndifferent mechanism than the first; other living polymerization schemes will be discussed in\nsubsequent sections.\n2.\nLabeled polymers. It is sometimes desired to attach a \u201clabel\u201d to a particular polymer, such as a\n\ufb02uorescent dye or radioactive group, which will permit subsequent tracking of the location of\nthe polymer in some process. By attaching the label to the end of the chain, the number of\nlabels is well-defined, and labeled chains can be dispersed in otherwise equivalent unlabeled\nchains in any desired proportion.\n3.\nChain coupling. Both block copolymers and regular branched architectures can be accessed by\ncoupling reactions between complementary functionalities on different chains.\n4.\nMacromonomers. If the terminal functional group is actually polymerizable, such as a carbon\u2014\ncarbon double bond, polymerization through the double bond can produce densely branched\ncomb or \u201cbottlebrush\u201d copolymers.\n5.\nGrafting to surfaces. As mentioned in the context of copolymer adsorption to a surface, a densely\npacked layer of polymer chains emanating from a surface forms a brush. Such brushes can also\nbe prepared by the grafting of end-functional chains, where the functionality is tailored to react with\n"]], ["block_5", ["7.\nNetwork precursors. Telechelic polymers can serve as precursors to network formation, when\ncombined with suitable multifunctional linkers or catalysts. For example, some silicone\nadhesives contain poly(dimethylsiloxane) chains with vinyl groups at each end. In the\nunreacted form, these polymers form a low-viscosity \ufb02uid that can easily be mixed with\ncatalyst and spread on the surfaces to be joined; the subsequent reaction produces an adhesive,\nthree-dimensional network in situ.\n8.\nReactive compatibilization. As noted previously, block copolymers can act as macromolecular\nsurfactants to stabilize dispersions of immiscible homopolymers. However, direct mixing of\nblock copolymers during polymer processing is not always successful, as the copolymers have\n"]], ["block_6", ["the surface. High grafting densities are hard to achieve by this strategy, however, due to steric\ncrowding; the first chains anchored to the surface make it progressively harder for further chain\nends to react.\n6.\nControlled-branched and cyclic architectures. Examples of branched structures will be given\nin the following section. Cyclic polymers can be prepared by intramolecular reaction of an\n\u201cor,m-heterotelechelic\u201d linear precursor, where the two distinct end groups can react. Such\nring-closing reactions have to be run at extreme dilution, to suppress interchain end linking.\n"]], ["block_7", ["Block Copolymers. End-Functional Polymers, and Branched Polymers\n133\n"]], ["block_8", ["a tendency to aggregate into micelles and never reach the interface between the two polymers.\nOne effective way to overcome this is to form the block copolymer at the targeted interface, by\nin situ reaction of suitable functional chains. Note that in this case it is not absolutely\nnecessary that the reactive groups be at the chain ends.\n"]]], "page_148": [["block_0", [{"image_0": "148_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "148_1.png", "coords": [15, 436, 312, 607], "fig_type": "figure"}]], ["block_2", ["Table 4.2\nExamples of Protection Strategies for Preparing End\u2014Functional\nPolymers by Living Anionic Polymerization of Styrenes and Dienes\n"]], ["block_3", ["Termination by short alkanes with a halide at one end and the protected functionality at the other.\n"]], ["block_4", ["Source: From Hirao, A. and Hoyashi, M., Acta Polymerica, 50, 219, 1999.\n"]], ["block_5", ["There are two general routes to end-functional chains: use a functional initiator or use a\nfunctional terminating group. The use of a functional terminating agent proves to be the more\n\ufb02exible strategy for a rather straightforward reason. Any functional group present in the initiator\nmust be inert to the polymerization, which can be problematic in the case of anionic polymeriza-\ntion. Thus, the functional group in the initiator must be protected in some way. In contrast, for the\nterminating agent all that is required are two functionalities: the desired one and another electro-\nphilic one to terminate the polymerization. However, the functionality that is designed to terminate\npolymerization must be substantially more reactive to carbanions than the other functionality, or\nmore than one chain end structure will result. Consequently, in most cases a protection strategy is\nalso employed for the terminating agent. Nevertheless, in the termination case the demands on the\nprotecting group are much reduced relative to initiation; in the former, the protecting group only\nneeds to be significantly less reactive than the electrophile, whereas in the latter the protecting\ngroup must be substantially less reactive than the monomer.\nFor the living anionic polymerization of styrene, butadiene, and isoprene, an effective termin-\nating strategy is to use alkanes that have bromo functionality at one end and the protecting group at\nthe other. The halide is very reactive to the carbanion, readily eliminating the LiBr salt as the chain\nis terminated. Of course, the protecting group must then be removed in a separate step. Examples\nof protecting groups and the desired functionalities are given in Table 4.2. Some of the same\nprotecting groups illustrated in Table 4.2 can also be used in functional initiators. For example, the\ntert-butyl dimethylsilyl moiety used to protect the thiol group can also be used to protect a\nhydroxyl group in the initiator, as in (3-(tert-buty1 dimethylsilyloxy-l-propyllithium)).\nAnother powerful strategy for preparing end-functional polymers by anionic polymerization\nwas implicitly suggested in the previous section, where addition of a nominally polymerizable\nmonomer (ethylene oxide in that instance) to a growing polystyryl anion resulted in the addition of\nonly one new monomer. It turns out that 1,1-dipheny1ethylene and derivatives thereof will only\nreact with organolithium salts to form the associated relatively stable carbanion; no further\npropagation occurs:\n"]], ["block_6", ["Si(CH3)3\n-NH2\n[:1/ \\Si(CH3)3\n"]], ["block_7", ["-COOH\n/C(OCH3)3\n"]], ["block_8", ["\"CECH\n\u2014CEC\"Si(CH3)3\n"]], ["block_9", ["Functional Group\nProtected functionality\n"]], ["block_10", ["._\nO\nOH\n/ \\Si(CH3)3\n"]], ["block_11", ["134\nControlled Polymerization\n"]], ["block_12", [{"image_2": "148_2.png", "coords": [228, 449, 292, 475], "fig_type": "molecule"}]], ["block_13", ["Sr((CH3)2t-Bu)\n"]]], "page_149": [["block_0", [{"image_0": "149_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "149_1.png", "coords": [31, 538, 287, 633], "fig_type": "figure"}]], ["block_2", ["This multifunctional terminating agent is then introduced directly into the reaction vessel\ncontaining the living polystyryl chains. The chains should be in stoichiometric excess to minimize\n"]], ["block_3", ["The kinds of synthetic methodology suggested in the previous section have been adapted to the\npreparation of a wide range of polymer structures with controlled branching [4]. The first\narchitecture to consider is that of the regular star, in which a predetermined number of equal\nlength arms are connected to a central core. There are two general strategies to prepare such a\npolymer by living anionic polymerization: use a multifunctional initiator, and grow the arms\noutwards simultaneously, or use a multifunctional terminating agent to link together premade\narms. The first route is an example of an approach known as grafting from, whereas the second is\ntermed grafting r0. Or, in anticipation of the discussion of dendrimers in Section 4.9, grafting from\nand grafting to are analogous to divergent and convergent synthetic strategies. Although both have\nbeen used extensively, grafting to is more generally applicable to anionic polymerization due to the\ndifficulty in preparing and dissolving small molecules with multiple alkyllithium functionalities.\nFurthermore, in order to achieve uniform arm lengths, it is essential that each initiation site be\nequally reactive and equally accessible to monomers in the reaction medium. If it is desired to\nterminate each star arm with a functional group, however, then grafting from may be preferred.\nShould the anionic polymerization be initiated by a potassium alkoxide group, as for example with\nthe polymerization of ethylene oxide suggested in the context of Reaction (4.H), then preparation of\ninitiators with multiple hydroxyl groups is quite feasible (see Reaction 4.EE for a specific example).\nSimilarly, if other living polymerization routes are employed, such as controlled radical polymer-\nization to be discussed in Section 4.6, then grafting from is more convenient than in the anionic case.\nThe preparation of an eight-arm polystyrene star by grafting to is illustrated in the following\nscheme. The most popular terminating functionality in this context is a chlorosilane, which reacts\nrapidly and cleanly with many polymeric carbanions, and which can be prepared with function-\nalities up to at least 32 without extraordinary effort. An octafunctional chlorosilane can be prepared\nstarting with tetravinylsilane and dichloromethylsilane, using platinum as a catalyst:\n"]], ["block_4", ["In this structure R\u2019 and R\u201d could be any of a variety of protected or even unprotected functional-\nities. Even more interesting is the fact that this carbanion can be used to initiate anionic polymer\u2014\nization of a new monomer (such as methyl methacrylate, dienes, etc.) or even to reinitiate the\npolymerization of styrene. In this way, diphenylethylene derivatives can be used to place particular\nfunctional groups at desired locations along a homopolymer or copolymer, not just at the terminus.\n"]], ["block_5", ["4.4.3\nRegular Branched Architectures\n"]], ["block_6", [{"image_2": "149_2.png", "coords": [35, 42, 282, 154], "fig_type": "figure"}]], ["block_7", ["Block \n135\n"]], ["block_8", [{"image_3": "149_3.png", "coords": [40, 556, 274, 627], "fig_type": "molecule"}]], ["block_9", ["R!\nR\u201d\n,.\nH\nO\nO\n\"\n+Li\n+\nH0\n\u2014+\n\u2014\n+-\n(4.1)\n2\nLI\n"]], ["block_10", ["~\u2014\\-\u2014/S\u20187\n"]], ["block_11", [{"image_4": "149_4.png", "coords": [49, 69, 179, 145], "fig_type": "molecule"}]], ["block_12", ["Me Cl\n\\ /\nCl\u2014Si\nl/\n2\n9'\n"]], ["block_13", ["+ CH3SIC|2H \u2014\u2014-\u2014 CIWSIA\u2019SK CI\n(4.1)\n/\nM9 CI\n8\nSi-Cl\n/ \\\nCI Me\n"]], ["block_14", [{"image_5": "149_5.png", "coords": [144, 537, 275, 640], "fig_type": "molecule"}]], ["block_15", [{"image_6": "149_6.png", "coords": [144, 44, 269, 159], "fig_type": "molecule"}]], ["block_16", ["R!\nR.\u201d\n"]]], "page_150": [["block_0", [{"image_0": "150_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "150_1.png", "coords": [14, 278, 462, 376], "fig_type": "figure"}]], ["block_2", ["the formation of a mixture of stars with different numbers of arms. This will necessitate separation\nof the unattached arms from the reaction mixture, but this is feasible. Moreover, an additional\nadvantage of the grafting to approach is thus exposed: the unattached arms can be characterized\n(for molecular weight, polydispersity, etc.) independently of the stars themselves, a desirable step\nthat is not possible when grafting from.\nThe scheme just outlined is not quite as straightforward as it might appear. The key issue is to make\nall eight terminating sites accessible to the polystyryl chains. As the number of attached arms grows, it\nbecomes harder and harder for new chain ends to find their way into the reactive core. In order to\nreduce these steric effects, more methylene groups can be inserted into the terminating agent to spread\nout the chlorosilanes. In some cases, polystyryl chains have been capped with a few butadienyl units to\nreduce the steric bulk of the chain end. Clearly, all of these issues grow in importance as the number of\narms increases. Note, however, that it is not necessary that all the chlorosilanes be equally reactive in\norder to preserve a narrow molecular weight distribution; it is only necessary that the attachment of the\nnarrowly distributed arms be driven to completion (which may take some time).\nAs the desired number of arms increases, it is practical to surrender some control over the exact\nnumber of arms in favor of a simpler method for termination. A scheme that has been refined to a\nconsiderable extent is to introduce a difunctional monomer, such as divinylbenzene, as a poly-\nmerizable linking agent. The idea is illustrated in the following reaction:\n"]], ["block_3", ["One divinylbenzene molecule can thus couple two polystyryl chains and leave two anions for\nfurther reaction. Each anion might add one more divinylbenzene, each of which could then add one\nmore polystyryl chain. At that point, the growing star molecule would have four arms, emanating\nfrom a core containing three divinylbenzene moieties and four anions. This process can continue\nuntil the divinylbenzene is consumed and the anions terminated. Clearly, there is potential for\n"]], ["block_4", ["a great deal of variation in the resulting structures, both in the size of the core and in the number\nof arms. However, by carefully controlling the reaction conditions, and especially the ratio of\ndivinylbenzene to living chains, reasonably narrow distributions of functionality can be obtained,\nwith average numbers of arms even exceeding 100.\nThe preceding strategy can actually be classified as grafting through, a third approach that is\nparticularly useful for the preparation of comb polymers. A comb polymer consists of a backbone to\nwhich a number of polymeric arms are attached; combs can be prepared by grafting from, grafting to,\nand grafting through. In the first case, the backbone must contain reactive sites that can used to initiate\npolymerization. The backbone can be characterized independently of the arms, but the arms them-\nselves cannot. In grafting to, the backbone must contain reactive sites such as chlorosilanes that can\nact to terminate the polymerization of the arms. Clearly in this case, as with stars, the arms and the\nbackbone can be characterized independently. The grafting through strategy takes advantage of what\nwe previously termed macromonomers: the arms are polymers terminated with a polymerizable group.\nThese groups can be copolymerized with the analogous monomers to generate the backbone. By\nvarying the ratio of macromonomer to comonomer, the spacing of the \u201cteeth\u201d of the comb can be\ntuned. Note that this process is not necessarily straightforward. In Chapter 5 we will consider\ncopolymerization in great detail, but a key concept is that of reactivity ratio. This refers to the relative\nprobability of adding one monomer to a growing chain, depending on the identity of the previous\nmonomer that attached. It is generally the case that there are significant preferences (i.e., the\nreactivity ratios of the two monomers are not unity), which means that the two monomers will not\n"]], ["block_5", ["136\nControlled Polymerization\n"]], ["block_6", [{"image_2": "150_2.png", "coords": [37, 303, 244, 370], "fig_type": "molecule"}]], ["block_7", ["H\nH\nx\n\\\n2\nH\nn\n_\n+\nCH2\n{1\nl\n(4K)\n0 O\n/\nH\n\u2014\nn\n"]], ["block_8", [{"image_3": "150_3.png", "coords": [45, 300, 246, 368], "fig_type": "figure"}]], ["block_9", [{"image_4": "150_4.png", "coords": [228, 270, 416, 377], "fig_type": "figure"}]], ["block_10", [{"image_5": "150_5.png", "coords": [233, 289, 435, 379], "fig_type": "molecule"}]]], "page_151": [["block_0", [{"image_0": "151_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "151_1.png", "coords": [31, 206, 327, 259], "fig_type": "molecule"}]], ["block_2", ["This macromonomer can then be copolymerized with methyl methacrylate, to produce a comb or\ngraft copolymer, witha poly(methyl methacrylate) backbone and polystyrene arms:\n"]], ["block_3", ["1.\nA single initiator species is often not sufficient in cationic polymerizations; frequently a\nsecond ingredient (or cocatalyst) is required.\n2.\nTotal dissociation of the cationic initiator is rather rare, which has implications for the ability\nto start all the chains growing at the same time.\n3.\nAlthough both ionic mechanisms clearly eliminate termination by direct recombination of\ngrowing chains, cationic species are much more prone to transfer reactions than their anionic\ncounterparts. Consequently, living cationic polymerization is much less prevalent than living\nanionic polymerization.\n4.\nMost monomers that can be readily polymerized by anionic mechanisms are also amenable to\nfree-radical polymerization. Thus, in commercial practice the rather more demanding anionic\nroute is only employed when a higher degree of control is required, for example, in the\npreparation of styrene\u2014diene block copolymers.\n5.\nIn contrast, although most monomers that can be polymerized by cationic mechanisms are also\namenable to free-radical polymerization, there are important exceptions. The most significant\nfrom a total production point of view is polyisobutylene (butyl rubber), which is produced\ncommercially by (both living and nonliving) cationic polymerization.\n"]], ["block_4", ["This last example reminds us that the variety of possible controlled branched architectures is\ngreatly enhanced when different chemistries are used for different parts of the molecule. If we\nconfine ourselves to the case of stars, a molecule in which any two arms differ in a deliberate and\nsigni\ufb01cant way has been termed a miktoarm star, from the Greek word for mixed [4]. A whole host of\ndifferent structures have been prepared in this manner. For example, an AZB miktoarm star contains\ntwo equal length arms of polymer A and one arm of polymer B. Among the structures that have been\nreported are AZB, A3B, A2B2, A4B4, and a variety of ABC miktoarm terpolymers. It is even possible\nto produce asymmetric stars, in which the arms consist of the same polymer but differ in length.\n"]], ["block_5", ["add completely to be understood before regular comb molecules with\nvariable branching density can be prepared by grafting through.\nThe grafting through approach can be illustrated through the following sequence [5]. Polystyryl\nchains 4.H) followed \nmethacryloyl \n"]], ["block_6", ["4.5\nCationic Polymerization\n"]], ["block_7", ["Just as anionic polymerization is a chain-growth mechanism that shares important parallels with the\nfree-radical route, so too cationic polymerizations can be discussed within the same framework:\ninitiation, propagation, termination, and transfer. However, there are important differences between\nanionic and cationic polymerizations that have direct impact on the suitability of the latter for living\npolymerization. The principal differences between the two ionic routes are the following:\n"]], ["block_8", ["Cationic Polymerization\n137\n"]], ["block_9", ["H\n.\na\u2014m\n.\nW\u2018\n(W\nCH2\n0\n'\n"]], ["block_10", ["CH2\n0\nCH\n_ _\nH\nn\nomiMe\n+\nMew/\u201c\\O/\n3 PMMA 9 PS\n(4 M)\n0\nCH2\n\u2018\n"]], ["block_11", [{"image_2": "151_2.png", "coords": [73, 115, 331, 177], "fig_type": "molecule"}]], ["block_12", ["CH2\n0\n"]]], "page_152": [["block_0", [{"image_0": "152_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "152_1.png", "coords": [13, 77, 366, 243], "fig_type": "figure"}]], ["block_2", [{"image_2": "152_2.png", "coords": [32, 490, 245, 558], "fig_type": "figure"}]], ["block_3", ["Note: Parentheses indicate not readily polymerized by this route.\n"]], ["block_4", ["Table 4.3\nGeneral Summary of Polymerizability of Various Monomer Types by\nthe Indicated Chain-Growth Modes\n"]], ["block_5", ["Ethylene\n\\/\n(\\/)\n><\na-Olefins\n><\n\\/\n(\\/)\n1,1-Dialkyl alkenes\nx\nx\n\\/\nHalogenated alkenes\n\\/\nx\nx\n1,3-Dienes\n\\/\n\\/\n\\/\n"]], ["block_6", ["Source: Adapted from Odian, G., Principles of Polymerization, 4th ed, Wiley-Interscience,\nHoboken, NJ, 2004.\n"]], ["block_7", ["4.5.1\nAspects of Cationic Polymerization\n"]], ["block_8", ["With insufficient cocatalyst these equilibria lie too far to the left, while excess cocatalyst can terminate\nthe chain or destroy the catalyst. Thus, the optimum proportion of catalyst and cocatalyst varies with\nthe speci\ufb01c monomer and polymerization solvent. In the case of protonic acids, the concentration of\nprotons depends on the position of the standard acid\u2014base equilibria, but in the chosen organic solvent:\n"]], ["block_9", ["Styrenes\n\\/\n\\/\n\\/\nAcrylates, methacrylates\n\\/\n\\/\n><sub>><></sub>sub><sub><sub>><></sub></sub><sub>><></sub>/sub>\nAcrylonitrile\n\\/\n\\/\n><sub>><></sub>sub><sub><sub>><></sub></sub><sub>><></sub>/sub>\nAcrylamide, methacrylamide\n\\/\n\\/\n><sub>><></sub>sub><sub><sub>><></sub></sub><sub>><></sub>/sub>\nVinyl esters\n\\/\n><sub>><></sub>sub><sub><sub>><></sub></sub><sub>><></sub>/sub>\n\\/\nVinyl ethers\nx\nx\n\\/\nAldehydes, ketones\n><sub>><></sub>sub><sub><sub>><></sub></sub><sub>><></sub>/sub>\n\\/\n\\/\n"]], ["block_10", ["Monomer\nRadical\nAnionic\nCationic\n"]], ["block_11", ["A brief summary of the applicability of the three chain\u2014growth mechanisms\u2014radical, anionic,\ncationic\u2014to various monomer classes is presented in Table 4.3. In the remainder of this section we\ndescribe general aspects of cationic polymerization and introduce some of the transfer reactions\nthat inhibit living polymerization. Then, we conclude by discussing the strategies that have been\nused to approach a living cationic polymerization.\n"]], ["block_12", ["In cationic polymerization, the active species is the ion formed by the addition of a proton\nfrom the initiator system to a monomer (partly for this reason the initiator species is often called a\ncatalyst, because it is not incorporated into the chain). For vinyl monomers the substituents which\npromote this type of polymerization are electron donating, to stabilize the propagating carboca-\ntion; examples include alkyl, 1,1-dialky1, aryl, and alkoxy. Isobutylene, a-methylstyrene, and vinyl\nalkyl ethers are examples of monomers commonly polymerized via cationic intermediates.\nThe initiator systems are generally Lewis acids, such as BF3, AlCl3, and TiCl4, or protonic\nacids, such as H2804, HClO4, and HI. In the case of the Lewis acids, 3 proton-donating coinitiator\n(often called a cocatalyst) such as water or methanol is typically used:\n"]], ["block_13", ["138\nControlled Polymerization\n"]], ["block_14", [{"image_3": "152_3.png", "coords": [36, 599, 161, 682], "fig_type": "molecule"}]], ["block_15", ["CH30H + TiCI4 \nH+ + C|4TiOCHg\n"]], ["block_16", ["H2804\n\u2014__=\u2018\u2014-_\nH+ + H804\"\n"]], ["block_17", ["H20 A'C'a\n\u2014\u2018-\u2014.\u2014\nH+ + CI3AIOH\u2018\n(4.N)\n"]], ["block_18", ["H20 + BF3\n=\nH+ + FsBOH'\n"]], ["block_19", ["Helo4\n_..__ H+ + CIO4\u2018\n(4.0)\n"]], ["block_20", ["H|\nH+ + l\u2014\n"]], ["block_21", [{"image_4": "152_4.png", "coords": [137, 90, 334, 228], "fig_type": "figure"}]]], "page_153": [["block_0", [{"image_0": "153_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "153_1.png", "coords": [28, 97, 219, 129], "fig_type": "molecule"}]], ["block_2", [{"image_2": "153_2.png", "coords": [29, 584, 200, 651], "fig_type": "molecule"}]], ["block_3", [{"image_3": "153_3.png", "coords": [33, 443, 90, 505], "fig_type": "molecule"}]], ["block_4", ["The structures expected and found are sketched here:\n"]], ["block_5", ["One of the side reactions that can cationic polymerization is the possibility of the ionic\nrepeat unit undergoing rearrangement during the polymerization. The following example illus\u2014\ntrates this situation.\n"]], ["block_6", ["The electron\u2014donating group helps to stabilize this \npolymerization, the separation of the ions and the possibility of ion pairing play important roles in\n"]], ["block_7", ["If we write formula the initiator system as H+B_, then the initiation and\npropagation steps monomer CH2 =CHR can be written as follows. The proton adds\nto the more electronegative carbon atom in the olefin to initiate chain growth:\n"]], ["block_8", ["Expected\n"]], ["block_9", ["Found\n"]], ["block_10", ["The conversion of the cationic intermediate of the monomer to the cation of the product occurs by\n"]], ["block_11", ["It has been observed that poly(1,1\u2014dimethyl propane) is the product when 3-methylbutene-1\n(CH2:CH\u2014CH(CH3)2) is polymerized with Al in ethyl chloride at \u2014130\u00b0C.T Draw structural\nformulas for the expected and observed repeat units, and propose an explanation.\n"]], ["block_12", ["Aldehydes can also be polymerized in this fashion, with the corresponding reactions for formal-\ndehyde being\n"]], ["block_13", ["the ease proceeds \n"]], ["block_14", ["Cationic \n139\n"]], ["block_15", ["a hydride shift between adjacent carbons:\n.\n"]], ["block_16", ["\u201dP. Kennedy and RM. Thomas, Makromol. Chem, 53, 28 (1962).\n"]], ["block_17", ["Solution\n"]], ["block_18", ["Example 4.2\n"]], ["block_19", [{"image_4": "153_4.png", "coords": [38, 512, 98, 545], "fig_type": "molecule"}]], ["block_20", [{"image_5": "153_5.png", "coords": [43, 177, 236, 208], "fig_type": "molecule"}]], ["block_21", [{"image_6": "153_6.png", "coords": [43, 251, 337, 303], "fig_type": "molecule"}]], ["block_22", [{"image_7": "153_7.png", "coords": [46, 260, 223, 299], "fig_type": "molecule"}]], ["block_23", ["R\nR\nR\nR\nMe\u2014<+B\u2018 +H2CZ< \u2014\u00bbMe>t/\\r+\n-\n(4Q)\nH\nH\nH\nH\n'3\n"]], ["block_24", ["H\nO=<\n+\nH\nH\nH\nH+\n(4R)\n\u2018\n\u2014\u2014\u2014+ H\n+\n-\nH\n-\nH\nB\n+0= Me\u2014(+\nB-\n(4.13)\n"]], ["block_26", ["Me\nMe\n"]], ["block_27", [{"image_8": "153_8.png", "coords": [55, 596, 195, 635], "fig_type": "molecule"}]], ["block_28", ["n\nMe Me\n"]], ["block_29", ["H\nH\n"]], ["block_30", ["Fl\n"]], ["block_31", ["Me\n"]], ["block_32", [{"image_9": "153_9.png", "coords": [127, 180, 230, 212], "fig_type": "molecule"}]], ["block_33", [{"image_10": "153_10.png", "coords": [200, 255, 323, 298], "fig_type": "molecule"}]]], "page_154": [["block_0", [{"image_0": "154_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "154_1.png", "coords": [24, 205, 385, 245], "fig_type": "molecule"}]], ["block_2", ["1.\nClearly, the reaction must be conducted in the absence of nucleophilic species that are capable\nof irreversible termination of the growing chain.\n2.\nSimilarly, the reaction should be conducted in the absence of bases that can participate in\nB\u2014proton transfer. As discussed above, the monomer itself is such as base, and therefore\n"]], ["block_3", ["The preceding discussion provides some insight into the obstacles to achieve a living cationic\npolymerization. Nevertheless, living cationic polymerization is by now a relatively common tool,\nand many of the controlled architectures (block copolymers, end\u2014functional chains, regular branched\nmolecules) that we discussed in the context of anionic polymerization have been accessed [7]. In this\nsection, we brie\ufb02y describe the general strategy behind living cationic polymerization; recall that the\nessential elements are the absence of termination or transfer reactions:\n"]], ["block_4", ["4.5.2\nLiving Cationic Polymerization\n"]], ["block_5", ["4.\nSpontaneous termination. This process, also known as chain transfer to counterion, is essen-\ntially a reversal of the initiation step, as a B-proton is transferred back to the anion (e.g., as in\nReaction 4.P, but with a growing chain rather than the first monomer).\n"]], ["block_6", ["This is a well-known reaction that is favored by the greater stability of the tertiary compared to the\nsecondary carbocation.\n"]], ["block_7", ["In the particular case of isobutylene the resulting primary carbocation is less stable than the\ntertiary one on the chain, so Reaction (4.T) is less of an issue than Reaction (4.8).\n3.\nIntermolecular hydride transfer. This is an example of transfer to polymer, and can be written\ngenerally as\n"]], ["block_8", ["The preceding example illustrates one of the potential complications encountered in cationic\npolymerization, but it is not in itself an impediment to living polymerization. There are several\nother potential transfer reactions, however, that collectively do impede a living cationic polymeri-\nzation. Four of these are the following:\n"]], ["block_9", ["The activated monomer can now participate in propagation reactions, whereas the previous\nchain is terminated. Note that in isobutylene there are two distinct B\u2014protons, and thus two\npossible structures for the terminal unsaturation of the chain. There is also a possibility that\nthese double bonds can react subsequently.\n2.\nHydride transfer from monomer. In this case, the transfer proceeds in the opposite direc-\ntion, but has the same detrimental net effect from the point of view of achieving a living\npolymerization:\n"]], ["block_10", ["14o\nControlled Polymerization\n"]], ["block_11", ["1.\nB\u2014Proton transfer. This is exemplified by the case of polyisobutylene. Protons on carbons\nadjacent (B) to the carbocation are electropositive, due to a phenomenon known as hypercon-\njagation; we can view this as partial electron delocalization through 0 bonds, in contrast\nto resonance, which is delocalization through 17 bonds. Consequently there is a tendency for\nB-protons to react with any base present, such as a vinyl monomer:\n"]], ["block_12", ["Me\nMe\nMe\nMe\nMe\n\u201d\\\\ufb02\n3\u2014\n+ H20=< \u2014+ H30\u2014<+\nB\u2014 +\u201d\\n\u2019\n+ \u201dY\n(4.8)\nMe\nMe\nMe\nCH2\nMe\n"]], ["block_13", ["+ Wyn/i9\nMe\n(4.T)\nMe\nMe\nMe\nN\u2019J\\]/+\nB\u2014\n+ HZC=<\n\u2014-\u2014+\nHZC=<\n"]], ["block_14", [{"image_2": "154_2.png", "coords": [54, 210, 267, 238], "fig_type": "molecule"}]], ["block_15", ["R\nFl\n"]], ["block_16", [{"image_3": "154_3.png", "coords": [114, 430, 279, 471], "fig_type": "molecule"}]], ["block_17", [{"image_4": "154_4.png", "coords": [127, 336, 295, 377], "fig_type": "molecule"}]], ["block_18", [{"image_5": "154_5.png", "coords": [170, 207, 362, 242], "fig_type": "molecule"}]]], "page_155": [["block_0", [{"image_0": "155_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The carbon\u2014iodide bond is then activated by the relatively weak Lewis acid az to allow insertion\nof the next monomer. The transition state for pr0pagation may be represented schematically as\n.01\nto\n"]], ["block_2", ["In this sequence, the first reaction generates the initiating proton, and the second and third reactions\ncorrespond to standard irreversible initiation and propagation steps involving monomer M. The fourth\nreaction is the key. The growing cationic i-mer P? is converted to a dormant, covalent species BC] by a\nreversible reaction. While the growing chain is in this form, it does not undergo transfer or propagation\nreactions, thereby extending its lifetime. The reversible activation/deactivation reaction must be\nsufficiently rapid to allow each chain to have many opportunities to add monomer during the\npolymerization, and the relative length of time spent in the active and dormant states can be controlled\nby the position ofthe associated equilibrium. This, in turn, offers many opportunities to tune a particular\nchemical system. For example, decreasing the polarity ofthe solvent or adding an inert salt that contains\n"]], ["block_3", ["Experimentally, the system exhibits many of the characteristics associated with a living polymeri-\nzation: polydispersities consistently below 1.1; Mn increasing linearly with conversion; the ability\n"]], ["block_4", ["a common ion (chloride in this case) both push the equilibrium toward the dormant state.\nWe will revisit this idea of a dormant reactive species in the next section on controlled radical\npolymerization, where it plays the central role. We conclude this section with a specific example of\na successful living cationic polymerization scheme. Isobutyl vinyl ether (and other vinyl ethers)\ncan be polymerized by a combination of HI and az [8]. The hydrogen iodide \u201cinitiator\u201d adds\nacross the double bond, but forms an essentially unreactive species:\n"]], ["block_5", ["4.\nCationic polymerization, like most chain-growth polymerization, is often highly exothermic.\nWith the additional feature of very rapid reaction, it becomes important to reduce the rate of\npolymerization in order to remove the excess heat. Low temperature is the first option in this\nrespect, followed by lower concentrations of growing chains.\n5.\nAnother way to view control in this context is to aim to extend the lifetime ofthe growing chain. As\na point ofreference, a living polystyryl carbanion can persist for years in a sealed reaction vessel; a\npolyisobutyl carbocation will probably not last for an hour under equivalently pristine conditions.\nWhile low temperature certainly aids in increasing the lifetime, another useful strategy is to make\nthe growing center inactive or dormant for a significant fraction ofthe elapsed reaction time. This is\ndone via the process of reversible termination, as illustrated by the sequence in Reaction (4.V):\n"]], ["block_6", ["3.\nGenerally, both propagation and transfer are very rapid reactions, with transfer having the\nhigher activation energy. Lower temperatures therefore favor propagation relative to transfer,\nas well as have the advantage of bringing both reactions under better control.\n"]], ["block_7", ["Cationic \n141\n"]], ["block_8", [{"image_1": "155_1.png", "coords": [39, 502, 195, 539], "fig_type": "molecule"}]], ["block_9", [{"image_2": "155_2.png", "coords": [47, 575, 190, 627], "fig_type": "molecule"}]], ["block_10", ["HCI + Till; I: TiCl5' + H+\n"]], ["block_11", ["H++ M \nP14-\n"]], ["block_12", ["on\nonl\n+ HI\n4.W\nc=Pi++ M \u00e9\nPi+1+\n"]], ["block_14", ["Pi++TiCI5' :: PiCI +TiC|4\n"]], ["block_15", ["cationic polymerization always has a \u201cbuilt-in\u201d transfer reaction. The key step, therefore, is to\nchoose reaction conditions to maximize the rate of propagation relative to transfer, given that\ntransfer probably cannot be completely eliminated.\n"]], ["block_16", ["\\[2\u2019\nOF!\nOR\n"]], ["block_17", ["I;\n|--Zn|2\n\u2014-\u2014\u2014h-\nreznlz\n"]], ["block_18", ["(4.V)\n"]], ["block_19", ["(4.X)\n"]]], "page_156": [["block_0", [{"image_0": "156_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this section we take up the topic of controlled radical polymerization, which represents one of\nthe most active \ufb01elds in polymer synthesis in recent years. The combination of the general\nadvantages of radical polymerization (a wide range of suitable monomers, tolerance to many\nfunctional groups, characteristically rapid reactions, relatively relaxed polymerization conditions)\nwith the unique features of a living polymerization (narrow molecular weight distributions,\ncontrolled molecular weights, end functionality, block copolymers, and other complex architec-\ntures) has tremendous appeal in many different areas of polymer science. In this section we outline\nfirst in general terms how this combination is achieved, and then give some specific examples of\nthe mechanistic details. We choose the term \u201ccontrolled\u201d rather than \u201cliving\u201d in this section,\nbecause irreversible termination reactions cannot be rigorously excluded.\n"]], ["block_2", ["to resume polymerization after addition of a new charge of monomer. These aspects are illustrated\nin Figure 4.5. The mechanism implied by Reaction (4.W) and Reaction (4.X) is consistent with the\nexperimental observation that Mn is inversely proportional to the concentration of HI, but inde-\npendent of the concentration of ZnIz. On the other hand, the polymerization rate increases with\nadded 21112. az is apparently a sufficiently mild activator that the polymerization is still living at\nroom temperature when conducted in toluene, whereas in the more polar solvent methylene\nchloride lower temperatures are required.\n"]], ["block_3", ["4.6\nControlled Radical Polymerization\n"]], ["block_4", ["Figure 4.5\nLiving cationic polymerization of isobutyl vinyl ether in methylene chloride at \u201440\u00b0C. The\ncalculated curve indicates the expected molecular weight assuming 100% initiator (HI) efficiency. After 100%\nconversion a new charge of monomer was added, demonstrating the ability of the chain ends to resume\npr0pagati0n. (Reproduced from Sawamoto, M., Okamoto, C., and Higashimura, T., Macromolecules, 20,\n2693, 1987. With permission.)\n"]], ["block_5", ["4.6.1\nGeneral Principles of Controlled Radical Polymerization\n"]], ["block_6", ["The first task is to resolve the apparent paradox: given that radicals can always combine to undergo\ntermination reactions, how do we approximate a living polymerization? To develop the answer, it\nis helpful to start by summarizing once again the essence of a chain-growth polymerization in\nterms of initiation, propagation, and termination rates:\n"]], ["block_7", ["142\nControlled Polymerization\n"]], ["block_8", [{"image_1": "156_1.png", "coords": [36, 47, 275, 247], "fig_type": "figure"}]], ["block_9", ["X\n0\nIE:\nIEC\nX\n.\n:\n[EC\n"]], ["block_10", ["\u20185;\n10\n-\nl\n"]], ["block_11", ["15 \nCalcd\n"]], ["block_12", ["5\n-\n5\n\u2014\n"]], ["block_13", ["0\nr\n\u2018\nI\nI\nI\n1.0\n"]], ["block_14", ["0\n100\n200\n"]], ["block_15", ["A\n. no?\n2 A\n\u2014 1 1\n"]], ["block_16", ["E\n\u2014- 1 2\nO\nI\nA\nA\ufb01.\n"]], ["block_17", ["Conversion, %\n"]], ["block_18", ["Monomer\n0\naddition\n"]]], "page_157": [["block_0", [{"image_0": "157_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The process of controlled radical polymerization can now be seen to take place as follows. The\ndormant species PX spontaneously dissociates into the active radical and the inert partner X.\nThe exposed radical may then undergo propagation steps, or simply recombine with X so that no net\nreaction takes place. If each radical spends most of its time in the dormant state, the instantaneous\nconcentration of radicals is small, and termination is very unlikely (but never impossible). During\nan average active period a given radical may add many new monomers, about one new monomer,\nor essentially no new monomers. It is actually the last situation that is most desirable, because it\nmeans that over time all radicals are equally likely to propagate, one monomer at a time. We can\nunderstand this concept in the following way. After the polymerization has proceeded for a\nreasonable time, so that each chain on average has experienced many active periods, the number\nof active periods per chain will follow the Poisson distribution (Equation 4.2.19). That is because\nwe are randomly distributing a large number of items (active periods) into a smaller number of\nboxes (growing chains). In the limit where the likelihood of adding a monomer per active period is\nsmall, the average number of monomers added per chain will be directly proportional to the\nnumber of active periods, and thus follow the Poisson distribution as well. Of course, we are\nneglecting any termination and transfer reactions.\nIn contrast, if radicals tend to add monomers in a burst during each active period, the molecular\nweight distribution will not be as narrow unless the total degree of polymerization involves many\nsuch bursts. In fact, the length distribution of the \u201cbursts\u201d will be the most probable distribution,\nwhich (recall Equation 2.4.10 and Equation 3.7.19) has a polydispersity approaching 2. We can\nactually rationalize an approximate expression for the polydispersity of the resulting polymers,\n"]], ["block_2", ["where species, [M] is the concentration of unreacted\nmonomer, and [19.] concentration of radicals of any length. The key to a living\npolymerization is that Rt 0, or equivalently in practice, that Rp >> Rt. From Table 3.3 and\nTable 3.4, we of kt are about four orders of magnitude larger than kp.\nTherefore, thanl\u2018i\u2019t we will need [M] to be 108 times larger\nthan [Po]. Given that [M] could be on the order of 1\u201410 mol L\"1 (i.e., in bulk or concentrated\nsolution), and therefore growing chains, will have to be\n104\u2014 10\u20148 mol L4. This is quite small, but from the calculations in Section 3.4.3 we know that it\nis quite feasible.\nWe can do even better than this, however, by a nifty trick already suggested in the context of\nliving cationic polymerization. that the absolute concentration of radical forming species\nis not that the vast majority of its time in an unreactive,\ndormant form. This is illustrated schematically below:\n"]], ["block_3", ["where PX is the dormant species and X is a group (atom or molecular fragment) that can leave and\nreattach to the radical rapidly. (Note that radical species are apparently not conserved in the way\nReaction (4.Y) is expressed, but as we will see subsequently this is not actually the case. Usually X\nis also a radical species, but one that is not capable of propagating.) If the equilibrium constant for\nthe activation process, Kw, is small, then the instantaneous [P-] will be small even if [PX] is\nreasonably large:\n"]], ["block_4", ["Controlled Radical Polymerization\n143\n"]], ["block_5", ["R. kit]\n(4.6.1)\nR, kprPoitM]\n(45-2)\nRt _\u2014_ kt[P\u00b0][P-]\n(4.6.3)\n"]], ["block_6", ["PX = P- + X\n(4.Y)\n"]], ["block_7", ["[PX]\n[P ]\nKact\n[X]\n(4.6.4)\n"]]], "page_158": [["block_0", [{"image_0": "158_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Equation 4.6.5 suggests that even if q is 10, a polydispersity of 1.1 is achievable if the total degree\nof polymerization exceeds 100. A more detailed analysis yields equations similar to Equation\n4.6.5, when the average degree of polymerization is sufficiently large [9]. It is worth noting that\nthere are several complications to this analysis, such as the fact that the value of q will actually\nchange during the polymerization, as [M] decreases.\nIt should be evident from the preceding discussion that termination processes are not rigor-\nously excluded in controlled radical polymerization, only significantly suppressed. Nevertheless,\npolydispersities MW/Mn < 1.1\u20141.2 are routinely obtained by this methodology. It should also be\nevident that the higher the average chain length, the more likely termination steps become. This\ncan be seen directly from Equation 4.6.2 and Equation 4.6.3; as time progresses, l\u2019i\u2019t remains\nessentially constant, whereas Rp decreases because [M] decreases with time. Consequently, the\nrelative likelihood of a termination event increases steadily as the reaction progresses. In fact,\nthere is really a three-way competition in designing a controlled radical polymerization scheme,\namong average molecular weight, polydispersity, and \u201cefficiency,\u201d where we use ef\ufb01ciency to\ndenote a combination of practical issues. For example, the higher the desired molecular weight,\nthe broader the distribution will become, due to termination reactions. This could be mitigated\nto some extent by running at even higher dilution, but this costs time and generates large volumes\nof solvent waste. Or, the reaction vessel could be replenished with monomer, to keep [M]\nhigh even as the reaction progresses, but this wastes monomer, or at least necessitates a\nrecovery process.\n"]], ["block_2", ["A rich variety of systems that fall under the umbrella of Reaction (4.Y) have been reported. Three\ngeneral schemes have so far emerged as the most prevalent, although there is no a priori reason\nwhy others may not become more pOpular in the years ahead. Each has particular advantages and\ndisadvantages relative to the others, but for the purposes of this discussion we are really only\ninterested in their evident success. All three have been the subject of extensive review articles; see\nfor example Refs. [10,11,12].\n"]], ["block_3", ["4.6.2\nParticular Realizations of Controlled Radical Polymerization\n"]], ["block_4", ["4.6.2.1\nAtom Transfer Radical Polymerization (ATRP)\n"]], ["block_5", ["based on what we already know. Suppose the mean number of monomers added per active period\nis q 2 1. The distribution of active periods from chain to chain still follows the Poisson distribution,\nso it is almost as though we were adding one q-length block per active period. Thus, the\npolydispersity index becomes that for the Poisson distribution (recall Equation 4.2.20) with a\nnew \u201ceffective monomer\u201d of molecular weight qMO:\n"]], ["block_6", ["In this approach the leaving group X in Reaction (4.Y) is a halide, such as a chloride or bromide,\nand it is extracted by a suitable metal, such as copper, nickel, iron, or ruthenium. The metal is\nchelated by ligands such as bipyridines, amines, and trialklyphosphines that can stabilize the metal\nin different oxidation states. A particular example of the activation/deactivation equilibrium using\ncopper bromide/2,2\u2019-bipyridine (bipy) can thus be written:\n"]], ["block_7", ["where the copper atom is oxidized from Cu(I)Br to Cu(II)Br2. Reaction (4.2) suggests that the\npolymerization could be initiated by the appropriate halide of the monomer in question, such as\nl-phenylethyl bromide when styrene is the monomer. Alternatively, a standard free-radical\n"]], ["block_8", ["144\nControlled Polymerization\n"]], ["block_9", ["MW\nm1+ \nMn\nMn\n(4.6.5)\n"]], ["block_10", ["PgBI\u2019 + CuBr(bipy)2 = B0 + CuBr2(bipy)2\n(4.2)\n"]]], "page_159": [["block_0", [{"image_0": "159_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["From itbeen reported that apparent propagation rate\nconstant on the order of 10\u20144 s\", where the apparent\nrate constant is What is the order of magnitude of the concentration\nof active \n"]], ["block_2", ["4.6.2.2\nStable Free-Radical Polymerization (SFRP)\n"]], ["block_3", ["In this variant, the leaving group X in Reaction (4.Y) is a free-radical, but a sufficiently stable one\nthat it does not initiate polymerization. The prime example of this class is the nitroxide radical,\nusually embedded in\nthe\n(2,2,6,6\u2014tetramethylpiperidinyloxy)\n\u201cTEMPO\u201d\ngroup\n[13]. When\nattached to a monomer analog or a growing polymer chain terminus through the alkoxyamine\n"]], ["block_4", ["Controlled Radical \n145\n"]], ["block_5", ["initiator of\nATRP is are amenable to this approach: styrene and substituted\nstyrenes, acrylates and methacrylates,and other vinyl monomers. Dienes and amine or carboxylic\n"]], ["block_6", ["acid\u2014containing \nThe following example illustrates some of the quantitative aspects of ATRP of styrene.\n"]], ["block_7", ["The adduct of styrene and TEMPO on the left-hand side of Reaction (4AA) can be prepared rather\nreadily, puri\ufb01ed, and stored inde\ufb01nitely. In contrast to other controlled radical polymerization\nschemes, this approach is based on a single initiating species; no cocatalyst or transfer agent is needed.\nEven in the presence of a large excess of styrene monomer, it is not until the system is brought to an\nelevated temperature such as 125\u00b0C that polymerization proceeds directly. The reaction can be run\nunder nitrogen, and the rigorous purification necessary for living ionic polymerizations is not required.\nMolecular weights well in excess of 105, with polydispersities in the range of 1.1\u20141.2, have been\nachieved. The range of accessible monomers is so far more restricted than with ATRP or reversible\naddition-fragmentation transfer (RAFT), with styrene, acrylate, and methacrylate derivatives being the\n"]], ["block_8", ["monomers of choice. However, the polymerization is relatively tolerant of functional groups, and\nmany functionalized initiators with TEMPO adducts have been designed. This makes SFRP an\nappealing alternative to living ionic polymerization for the production of end-functional polymers\n(recall Section 4.4), and by extension block copolymers and branched architectures, once the initiator\nis available.\n"]], ["block_9", ["From Equation 4.6.2, we can see that kgpp thus defined is actually equal to kp [Po ]. From Table 3.4 in\nChapter 3 we know that a typical value for kp for free-radical polymerization is 102\u2014103 L mol\u2018l s- l,\nand on this basis direct substitution tells us that [Po] is about 104 3\u20141/102'3 L 11101\u201c s\u20141 10\"5\u201410\u20147\nmol L\u2019 '. This is in line with the estimate given in the previous section, of the target concentration of\nactive radicals needed to make the rate of termination small with respect to the rate of propagation.\n"]], ["block_10", ["C\u2014ON bond, homolytic cleavage of the C\u20140 bond produces the stable TEMPO radical and an\nactive radical species. This reaction is illustrated below for the case of styrene:\n"]], ["block_11", ["Solution\n"]], ["block_12", ["lK. Matyjaszewski, T.E. Patten, and J. Xia, J. Am. Chem. 306., 119, 674 (1997).\n"]], ["block_13", ["Example \n"]], ["block_14", [{"image_1": "159_1.png", "coords": [44, 395, 245, 465], "fig_type": "molecule"}]], ["block_15", [{"image_2": "159_2.png", "coords": [45, 397, 251, 461], "fig_type": "figure"}]], ["block_16", ["Me\nOH.\u2018\nMe\n.\nO\u2018M.\nN\n'\nN\n_._\n+\n(4AA)\nMe\nMe\nMe\nMe\n"]], ["block_17", [{"image_3": "159_3.png", "coords": [125, 396, 244, 454], "fig_type": "molecule"}]]], "page_160": [["block_0", [{"image_0": "160_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "160_1.png", "coords": [27, 606, 258, 659], "fig_type": "molecule"}]], ["block_2", ["4.6.2.3\nReversible Addition-Fragmentation Transfer (RAFT) Polymerization\n"]], ["block_3", ["The principal distinction between RAFT polymerization on the one hand and ATRP or SFRP on\nthe other is that RAFT polymerization involves a reversible chain transfer, whereas the other two\ninvolve reversible chain termination. The key player in the RAFT process is the chain transfer\nagent itself; the radicals are generally provided by conventional free-radical initiators such as\nAIBN. Dithioesters (RCSSR\u2019) such as cumyl dithiobenzoate are often used; in this instance R is a\nphenyl ring and R\u2019r is a cumyl group. The growing radical chain P,\u00bb reacts with the transfer agent,\nand the cumyl group departs with the radical:\n"]], ["block_4", ["A different growing radical PJ-o can also react in the analogous manner:\n"]], ["block_5", ["1\u2018CJ. Hawker, G.G. Barclay, and J. Dao, J. Am. Chem. Soc, 118, 11467 (1996).\n"]], ["block_6", ["A 1:] mixture of the two initiators was added to styrene monomer and heated to the polymer-\nization temperature. At various times, the reaction mixture could be cooled, and analyzed. If the\nexchange of TEMPO groups was rapid, then one would expect four distinct chain populations,\nwith roughly equal proportions: one with no hydroxyls, one with a hydroxyl at each end, one\nwith a hydroxyl at the terminus, and one with a hydroxyl at the initial monomer. On the other\nhand, if there was little exchange, there should be just two populations: one with no hydroxyls\nand one with two. Liquid chromatographic analysis gave results that were fully consistent with\nthe former scenario.\n"]], ["block_7", ["The unimolecular nature of the TEMPO-based initiator, plus its susceptibility to functionalization,\noffers a convenient solution, as has been demonstrated.)r These authors prepared the styrene\u2014-\nTEMPO adduct shown in Reaction (4.AA), plus a dihydroxy-functionalized variant:\n"]], ["block_8", ["An interesting question arises upon examination of Reaction (4.AA): does each TEMPO radical\nremain associated with the same chain during the polymerization, or does it migrate freely through\nthe reaction medium? In the case of anionic polymerization in a nonpolar solvent, the counterion is\ncertainly closely associated with the chain end, due to the requirement of electrical neutrality. In\nthe case, of conventional free-radical polymerization, we considered the \u201ccaging effect\u201d that can\nseverely limit the efficiency of an initiator (see Section 3.3). In this case, the relatively high\ntemperature should enhance both the mobility of the individual species and the ability to escape\nfrom whatever attractive interaction would hold the two radical species in proximity. How could\none test this intuition experimentally?\n"]], ["block_9", ["Solution\n"]], ["block_10", ["146\nControlled Polymerization\n"]], ["block_11", ["Example 4.4\n"]], ["block_12", [{"image_2": "160_2.png", "coords": [36, 550, 259, 582], "fig_type": "molecule"}]], ["block_13", [{"image_3": "160_3.png", "coords": [42, 255, 157, 315], "fig_type": "molecule"}]], ["block_14", ["P\" RASH' \n3\nFl\nFl\n"]], ["block_15", ["..\n\u2014-\u2014\n\u00b0\n..\n4.CC\nP! HAS\nHASP,\nP' \u201dHa/\u201c\\sp,\n(\n)\n"]], ["block_16", ["OKN\nHO\n"]], ["block_17", [{"image_4": "160_4.png", "coords": [59, 616, 250, 646], "fig_type": "molecule"}]], ["block_18", ["SP,-\nSP,-\n.\n_.._\n.\n4.BB\n13R.__R.. x\n<\n)\ns\n"]], ["block_19", ["Me\nMe\nOH\n"]], ["block_20", ["Me\nMe\n"]]], "page_161": [["block_0", [{"image_0": "161_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "161_1.png", "coords": [30, 427, 194, 485], "fig_type": "molecule"}]], ["block_2", [{"image_2": "161_2.png", "coords": [32, 567, 182, 597], "fig_type": "molecule"}]], ["block_3", ["In this from chain to chain. of\nthis reaction narrow molecular weight \nthat the independent of the length of the \nchain, or between group.\nThere is an important feature of this scheme that is different from the other two controlled\nradical number of chains is not determined by the number of\ninitiators, initiators (e.g., AIBN) and those from \nRAFT agent, e.g., cumyl radicals (R\u2019\u00b0 in Reaction 4.BB). In fact, given that the decomposi-\ntion of AIBNto use an excess of the RAFT agent,\nthereby dictatingof initiated via R\u201c, which in turn is proportional to the\nconcentration of RCSSR\u2019. This is facilitated by the fact that such dithioesters have\nvery large chain transfer constants (recall Section 3.8), and thus a chain initiated by AIBN\nor by R\" is rapidly transformed into a dormant form, before achieving a significant degree\nof polymerization. The RAFT approach has been successful with a very wide variety of\ndifferent monomers.\n"]], ["block_4", ["The last term indicates that the equilibrium constant is the inverse of the equilibrium monomer\nconcentration, because the concentrations of i-mer and (i + l)-mer must be nearly equal (recall\nEquation 3.7.3). The reason we have not emphasized the possibility of equilibrium so far is that\nalmost all polymerization reactions are run under conditions where the equilibrium lies far to the\nright, in favor of products; the residual monomer concentration is very small. This is not always the\n"]], ["block_5", ["where the reaction quotient, (21,015,, is the same ratio of product and reactant concentrations as K,\nbut not necessarily at equilibrium. The free energy of polymerization is the difference between the\nfree energies of the products and the reactants, in kJ/mol, where for polymeric species we consider\nmoles of repeat units. The superscript 0 indicates the standard quantity, where all species are at\nsome specified standard state (e.g., pure monomer and repeat unit, or perhaps at\n1 mol L\u20141\nin solution). For the polymerization reaction to proceed spontaneously, AGpoly <0. When the\n"]], ["block_6", ["Up to this point we have tended to write chain-growth propagation steps as one-way reactions, with\na single arrow pointing to the product:\n"]], ["block_7", ["case, however, as we shall now discuss.\nThe state of equilibrium is directly related to the Gibbs free energy of polymerization:\n"]], ["block_8", ["In fact, as a chemical reaction, there must be a reverse depropagation or depolymert'zation step, and\nthe possibility of chemical equilibrium:\n"]], ["block_9", ["This equilibrium constant for polymerization, Kpoly, can be written as the ratio of the forward and\nreverse rate constants, and as the appropriate ratio of species concentrations at equilibrium:\n"]], ["block_10", ["4.7\nPolymerization Equilibrium\n"]], ["block_11", ["Polymerization Equilibrium\n147\n"]], ["block_12", ["K\n: =\n[PEI-(+1]\ng\n1\nW\u201d\nk...\n[Pr][M1\n[M].q\n(4.7.1)\n"]], ["block_13", ["It\nPik+M\u2014p>P?c+l\n"]], ["block_14", ["now], A0301), + RT 1n QM\n(4.7.2)\n"]], ["block_15", ["Kpolr\nPik\u2018l'Mpl-l\n"]]], "page_162": [["block_0", [{"image_0": "162_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "162_1.png", "coords": [20, 497, 412, 650], "fig_type": "figure"}]], ["block_2", ["Note: The enthalpy corresponds to the conversion of liquid monomer (gas1n the case of ethylene) to amorphous\n(or slightly crystalline) polymer. The entropy corresponds to conversion of a 1 mol L\u2018I solution of monomer to\npolymer.\n"]], ["block_3", ["reaction is allowed to come to equilibrium, Q1301), \npoly \nknown relation\n"]], ["block_4", ["The free energy change per repeat unit upon polymerization may be further resolved into enthalpic\n(H) and entropic (S) contributions:\n"]], ["block_5", ["Ethylene\n\u2014\u201493\n\u2014 155\n\u201447\nPropylene\n\u201484\n\u20141 16\n\u201449\nIsobutylene\n\u201448\n\u2014121\n\u201412\n1,3-Butadiene\n\u201473\n\u201489\n\u201446\nIsoprene\n\u201475\n\u20141 01\n\u201445\nStyrene\n\u201473\n\u2014104\n\u201442\na-Methylstyrene\n\u201435\n\u20141 10\n\u20142\nTetrafluoroethylene\n-\u2014163\n\u20141 12\n\u2014130\nVinyl acetate\n\u201488\n-\u20141 10\n\u201455\nMethyl methacrylate\n\u201456\n\u20141 17\n\u201421\n"]], ["block_6", ["From Equation 47.3, we can see that the statement that Kpoly is large, favoring products, is\nequivalent to saying that AGpoly18 large and negative. From Equation 47.4, we can see that facile\npolymerization requires that either A1130\u201d13 large and negative, thatIS, the reaction is exothermic,\nor that ASpol), is large and positive. In fact, A5130]y is usually negative, the monomers lose\ntranslational entropy when bonded together1n a polymer. However,\npoly \nthe extra energy of a carbon\u2014carbon double bond relative to a single bond is released. In fact, we\nshould have anticipated this conclusion from the outset: polymers could not be made inexpensively\nin large quantities if we had to put in energy for each propagation step.\nTable 4.4 provides examples of the standard enthalpy and entropy of polymerization for a few\ncommon vinyl monomers. In all cases both the enthalpy and the entropy changes are negative, as\nexpected; furthermore, AGED\u201d, is negative at room temperature (300 K). Starting with ethylene as\nthe reference, the relative enthalpies of polymerization can be understood in terms of two general\neffects. The first is the possibility of resonance stabilization of the double bond in the monomer\nthat is lost upon polymerization. This results in lower exothermicity for butadiene, isoprene,\nstyrene, and a-methylstyrene, for example. The second is steric hindrance in the resulting polymer.\nFor example, disubstituted carbons in the polymer can lead to signi\ufb01cant interactions between\nsubstituents on every other carbon, which therefore destabilize the polymer, as in the case of\nisobutylene, or-methylstyrene, and methyl methacrylate. Tetra\ufb02uoroethylene, with its unusually\nlarge exothennicity, is included in this short table in part to remind us that there are examples\nwhere we may not have a simple explanation. Further discussion of these issues is provided in\nSection 5.4.\nEquation 4.7.4 indicates that as the polymerization temperature increases, the relative import-\nance of entropy increases as well. As AS favors depolymerization, it is possible to reach a\n"]], ["block_7", ["Table 4.4\nValues of the Standard Enthalpy and Entropy of Polymerization, as Reported in\nOdian [6]\n"]], ["block_8", ["Monomer\nAHgoly (kJ mol-l)\nA5130;y (J K Hmol\u2018 )\nAGpoly at 300 K (kJ mol\")\n"]], ["block_9", ["148\nControlled Polymerization\n"]], ["block_10", ["AGO\nRTln K1301),\n(4.7.3)\npoly:\n"]], ["block_11", ["AGO\n= M301, T453013,\n(4.7.4)\nPoly\n"]]], "page_163": [["block_0", [{"image_0": "163_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "163_1.png", "coords": [30, 85, 189, 114], "fig_type": "molecule"}]], ["block_2", ["Using this relation, the data in Table 4.4, and assuming [M] 1 mol L\u2018l, the ceiling temperature is\n45\u00b0C for poly(0t-methylstyrene) and 206\u00b0C for poly(methyl methacrylate). Note the important fact\nthat according to Equation willdepend on the monomer concentration and will therefore\nbe different for a polymerization in dilute solution compared to one in bulk monomer.\nFigure 4.6 illustrates the application of Equation 4.7.6 to poly(a\u2014methylstyrene). The equilib-\nrium monomer concentration is plotted against temperature according to Equation 4.7.6 and\nthe indicated values of AH0 and A30. The smooth curve corresponds to the ceiling temperature\nand the horizontal line indicates a 1 mol L\u20181 solution. Any solution of monomer that falls to the\nright of, or below, the curve (i.e., any combination of [M] and T) will simply not polymerize. Any\nsolution to the left of, or above, the curve can polymerize, but only until the equilibrium monomer\nconcentration is reached. For example, a 1 mol L\u20141 solution at 0\u00b0C could polymerize, but as\n"]], ["block_3", ["the equilibrium monomer concentration is about 0.1 mol LT], the maximum conversion would\nonly be 90%.\nInterestingly, there are a few instances in which polymerization is driven by an increase in\nentropy, and where the enthalpy gain is almost negligible. Examples include the polymerization of\ncyclic oligomers of dimethylsiloxane, such as the cyclic trimer and tetramer, which we will discuss\nin the next section. In this case, the bonds that are broken and reformed are essentially the same,\n"]], ["block_4", ["and combining 4.7.1 through Equation 4.7.3 \ufb01nd\n"]], ["block_5", ["temperature \nspecial temperature is to the ceiling temperature, TC. From have\n"]], ["block_6", ["Figure 4.6\nIllustration of the relation between equilibrium monomer concentration and ceiling temperature\nfor poly(a-methylstyrene).\n"]], ["block_7", ["[\n]\n0.1-\n-:\n_ a\u2014methylstyrene\n"]], ["block_8", ["polymerization Equilibrium\n149\n"]], ["block_9", [{"image_2": "163_2.png", "coords": [41, 476, 284, 608], "fig_type": "figure"}]], ["block_10", ["0.001 \n-50\n0\n50\n100\n150\n"]], ["block_11", ["AHOpoly\n, \n4.7.6\nT\nA8301),+1\u2019i\u2018ln[M],,,l\n(\n)\n"]], ["block_12", ["AGO\npoly \npoly TCASO\n(4-7-5)\nPoly\n"]], ["block_13", ["001:\nA802\u2014110Jmol_1K_1\n2\n"]], ["block_14", [{"image_3": "163_3.png", "coords": [59, 367, 278, 634], "fig_type": "figure"}]], ["block_15", [{"image_4": "163_4.png", "coords": [129, 509, 263, 567], "fig_type": "figure"}]], ["block_16", ["T,\u00b0C\n"]], ["block_17", ["AH\u201d \u201435,000 J mol\"1\n"]]], "page_164": [["block_0", [{"image_0": "164_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The following values of We and A58; per methylene unit have been estimated at room tempera-\nture (as reported in Ref. [10]) for the process\n"]], ["block_2", ["The subscript \u201c1c\u201d denotes the liquid-to\u2014crystal aspect of the process; as we will see in Chapter 13,\nhigh molecular weight linear polyethylene, the product of the hypothetical polymerization\n"]], ["block_3", ["In this instance the ring contains four atoms, and the A\u2014B bond is the one that is preferentially\ncleaved. The propagating chain is shown with B containing the active center; it is often the case\nthat ROP proceeds by an ionic mechanism. Comparison of monomer and repeat unit structures in\nReaction (4.DD) reveals that the bonding sequence is the same in both cases, in marked contrast to\neither a chain-growth polymerization through, for example, a carbon\u2014carbon double bond or a\nstep-growth polymerization through, for example, condensation of acid and alcohol groups. In\nlight of the previous section, where we considered the thermodynamics of polymerization, a basic\nquestion immediately arises: if the bonding is the same in monomer and polymer, what is the\nprimary driving force for polymerization? The answer is ring strain. The linkage of the atoms into\na ring generally enforces distortion of the preferred bond angles and even bond lengths, effects that\nare grouped together under the title ring strain. The amount of ring strain is a strong function of the\nnumber of atoms in the ring, r. For example, ethylene oxide, with r 3, is quite a explosive gas at\nroom temperature. On the other hand, cyclohexane, with r=6, is almost inert. In fact, r=6\nrepresents a special case, at least for all carbon rings, as the natural sp3 bond angles and lengths\ncan be almost perfectly matched. The following example illustrates the effect of r on the\nthermodynamics of polymerization, for cyclic alkanes.\n"]], ["block_4", ["4.8\nRing-Opening Polymerization (ROP)\n"]], ["block_5", ["4.8.1\nGeneral Aspects\n"]], ["block_6", ["Cyclic molecules, in which the ring contains a modest number of atoms, say 3\u20148, can often be\npolymerized by a ring-opening reaction, in which a particular bond in the cycle is ruptured, and\nthen reformed between two different monomers in a linear sequence. This process is illustrated in\nthe following schematic reaction:\n"]], ["block_7", ["hence M301), :3 0. On the other hand, possibly because of the greater conformational freedom in\nthe linear polymer versus the small cycles, A5301), is positive. In such a case when M301), is slightly\npositive, it is possible in principle to have a\ufb02oor temperature, below which polymerization cannot\noccur at equilibrium. Furthermore, for. the living polymerization of \u201cD3,\u201d the cyclic trimer of\ndimethylsiloxane, for example using an anionic initiator, 111G130]y is never particularly favorable\ncompared to the typical vinyl monomer case. Consequently, one has to be aware of the law of mass\naction, just as in a polycondensation. In other words, the reaction cannot be allowed to go to\ncompletion, because rather than achieving essentially 100% conversion to polymer, the reaction\nmixture stabilizes at the equilibrium monomer concentration. Then, random propagation and\ndepropagation steps will degrade the narrow molecular weight distribution that was initially\nsought. This problem can be circumvented by adding more monomer than is necessary to achieve\nthe target molecular weight, and using trial and error to determine the time (and fractional\nconversion) at which the desired average molecular weight has been achieved.\n"]], ["block_8", ["Example 4.5\n"]], ["block_9", ["150\nControlled Polymerization\n"]], ["block_10", [{"image_1": "164_1.png", "coords": [39, 326, 184, 358], "fig_type": "molecule"}]], ["block_11", ["MB*+pl:||3\u2014\u2014\u2014p\u2014\nMB\u2014A\n3*\n(4.DD)\n"]], ["block_12", ["\u2014(CH2), (liquid ring)\n41>\n\u2014\u2014(CH2),, (crystalline linear polymer)\n"]]], "page_165": [["block_0", [{"image_0": "165_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "165_1.png", "coords": [19, 93, 204, 190], "fig_type": "figure"}]], ["block_2", [{"image_2": "165_2.png", "coords": [26, 239, 175, 346], "fig_type": "figure"}]], ["block_3", ["The preceding example nicely illustrates the importance of ring strain, but the fact is that the\nprimary utility of ROP is not to produce polyethylene from cyclic alkanes. Rather, it is to produce\ninteresting polymer structures from readily accessible cyclic monomers, structures that cannot be\nprepared more conveniently by classical step\u2014growth or chain-growth polymerization. Examples of\nseven different classes of cyclic monomers and the resulting polymer structures are given in Table\n4.5. In all cases the ring contains one or more heteroatoms, such as O, N, and Si. These participate\nin the bond-breaking process that is essential to ROP; in contrast, it is actually rather difficult in\npractice to polymerize cyclic alkanes, even when free energy considerations favor it. In Chapter 1\nwe suggested that the presence of a heteroatom in the backbone was often characteristic of a step\u2014\ngrowth polymerization. One of the beauties of ROP is that it is a chain-growth mechanism,\nenabling the ready preparation of high molecular weight materials. For example, Entry 5 (a\npolyamide, poly(s\u2014caprolactam)) and Entry 6 (a polyester, polylactide) are polymers that could\nbe prepared by condensation of the appropriate AB monomer. However, by using the cyclic\nmonomer, the condensation step has already taken place, and the small molecule byproduct is\n"]], ["block_4", ["7\u2018, while remaining consistently negative. However, we need to recognize that these values\nincorporate the relief of the ring strain, the incorporation of monomer into polymer, and the\nchanges associated with crystallization of the liquid polymer. The opening of the ring affords\nmore degrees of freedom to the molecule, increasing the entropy, but both subsequent polymer-\nization and crystallization reduce it. Consequently, it is dangerous to overinterpret particular\nvalues of A80.\n"]], ["block_5", ["The results indicate that for all but r: 6, polymerization is favored, consistent with the known\nstability of six\u2014membered carbon rings. From the point of view of polymerization, the driving\nforce should be ranked according to r:3, 4 >r= 8 >r=5, 7. These are general trends, and\ndifferent substituents or heteroatoms within the ring can change the numerical values signifi-\ncantly. Finally, while AHO shows a distinct maximum at r 6, A50 decreases monotonically with\n"]], ["block_6", ["We use the relationships AG\u00b0=AH\u00b0\u2014TASO (Equation 4.7.4) and AGO: \u2014RT In K (Equation\n4.7.3), with T:298 K and R 8.314 J moi\u20141 K\u2018l, to obtain the following table:\n"]], ["block_7", ["3\n\u201492\n2 x 1016\n4\n\u201489\n3 x 1015\n5\n\u20148.5\n30\n6\n+ 6.0\n0.09\n7\n\u201417\n103\n8\n\u201434\n8 x 105\n"]], ["block_8", ["_._._\u2014\u2014\u2014\nr\nA119, (kJ mol\u201c)\nASE, (J nior1 K\u201c)\n"]], ["block_9", ["3\n\u20141 13\n\u201469.1\n4\n4105\n\u201455.3\n5\n\u201421.2\n\u201442.7\n6\n+ 2.9\n\u201410.5\n7\n\u201421.8\n\u201415.9\ng\n\u201c34.8\n\u20143.3\n"]], ["block_10", ["reaction, is crystalline at equilibrium at room temperature. Evaluate AGPC and K for this process\nand interpret \n"]], ["block_11", ["Solution\n"]], ["block_12", ["r\nnot: (H moi\u20141)\nK\n"]], ["block_13", ["Ring-Opening \n151\n"]]], "page_166": [["block_0", [{"image_0": "166_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "166_1.png", "coords": [32, 76, 267, 327], "fig_type": "figure"}]], ["block_2", ["4.8.2\nSpecific Examples of Living Ring-Opening Polymerizations\n"]], ["block_3", ["4.8.2.1\nP01y(ethy1ene oxide)\n"]], ["block_4", ["removed. Thus, the law of mass action that typically limits the molecular weight of a step-growth\npolymerization is overcome. (This is not to say that polymerization equilibrium will not be an\nissue. In fact, Entry 6 and Entry 7 both indicate six-membered rings, and the correspondingly lower\nring-strain does bring equilibration into play.)\nFrom the perspective of this chapter the main point of ROP is not just its chain-growth\ncharacter, but the fact that in many cases living ROP systems have been designed. As indicated\nabove, most ROPs proceed via an ionic mechanism, which certainly invites attempts to achieve a\nliving polymerization. We will brie\ufb02y present specific examples of such systems, for three\ndisparate but rather interesting and important polymers in Table 4.5: poly(ethy1ene oxide) (Entry\n1), polylactide (Entry 6), and poly(dimethylsiloxane) (Entry 7). We will also consider a class of\nROP that can produce all carbon backbones, via olefin metathesis.\n"]], ["block_5", ["P01y(ethy1ene oxide) represents one of the most versatile polymer structures for both fundamental\nstudies and in commercial applications. It is readily prepared by living anionic polymerization,\nwith molecular weights ranging all the way up to several millions. It is water soluble, a highly\ndesirable yet relatively unusual characteristic of nonionic, controlled molecular weight polymers.\nFurthermore, it appears to be more or less benign in humans, thereby allowing its use in many\n"]], ["block_6", ["Table 4.5\nExamples of Monomers Amenable\nto Ring-Opening Polymerization\n"]], ["block_7", ["6. Lactones (cyclic esters)\n\u2018z\n"]], ["block_8", ["/ /\n,O-Sk\nMe\nMe\n7. Siloxanes\n:Si\n0\n()Si: ,)\nb\u2014Si:\n0\n"]], ["block_9", ["1 52\nControlled Polymerization\n"]], ["block_10", [".\n.\n.\nT7\nH\n4. Immes (cyclic amines)\nN\nN\nH\nW \\l\n"]], ["block_11", ["2. Cyclic ethers\n[:0\nM04\n"]], ["block_12", ["5. Lactams (cyclic amides)\nGH\n(\u201895\nE4\n"]], ["block_13", ["Monomer class\nExample\nRepeat unit\n"]], ["block_14", ["/\\\n.\n0\n0\n3. Cyclic acetals\nOMS\n(/O\\/ p?\n"]], ["block_15", ["W0\u00bb)\n[:0\n1. Epoxides\n"]], ["block_16", [{"image_2": "166_2.png", "coords": [134, 354, 190, 404], "fig_type": "molecule"}]], ["block_17", [{"image_3": "166_3.png", "coords": [143, 298, 265, 351], "fig_type": "molecule"}]], ["block_18", ["0\no\no\n0\n"]], ["block_19", ["(x w\n0\n0\nMe\n"]], ["block_20", [{"image_4": "166_4.png", "coords": [221, 301, 263, 346], "fig_type": "molecule"}]]], "page_167": [["block_0", [{"image_0": "167_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["is not an effective counter ion in this case. (As noted in the context of Reaction 4.H, this feature is\nactually convenient when it is desired to use ethylene oxide to end-functionalize such polymers.)\nThe following scheme represents an example of a three-arm star prepared by grafting from; the\ninitiator is trimethylol propane, which has three equivalent primary alcohols that can be activated\nby diphenylmethyl potassium. of ethylene oxide monomer is straightforward, and the\nresulting polymer is terminated with acidic methanol to yield a terminal hydroxyl-functionality on\neach arm.\n"]], ["block_2", ["4.8.2.2\nPolylactide\n"]], ["block_3", ["Polylactide is biodegradable and can be derived from biorenewable feedstocks such as corn. The\nformer feature enables a long-standing application as resorbable sutures; after a period of days to\nweeks, the suture degrades and is metabolized by the body. The latter property underscores recent\ninterest in the large-scale commercial production of polylactide for a wide variety of thermoplastic\napplications.\nThe structure of lactic acid is\n"]], ["block_4", ["Clearly, as it contains both a hydroxyl and a carboxylic acid, it could be polymerized directly to the\ncorresponding polyester via condensation. However, if the starting material is lactide, the cyclic\n"]], ["block_5", ["ene oxide that example in proposing the Poisson distribution for chain-growth\npolymers prepared from a \ufb01xed number of initiators with rapid propagation [14].\nIn practice, the living anionic polymerization of ethylene oxide has been achieved by a variety\nof initiator systems, following the general principles laid out in Section 4.3. Examples include\nmetal hydroxides, alkoxides, alkyls, and aryls. In contrast to styrenes and dienes, however, lithium\n"]], ["block_6", ["consumer products, biomedical formulations, etc. In fact, a grafted layer of short chain poly(ethyl-\nene oxide)s (see confer long-term stability against protein adsorption or deposition\nof other biochemical arena short chain poly(ethylene oxide)s (with\nhydroxyl groups at both ends) are more commonly referred to as poly(ethylene glycol)s or PEGs.\nThe grafting of PEG molecules onto a biomacromolecule or other substrate has become such a\nuseful name: PEGylation. Poly(ethylene oxide) crystallizes\nrather readily, witha typical melting temperature near 65\u00b0C, which has led to its use in many\nfundamental crystallization (see Chapter 13). The first block copolymers\nto become the so-called polyoxamers, diblocks and triblocks of\npoly(ethylene oxide) and poly(propylene oxide). Historically, it was the polymerization of ethyl-\n"]], ["block_7", ["a cationic ring-opening mechanism; the high ring-strain of ethylene oxide makes it the exception to\nthis rule.\n"]], ["block_8", ["An interesting feature of this reaction is that it can be carried out in THF, which like ethylene oxide\nis a cyclic ether. This again illustrates the importance of ring-strain in facilitating, or, in this case\nsuppressing, polymerization. In fact, cyclic ethers including THF are usually polymerized only by\n"]], ["block_9", ["Ring-Opening Polymerization (ROP)\n153\n"]], ["block_10", [{"image_1": "167_1.png", "coords": [43, 336, 282, 398], "fig_type": "figure"}]], ["block_11", [{"image_2": "167_2.png", "coords": [44, 335, 293, 401], "fig_type": "molecule"}]], ["block_12", ["Me\nH\nHO\nO\n"]], ["block_13", ["Me/q(\\OH\n"]], ["block_14", [{"image_3": "167_3.png", "coords": [48, 592, 98, 634], "fig_type": "molecule"}]], ["block_15", [{"image_4": "167_4.png", "coords": [49, 340, 205, 393], "fig_type": "molecule"}]], ["block_16", ["K+\nO\nK+ 01\nMeOH\nH0\nK+ _\n"]], ["block_17", ["0\n"]], ["block_18", ["OH\nK+ \n(4.EE)\n"]], ["block_19", ["+\n3\n\u2014p-\nMe\nO\u2014 \n-\u2014-\u2014-\u2014\u2014\u2014\u2014-i-- Fl((CH2CH20)nH)3\n"]], ["block_20", [{"image_5": "167_5.png", "coords": [197, 338, 428, 395], "fig_type": "molecule"}]], ["block_21", ["\u2014\n"]]], "page_168": [["block_0", [{"image_0": "168_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "168_1.png", "coords": [32, 377, 280, 457], "fig_type": "figure"}]], ["block_2", ["4.8.2.3\nPoly(dimethylsiloxane)\n"]], ["block_3", ["Note that the central carbon in lactic acid is chiral, and therefore the corresponding two carbons in\nboth lactide and the polylactide repeat unit are stereocenters. Consequently, there are three possi-\nbilities for the lactide monomer, according to the absolute configuration of these carbons: R, R; S, S;\nand R,S. Polymerization of either of the first two leads to the corresponding stereochemically pure\nPDLA and PLLA, which are crystallizable; the meso dyad leads to an atactic polymer. In this case,\ntherefore, the responsibility of producing a particular tacticity is largely transferred from the catalyst\n(see Chapter 5) to the purification of the starting material. Of course, the catalyst has to guide the\npolymerization in such a way that the stereochemistry is not scrambled or epimerized.\nFrom the point of view of designing a living polymerization of polylactide, there are two general\nissues to confront. First, as polylactide is a polyester, it is susceptible to transesterification reactions.\nThis constraint favors lower reaction temperatures, conditions that are neither too basic nor acidic,\nand acts against the normal desire to make the catalyst more \u201cactive.\u201d Ironically, the advantageous\ndegradability of polylactide through the hydrolysis of the ester linkage is thus a disadvantage from\nthe point of view of molecular weight control. The second problem is equilibration. Recall from\nExample 4.5 that for cyclic alkanes, the six-membered ring has no ring~strain to speak of, and is\ntherefore not polymerizable. Although lactide is a six-membered ring, it does possess sufficient ring-\nstrain, but not a lot. Consequently, narrow molecular weight distributions are usually obtained only\nby terminating well before the reaction has approached completion.\nA typical catalyst is based on a metal alkoxide, such as LMOR\u2019, where L is a \u201cspectator\u201d ligand\nand R\u2019 is a small alkyl group. The initiation step can be written as\n"]], ["block_4", ["where the lactide ring is cleaved at the bond between the oxygen and the carbonyl carbon. The\nsubsequent propagation steps involve the same bond cleavage, with addition of the new monomer\ninto the oxygen\u2014metal bond at the growing chain end. In fact, this kind of polymerization has been\nclassified as \u201canionic coordination,\u201d in distinction to anionic polymerization, as the crucial step is\ncoordination of the metal with the carbonyl oxygen, followed by insertion of the alkoxide into the\npolarized C\u2014O bond. The most commonly employed catalyst for polylactide is tin ethylhexanoate,\nbut more success in terms of achieving living conditions has been realized with aluminum\nalkoxides. Interestingly, these aluminum species have a tendency to aggregate in solution, with\nthe result that the reaction kinetics can become rather complicated; different aggregation states can\nexhibit very different pr0pagation rates. This situation is reminiscent of the aggregation of\ncarbanions in anionic polymerization in nonpolar solvents discussed in Section 4.3.\n"]], ["block_5", ["Poly(dimethylsiloxane) has one of the lowest glass transition temperatures (Tg, see Chapter 12) of\nall common polymers. This is due in part to the great \ufb02exibility of the backbone structure (see\nChapter 6), which re\ufb02ects the longer Si\u2014O bond compared to the C\u2014C bond, the larger bond angle, and\nthe absence of substituents on every other backbone atom. It is also chemically quite robust. It is used if.\n"]], ["block_6", ["dimer of the corresponding ester, then ROP produces the same polymer structure but by a Chain-\ngrowth process:\n"]], ["block_7", ["154\nControlled Polymerization\n"]], ["block_8", [{"image_2": "168_2.png", "coords": [36, 75, 226, 141], "fig_type": "figure"}]], ["block_9", [{"image_3": "168_3.png", "coords": [38, 383, 245, 450], "fig_type": "molecule"}]], ["block_10", [{"image_4": "168_4.png", "coords": [43, 83, 222, 127], "fig_type": "molecule"}]], ["block_11", ["I'_M\nMe\nO\n0\nMe\nOM\nI I\n+LMOR'\nl\ne\no\n(4-GG)\no\n0\nMe\n0\nOJY\non'\n"]], ["block_12", ["Me\n0\no\n0\nMe\nI f\n..\n0%0\n(4.1212)\n0\n0\nMe\nMe\n0\nn\n"]]], "page_169": [["block_0", [{"image_0": "169_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "169_1.png", "coords": [28, 188, 265, 243], "fig_type": "molecule"}]], ["block_2", ["Although Reaction (4.11) illustrates the net outcome, it says nothing about the mechanism.\nMetathesis reactions are catalyzed by transition metal centers, with the associated ligand package\nproviding tunability of reaction characteristics such as rate, selectivity, and stereochemistry of\naddition. In the case of ROMP, the metal forms a double bond with a carbon at one end of the\nchain; thus in Reaction (4.11) the RICH group would be replaced by the metal and its ligands\n(MLn). For a propagation step, R2 would denote a previously polymerized chain, P,. In the\nmonomer to be inserted into the chain, between the metal and the end of the previously polymer-\nized chain, R3 and R4 are covalently linked to form a ring. Thus, the ring plays two key roles:\nthe ring-strain provides the driving force for polymerization, and the ring structure provides the\npermanent connectivity between the two carbon atoms whose double bond is broken. The ROMP\nanalog to Reaction (4.11) can thus be described schematically as\n"]], ["block_3", ["any number of lubrication and adhesive applications (\u201csilicones\u201d), as well as a variety of rubber\nmaterials. of the cyclic hexamethylcyclotrisiloxane, D3, has been\nachieved that this monomer is a six\u2014membered ring, we can anticipate that\nthe polymerization is not strongly favored by thermodynamics. Consequently, narrow molecular weight\ndistributions are achieved by terminating the reaction well before the consumption of all the monomer.\nAn example of a successful protocol is the following. A modest amount of cyclic trimer is\ninitiated solution. Under these conditions initiation is rather\nslow, but propagation is almost thereby allowing for complete initiation. Presumably,\nthe lack of propagation is due to ion clustering as discussed in Section 4.3. The addition of THF, as a\npolar modifier, plus more monomer allows propagation to proceed for an empirically determined\ntime interval. Termination achieved with trimethylchlorosilane (TMSCl):\n/\nO\u2014Si\nO\nO\nO\u2014\nCHX\n+\n:SI/\n\\O+FlOK\u2014I-\u2014 Ro/>SiR1 \n--.\u2014\u2014\n(4.11)\n"]], ["block_12", [{"image_4": "169_4.png", "coords": [48, 391, 213, 463], "fig_type": "figure"}]], ["block_13", ["H\nH\n"]], ["block_14", [{"image_5": "169_5.png", "coords": [81, 248, 286, 293], "fig_type": "molecule"}]], ["block_15", ["THF\nD3\nTMSCI\nO\n"]], ["block_16", ["\\\nn\n[\n"]]], "page_170": [["block_0", [{"image_0": "170_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "170_1.png", "coords": [25, 495, 217, 562], "fig_type": "figure"}]], ["block_2", [{"image_2": "170_2.png", "coords": [33, 490, 374, 571], "fig_type": "figure"}]], ["block_3", ["Dendrimers are an interesting, unique class of polymers with controlled structures. For example,\nthey can have precisely defined molecular weights, even though the elementary addition steps are\nusually of the condensation variety. From an application point of view it is the structure of the\ndendrimer, rather than its molecular weight per se that is the source of its appeal. A cartoon\nexample of a dendrimer is provided in Figure 1.2. The term comes from the Greek word dendron,\nor tree, and indeed a dendrimer is a highly branched polymer molecule. In particular, a dendrimer\nis usually an approximately spherical molecule with a radius of a few nanometers. Thus, a\n"]], ["block_4", ["Note that the substituent on the metal\u2014carbon double bond will become attached to the nonpropa-\ngating terminus of the chain. Collectively, catalysts in these two families have proven capable of\nachieving controlled polymerization of a wide variety of cyclic olefins, including those containing\nfunctional groups. In particular, while the Schrock catalysts tend to be more active, the Grubbs\ncatalysts are more tolerant of functional groups, oxygen, and protic solvents. Reaction conditions\nare often mild, that is, near room temperature, and in some cases the polymerization can be\nconducted in water. Overall control is often quite good, and many block copolymers have\nbeen prepared by ROMP. Some ROMP systems have even been commercialized, including the\npolymerization of norbornene:\n"]], ["block_5", ["4.9\nDendrimers\n"]], ["block_6", ["In this case, the monomer is cyclooctene. After the monomer insertion or propagation step, an\nactive metal carbene remains at the chain terminus, and one carbon\u2014carbon double bond remains in\nthe backbone for each repeat unit.\nThere is a large literature on metathesis catalysts and associated mechanisms, many of\nwhich incorporate multiple components beyond the active metal. However, for the purposes of\ncontrolled ROMP, there are currently two families of single species catalysts that are highly\nsuccessful. One, based on tungsten or molybdenum, is known as a Schrock catalyst [I], and the\nother, based on ruthenium, is a Grubbs catalyst [II]. These investigators were corecipients of the\n2005 Nobel Prize in chemistry (with Y. Chauvin) for their work on the metathesis reaction.\nThe structures of representative examples are given below, where the symbol Cy denotes a\ncyclohexyl ring:\n"]], ["block_7", ["1 56\nControlled Polymerization\n"]], ["block_8", [{"image_3": "170_3.png", "coords": [41, 189, 136, 267], "fig_type": "molecule"}]], ["block_9", [{"image_4": "170_4.png", "coords": [42, 498, 213, 563], "fig_type": "molecule"}]], ["block_10", [{"image_5": "170_5.png", "coords": [44, 289, 127, 364], "fig_type": "molecule"}]], ["block_11", ["N\\\n0\nemail\n"]], ["block_12", ["CI\" I?\u201d\nPCY3\n"]], ["block_13", ["Grubbs catalyst\n"]], ["block_14", ["Ins _\n[II]\n"]], ["block_15", ["\u2014...W\n(4.KK)\n"]]], "page_171": [["block_0", [{"image_0": "171_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "171_1.png", "coords": [15, 465, 235, 540], "fig_type": "molecule"}]], ["block_2", [{"image_2": "171_2.png", "coords": [30, 551, 248, 632], "fig_type": "molecule"}]], ["block_3", [{"image_3": "171_3.png", "coords": [32, 409, 211, 443], "fig_type": "molecule"}]], ["block_4", ["dendrimer and also a well-defined nanoparticle. The outer\nsurface density of functional groups that govern the\ninteractions These exterior groups have the advan-\ntages of being numerous, and readily accessible for chemical transformation. The interior of the\ndendrimer can distinct functionality that can endow the molecule with desirable\nproperties. For example, the dendrimer might incorporate a highly absorbing group, for \u201clight\nharvesting,\u201d or \ufb02uor0phore, for ef\ufb01cient emission. Other possibilities include catalytic centers or\nelectrochemically active groups. By being housed within the dendrimer, this functional unit can be\nprotected from unwanted interactions with the environment. The functional unit may be covalently\nbound within the or it may simply be encapsulated. The possibility of controlling\nuptake and release of specific agents by the dendrimer core also makes them appealing as possible\ndelivery vehicles for pharmaceuticals or other therapeutic agents. As nanoparticles, dendrimers\nshare certain attributes with other objects of similar size, such as globular proteins, surfactant and\nblock copolymer micelles, hyperbranched polymers, and colloidal nanoparticles. Although beyond\nthe scope of this chapter, it is interesting to speculate on the possible advantages and disadvantages\nof these various structures (see Problem 16).\nThere are two distinct, primary synthetic routes to prepare a dendrimer, termed divergent and\nconvergent. In a divergent approach, the dendrimer is built-up by successive additions of mono-\nmers to a central, branched core unit, whereas in the convergent approach branched structures\ncalled dendrons are built-up separately, and then ultimately linked together to form the dendrimer\nin a final step. The divergent approach was conceived first, and is the more easily visualized. The\nprocess is illustrated schematically in Figure 4.7. The core molecule in this case has three\nfunctional groups denoted by the open circles. These are reacted with three equivalents of another\nthree-functional monomer, but in this case two of the functional groups are protected (filled\ncircles). After this reaction is complete, the growing molecule has six functional groups that are\n"]], ["block_5", ["Figure 4.7\nSchematic illustration of the divergent synthesis of a third-generation dendrimer from a\ntrifunctional core. The open circles denote reactive groups and the \ufb01lled circles protected groups.\n"]], ["block_6", ["Dendrimers\n1 57\n"]], ["block_7", ["G2\n+12\n"]], ["block_8", [{"image_4": "171_4.png", "coords": [47, 389, 230, 462], "fig_type": "figure"}]], ["block_9", [{"image_5": "171_5.png", "coords": [52, 395, 208, 458], "fig_type": "molecule"}]], ["block_10", ["A. J/ \n"]], ["block_11", ["J, Deprotect\n"]], ["block_12", [{"image_6": "171_6.png", "coords": [95, 473, 229, 530], "fig_type": "molecule"}]], ["block_13", ["\\k\ufb01%\\<\n"]], ["block_14", ["Deprotect\n\\\n"]]], "page_172": [["block_0", [{"image_0": "172_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Thus, a perfect G5, G6, and G7 dendrimer would have 96, 192, and 384 functional groups,\nrespectively. Note that Equation 4.9.1 would need to be modified in the case of, for example, a\ntetrafunctional core. We have introduced the term \u201cperfect\u201d here to emphasize that it is certainly\npossible for a dendrimer molecule to have defects, or missing functional groups, which will\npropagate through all subsequent generations. For the first few generations it is usually not too\ndifficult to approach perfection, but for G5 and above, the functional groups become rather\ncongested, which makes complete addition of the next generation difficult. It also becomes harder\nto separate out defective structures. It is typically not practical to go beyond G8.\nFurther consideration of the divergent approach in Figure 4.7 reveals that, compared with most\npolymerization reactions, it is rather labor intensive. For example, the addition of each generation\nrequires both an addition step and a deprotection step. The addition will typically be conducted\nin the presence of a substantial excess of protected monomer, to drive the completion of the\nnew layer. The resulting products will need to be separated, in order to isolate the perfect\ndendrimer structure from all other reaction products and reagents. Similarly, the deprotection\nstep needs to be driven to completion, and the pure product isolated. Thus, in the end there are two\nreaction steps and two purification steps required for each generation. This requires a significant\namount of time, and it is challenging to prepare commercial-scale quantities of perfect, high\ngeneration products.\nAs a specific example of a divergent synthesis, we will consider the formation of the poly-\namidoamine (PAMAM) system. In this case, there are two monomers to be added sequentially in\neach generation, rather than one addition and one deprotection step. The core molecule and one of\nthe monomers is typically ethylene diamine and the other monomer is methyl acrylate. The first\nstep is addition of four methyl acrylate molecules to ethylene diamine in a solvent such\nas methanol. The Michael addition\u2014type mechanism involves nucleophilic attack of the electron\npair\non\nthe\nnitrogen\nto\nthe\ndouble\nbond\nof\nthe\nacrylate,\nwhich\nis\nactivated\nby\nthe\nelectron withdrawing character of the ester group:\n"]], ["block_2", ["then deprotected. At this stage the molecule is termed afirst generation (G1) dendrimer. Another\naddition reaction is then performed, but now six equivalents of the protected monomer are required\nto complete the next generation. After deprotection the resulting G2 dendrimer has 12 functional\ngroups. It is straightforward to see that the number of functional groups on the surface grows\ngeometrically with the number of generations, 3:\n"]], ["block_3", ["The next second step involves amidation of each ester group by nucleophilic attack of the nitrogen\non the electropositive carbonyl carbon, with release of methanol:\n"]], ["block_4", ["158\nControlled Polymerization\n"]], ["block_5", [{"image_1": "172_1.png", "coords": [43, 492, 193, 530], "fig_type": "molecule"}]], ["block_6", [{"image_2": "172_2.png", "coords": [44, 461, 360, 564], "fig_type": "figure"}]], ["block_7", ["Number of functional groups 3 X 23\n(4.9.1)\n"]], ["block_8", ["H\nO\nN\nH2N \ufb02\u201d\\n/\n\\Me\n\u2014\u2014\u2014\u2014\u2014\u20141-- Ir\u201d\\\\u201d/ \\/\\NH2 + MeOH\n(4.MM)\n"]], ["block_9", [{"image_3": "172_3.png", "coords": [66, 607, 370, 662], "fig_type": "molecule"}]], ["block_10", ["O\n"]], ["block_11", ["O\nO\n"]], ["block_12", [{"image_4": "172_4.png", "coords": [179, 461, 322, 523], "fig_type": "molecule"}]], ["block_13", [{"image_5": "172_5.png", "coords": [184, 450, 319, 574], "fig_type": "molecule"}]], ["block_14", ["(DE/Q1219\nMe\\0J1\n"]], ["block_15", [{"image_6": "172_6.png", "coords": [190, 453, 327, 567], "fig_type": "figure"}]], ["block_16", [{"image_7": "172_7.png", "coords": [192, 492, 304, 565], "fig_type": "molecule"}]], ["block_17", ["\u2018i\nM\n"]], ["block_18", ["O\n"]], ["block_19", ["9\n"]], ["block_20", ["f\n1Y0\n"]], ["block_21", ["O\\\n0\nMe\n"]], ["block_22", ["(4.LL)\n"]]], "page_173": [["block_0", [{"image_0": "173_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "173_1.png", "coords": [16, 423, 250, 648], "fig_type": "figure"}]], ["block_2", [{"image_2": "173_2.png", "coords": [21, 432, 186, 471], "fig_type": "molecule"}]], ["block_3", [{"image_3": "173_3.png", "coords": [23, 536, 276, 628], "fig_type": "figure"}]], ["block_4", [{"image_4": "173_4.png", "coords": [26, 473, 221, 534], "fig_type": "molecule"}]], ["block_5", ["An alternative general strategy for preparing dendrimers and dendritic fragments, or dendrons,\nis the so-called convergent approach. This is illustrated schematically in Figure 4.8. As the name\nimplies, these molecules are made \u201cfrom the outside in,\u201d that is, the eventual surface group,\ndenoted by \u201cx\u201d in Figure 4.8, is present in the initial reactants. The first reaction produces a\nmolecule with two surface groups and one protected reactive group. The second step, after\ndeprotection, doubles the number of surface groups, and so on. At any stage, a suitable multifunc-\ntional core molecule can be used, to stitch the appropriate number of dendrons (usually 3 or 4)\ntogether. Each growing wedge-shaped dendron possesses only one reactive group, which pre-\nsents a significant advantage in terms of purification. At each growth step, a dendron either\nreacts or it does not, but the product and reactant are significantly different in molecular weight.\nBy contrast, in the divergent approach, the surface of the dendrimer has many reactive groups,\nand it may not be easy to separate a G3 dendrimer with 24 newly added monomers from one with\nonly 23 monomers. Furthermore, because there are so many more reactive groups on the\ndendrimer than on the added monomers, the monomers must be present in huge molar excess\nto drive each reaction to completion. In the convergent approach shown, there are only twice as\nmany dendrons as new coupling molecules at stoichiometric equivalence, so a large excess of\ndendrons is not necessary.\n"]], ["block_6", ["The structure of the resulting Gl PAMAM dendrimer is therefore the following:\n"]], ["block_7", ["X\u2014O\n2\nX\u2014-O\n+ P \u2014\u2014-\u2014>-\n>\u2014.\nX\u2014o\n"]], ["block_8", ["X\u2014o\n2\n>\u20140\n+ P\u2014p\u2014x\nX\u2014o\nX\u2014o\n"]], ["block_9", ["Figure 4.8\nIllustration of the convergent approach to dendrimer synthesis. Each dendron is built-up by\nsuccessive 2:1 reactions, before the final coupling step.\n"]], ["block_10", ["Dendrimers\n159\n"]], ["block_11", [{"image_5": "173_5.png", "coords": [35, 536, 235, 610], "fig_type": "molecule"}]], ["block_12", [{"image_6": "173_6.png", "coords": [36, 542, 193, 586], "fig_type": "molecule"}]], ["block_13", ["2 \n:>\u2014c\n"]], ["block_14", [{"image_7": "173_7.png", "coords": [37, 68, 238, 205], "fig_type": "figure"}]], ["block_15", ["NH2\nHKEHE\n1\nO\n"]], ["block_16", ["+ P \u2014\"\nX\u2014o\n"]], ["block_17", ["X\u2014\u2014O\n"]], ["block_18", ["X\u2014\u2014o\n"]], ["block_19", [{"image_8": "173_8.png", "coords": [56, 73, 196, 202], "fig_type": "molecule"}]], ["block_20", ["HN\n0\n HM) \nK/NH 2\n"]], ["block_21", ["N/\\/N\nf\n\\kfo\nNH2\n"]], ["block_22", [{"image_9": "173_9.png", "coords": [142, 528, 302, 642], "fig_type": "figure"}]], ["block_23", [")(\u2014\u2014-\u2014-\u00abo\n"]], ["block_24", ["X\n"]], ["block_25", [{"image_10": "173_10.png", "coords": [155, 534, 277, 641], "fig_type": "molecule"}]], ["block_26", ["f?\n1111\n"]]], "page_174": [["block_0", [{"image_0": "174_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "174_1.png", "coords": [33, 141, 430, 254], "fig_type": "figure"}]], ["block_2", ["1.\nWhen a living polymerization is conducted such that the rate of initiation is effectively\ninstantaneous compared to propagation, it is possible to approach a Poisson distribution of\nmolecular weights, where the polydispersity is l +(1/Nn).\n2.\nAnionic polymerization is the most established method for approaching the ideal living\npolymerization, and effective protocols for a variety of monomers have been established.\n3.\nCationic polymerization can also be living, although it is generally harder to do so than for the\nanionic case, in large part due to the prevalence of transfer reactions, including transfer to\nmonomer.\n4.\nUsing the concept of a reversibly dormant or inactive species, free-radical polymerizations\nhave also been brought under much greater control. Three general \ufb02avors of controlled\nradical polymerization, known as ATRP, SFRP, and RAFT, are currently undergoing rapid\ndevelopment.\n5.\nLiving polymerization in general, and anionic polymerization in particular, can be used to\nproduce block copolymers, end-functional polymers, and well-defined star and graft polymers,\nfor a variety of possible uses.\n6.\nThrough basic thermodynamic considerations the concepts of equilibrium polymerization,\nceiling temperature, and \ufb02oor temperature have been explored.\n7.\nThe utility of ROPs has been established, where the thermodynamic driving force for chain\ngrowth relies on ring-strain. Specific systems of nearly living ring-opening polymerizations\nhave been introduced, including important metal-catalyzed routes such as ROMP.\n"]], ["block_3", ["In this chapter, we have considered a wide variety of synthetic strategies to exert greater control\nover the products of a polymerization, compared to the standard step-growth and chain-growth\napproaches. Although access to much narrower molecular weight distributions has been the\nprimary focus, production of block copolymers, end-functional polymers, and controlled-branched\narchitectures has also been explored. The central concept of the chapter is that of a living\npolymerization, de\ufb01ned as a chain-growth process that proceeds in the absence of irreversible\nchain termination or chain transfer:\n"]], ["block_4", ["4.10\nChapter Summary\n"]], ["block_5", ["The initial demonstration of this approach was based on the following scheme [15]. The\nbuilding blocks were 3,5-dihydroxybenzyl alcohol and a benzylic bromide. The first reaction,\nconducted in acetone, coupled two of the bromides with one alcohol. The surviving benzylic\nalcohol was then transformed back to a bromide functionality with carbon tetrabromide in the\npresence of triphenyl phosphine. Introduction of more 3,5-dihydroxybenzyl alcohol began the\nformation of the next generation dendron, and the process continued.\n"]], ["block_6", ["160\nControlled Polymerization\n"]], ["block_7", [{"image_2": "174_2.png", "coords": [156, 150, 360, 245], "fig_type": "figure"}]], ["block_8", [{"image_3": "174_3.png", "coords": [276, 128, 431, 254], "fig_type": "figure"}]], ["block_9", [{"image_4": "174_4.png", "coords": [294, 124, 429, 263], "fig_type": "molecule"}]], ["block_10", ["(4.NN)\n"]]], "page_175": [["block_0", [{"image_0": "175_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n161\n"]], ["block_2", ["8.\nA particular class of highly branched, precisely controlled polymers called dendrimers can be\nprepared by either of two step-growth routes, referred to as convergent and divergent,\nrespectively.\n"]], ["block_3", ["Problems\n"]], ["block_4", ["TA. Soum and M. Fontanille, in Anionic Polymerization, J.E. McGrath, Ed., ACS Symposium Series, Vol1981.\ninn. Plesch, Ed, Cationic Polymerization, Macmillan, New York, 1963.\n"]], ["block_5", ["1.\nExperimental data cited in Example 4.1 for the anionic polymerization of styrene do not really\ntest the relationship between conversion, p and time; why not? What additional experimental\ninformation should have been obtained if that were the object?\nAlthough the polydispersities described in Example 4.1 are very low, they consistently exceed\nthe theoretical Poisson limit. List four assumptions that are necessary for the Poisson distri-\nbution to apply, and then identify which one is most likely not satisfied. Justify your answer,\nbased on the data provided.\nFor the living anionic polymerization of styrene discussed in Example 4.1, the solvent used\nwas cyclohexane, and the kinetics are known to be 0.5 order with respect to initiator. What is\nthe predominant species in terms of ion pairing, and what is the approximate dissociation\nconstant for this cluster if kp is actually 1000 mol L\u20181 s\u2018l?\nOne often-cited criterion for judging whether a polymerization is living is that 114,, should\nincrease linearly with conversion. Why is this not, in fact, a robust criterion?\nA living polymerization of 2\u2014vinyl pyridine was conducted using benzyl picolyl magnesium as\nthe initiator.)r Values of Mn were determined for polymers prepared with different initiator\nconcentrations and different initial concentrations of monomer, as shown below. Calculate the\nexpected Mn assuming complete conversion and 100% initiator efficiency; how well do the\ntheoretical and experimental values agree?\n"]], ["block_6", [{"image_1": "175_1.png", "coords": [41, 359, 270, 454], "fig_type": "figure"}]], ["block_7", ["The following table shows values of AH0 at 298 K for the gas phase reactions X(g) + H+(g)\n"]], ["block_8", ["\u2014> HX+ (g), where X is an olefin: Use these data to comment quantitatively of the\nfollowing points:\n"]], ["block_9", ["0.48\n82\n20\n0.37\n85\n25\n0.17\n71\n46\n0.48\n71\n17\n0.58\n73\n14\n0.15\n150\n115\n"]], ["block_10", ["\u20145 .8 k] mol_l, respectively, evaluate the AH for the rearrangement sec-butyl\nions and compare with the corresponding isomerization for the propyl cation.\n4.\nOf the monomers shown, only isobutene undergoes cationic polymerization to any\nsignificant extent. Criticize or defend the following proposition: the data explain this\nfact by showing that this is the only monomer listed that combines a sufficiently negative\nAH for protonation, with the freedom from interfering isomerization reactions.\n"]], ["block_11", ["1.\nThe cation is stabilized by electron-donating alkyl substituents.\n2.\nThe carbonium ion rearrangement of n-propyl ions to i-propyl ions is energetically\nfavored.\n3.\nWith the supplementary information that AH? of l\u2014butene and cis-Z-butene are +1.6 and\n"]], ["block_12", ["[I](mmolL_1)\n[M]0(mmolL\"l)\nMn (kg 11101\u20141)\n"]]], "page_176": [["block_0", [{"image_0": "176_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018A. Duda and S. Penczek Macromolecules, 23, 1636 (1990).\n"]], ["block_2", ["162\n"]], ["block_3", ["13.\n"]], ["block_4", ["14.\n"]], ["block_5", ["10.\n"]], ["block_6", ["11.\n"]], ["block_7", ["12.\n"]], ["block_8", [{"image_1": "176_1.png", "coords": [41, 61, 331, 147], "fig_type": "figure"}]], ["block_9", [{"image_2": "176_2.png", "coords": [51, 539, 280, 612], "fig_type": "figure"}]], ["block_10", ["In the study discussed in Example 4.3, a solution ATRP of styrene gave an apparent\npropagation rate constant of 3.9 x 10\u20145 s\u2018]. Given that the initial monomer concentration\nwas 4.3 mol L\u2018 1, and that the initial concentrations of initiator and CuBr were 0.045 mol L\u201d,\nestimate the equilibrium constant K for activation of the chain end radical.\nFor the solution polymerization of lactide with [M] =1 mol L\u201d, Duda and Penczek\ndetermined AHO \u201422.9 kJ mol\u20181 and A50: 411 J mol\" K\u2018\u201d. What is the associated\nceiling temperature for an equilibrium monomer concentration of\n1 mol L\u2018l? Does the\nvalue you obtain suggest that equilibration is an issue in controlled polymerization of\nlactide?\nCompare\nthese thermodynamic quantities with\nthose for the cyclic alkanes\nin Table 4.4; how do you account for the differences between the six-membered alkane\nand lactide?\nFor the polymerization system in Problem 8, calculate the equilibrium monomer con-\ncentration that would actually be obtained, and the conversion to polymer, at 80\u00b0C\nand 120\u00b0C.\nA typical propagation rate constant, kp, for the anionic ROP of hexamethylcyclotrisiloxane\n(D3) is 0.1 L mol\u20181 5\". Design a polymerization system (initial monomer and initiator\nconcentrations) to obtain a narrow distribution poly(dimethylsiloxane) with Mn=50,000,\nassuming that the reaction will be terminated at 50% conversion. At what time should the\npolymerization be terminated?\nGiven that ROP is often conducted under conditions in which reverse reactions are\npossible, do we need to worry about cyclization of the entire growing polymer? Why or\nwhy not?\nSuggest a scheme to test the hypothesis that in lactide polymerization it is the acyl carbon\u2014\noxygen bond that is cleaved, rather than the alkyl carbon\u2014oxygen bond.\nBoth ethyleneimine and ethylene sul\ufb01de are amenable to ROP. The former proceeds in the\npresence of acid, whereas the latter can follow either anionic or cationic routes. Propose\nstructures for the three propagating chain ends and the resulting polymers.\n"]], ["block_11", ["Draw repeat unit structures for polymers made by ROMP of the following three monomers:\n"]], ["block_12", ["H\nN\n8\nAA\nA\n"]], ["block_13", ["CH3CH2CH CH2\nCH3CH2CH2CH2+\n\u2014682\nCH3CH CHCH3\nCH3CH2C+HCH3\n\u2014782\n(CH3)2C CH2\n(CH3)2CHCH2+\n\u2014695\n"]], ["block_14", ["X\nHX+\nAHO at 298 K (kJ mol\u201c)\n"]], ["block_15", ["CH2 CH2\nCH3CH;\n\u2014640\nCHgCH CH2\nCH3CH2CH2+\n\u201469O\nCH3CH CH2\nCH3C+HCH3\n\u2014757\n"]], ["block_16", ["Me\n"]], ["block_17", ["Me\nMW\nOH\n"]], ["block_18", [{"image_3": "176_3.png", "coords": [201, 546, 276, 593], "fig_type": "molecule"}]], ["block_19", ["Controlled Polymerization\n"]]], "page_177": [["block_0", [{"image_0": "177_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Further Readings\n163\n"]], ["block_2", ["N. Hadjichristidis, S. PiSpas, and G. Floudas, Block Copolymers: Synthetic Strategies, Physical Pr0perties,\nand Applications, John Wiley & Sons, Inc., New York, 2003.\n"]], ["block_3", ["References\n"]], ["block_4", ["15.\n"]], ["block_5", ["16.\n"]], ["block_6", ["18.\n"]], ["block_7", ["Further Readings\n"]], ["block_8", ["17.\n"]], ["block_9", ["12.\n"]], ["block_10", ["13.\n14.\n15.\n"]], ["block_11", ["11.\n"]], ["block_12", ["9\u00b0\n"]], ["block_13", ["1.\n"]], ["block_14", ["249).\u201d?\n"]], ["block_15", ["G. Odian, Principles of Polymerization, 4th ed., Wiley\u2014Interscience, Hoboken, NJ, 2004.\nJ.E. Puskas, P. Antony, Y. Kwon, C. Paulo, M. Kovar, P.R. Norton, G. Kaszas, and V. Altst'adt,\nMacromol. Mater. Eng, 286, 565 (2001).\nM. Sawamoto, C. Okamoto, and T. Higashimura, Macromolecules, 20, 2693 (1987).\n"]], ["block_16", ["Rev., 101, 3661 (2001).\nG. Moad, E. Rizzardo, S.H. Thang, Aust. J. Chem, 58, 379 (2005).\nCJ. Hawker, Acc. Chem. Res, 30, 373 (1997).\nR]. Flory, Nature (London), 62, 1561 (1940).\nC]. Hawker and J.M.J. Fr\u00e9chet, J. Am. Chem. Soc., 112, 7639 (1990).\n"]], ["block_17", ["TR. Darling, T.P. Davis, M. Fryd, A.A. Gridnev, D.M. Haddleton, S.D. Ittel, R.R. Matheson, Jr.,\nG. Moad, and E. Rizzardo, J. Polym. Sci., Polym. Chem. Ed., 38, 1706 (2000).\nM. Szwarc, Nature, 178, 1168 (1956).\nH.L. Hsieh and RP. Quirk, Anionic Polymerization, Principles and Practical Applications, Marcel\nDekker, Inc., New York, NY, 1996.\nN. Hadjichristidis, M. Pitsikalis, S. Pispas, and H. Iatrou, Chem. Rev., 101, 3747 (2001).\nY. Tsukahara, K. Tsutsumi, Y. Yamashita, and S. Shimada, Macromolecules, 23, 5201 (1990).\n"]], ["block_18", ["J .E. Puskas, G. Kaszas, and M. Litt, Macromolecules, 24, 5278 (1991).\nK. Matyjaszewski, (Ed.), Controlled/Living Polymerization: Progress in ATRP, ACS\nSymposium Series 768, Oxford University Press (2000).\nK. Matyjaszewski, Chem. Rev., 101, 2921 (2001); C]. Hawker, A.W. Bosman, and E. Harth, Chem.\n"]], ["block_19", ["Compare and contrast dendrimers with block copolymer micelles, globular proteins, inor-\nganic nanoparticles in terms of attributes and likely utility in the following applications: (a)\ndrug delivery; (b) homogenous catalysis; and (c) solubilization.\nIn an ideal living polymerization, how should Mn and MW/Mn vary with conversion\nof monomer to polymer? How should Mn of the formed polymer vary with time?\nCompare these to a radical polymerization with termination by disproportionation, and no\ntransfer.\nThe following criteria have all been suggested and/or utilized as diagnostics for whether a\npolymerization is living or not. For each one, explain why it might be useful, and then decide\nwhether or not it is a robust criterion, that is, can you think of a situation in which the\ncriterion is satis\ufb01ed but the polymerization is not living? (See also Problem 4.)\nl.\nPolymerization proceeds until all monomer is consumed. Polymerization continues if\nmore monomer is then added.\n2.\nThe number of polymer molecules is constant, and independent of conversion.\n3.\nNarrow molecular weight distributions are produced.\n4.\nThe concentration of monomer decreases to zero, exponentially with time.\n"]], ["block_20", ["Suggest monomer structures that will lead to the following repeat unit structures following\nROMP:\n/\nN\nW\nONO\n"]], ["block_21", [{"image_1": "177_1.png", "coords": [89, 79, 334, 157], "fig_type": "figure"}]], ["block_22", ["I\nMe\n"]]], "page_178": [["block_0", [{"image_0": "178_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["P. Rempp and E.W. Merrill, Polymer Synthesis, 2nd ed., Hijthig & Wepf, Basel, 1991.\nM. Szwarc and M. van Beylen, Ionic Polymerization and Living Polymers, Chapman and Hall, New York, 1993.\n"]], ["block_2", ["H.L. Hsieh and RP. Quirk, Anionic Polymerization, Principles and Practical Applications, Marcel Dekker,\nInc., New York, NY, 1996.\nG. Odian, Principles of Polymerization, 4th ed., Wiley\u2014Interscience, Hoboken, NJ, 2004.\n"]], ["block_3", ["164\nControlled Polymerization\n"]]], "page_179": [["block_0", [{"image_0": "179_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["All polymer molecules have unique features of one sort or another at the level of the individual\nrepeat units. Occasional head-to-head or tail-to-tail orientations, random branching, and the\ndistinctiveness of chain ends are all examples of such details. In this chapter, we shall focus\nattention on two other situations that introduce structural variation at the level of the repeat unit:\nthe presence of two different monomers, and the regulation of configuration of successive repeat\nunits. In the former case copolymers are produced, and in the latter case polymers with differences\nin tacticity. In the discussion of these combined topics, we use statistics extensively because the\ndescription of microstructure requires this kind of approach. This is the basis for merging a\ndiscussion of copolymers and stereoregular polymers into a single chapter. In other respects\nthese two classes of materials and the processes that produce them are very different, and their\ndescription leads us into some rather diverse areas. Copolymerization offers a facile means to tune\nmaterial properties, as the average composition of the resulting polymers can often be varied across\nthe complete composition range. Similarly, control of stereoregularity plays an essential role in\ndictating the crystallinity of the resulting material, which in turn can exert a profound in\ufb02uence on\nthe resulting physical properties.\nThe formation of copolymers involves the reaction of (at least) two kinds of monomers. This\nmeans that each must be capable of undergoing the same propagation reaction, but it is apparent\nthat quite a range of reactivities are compatible with this broad requirement. We shall examine\nsuch things as the polarity of monomers, the degree of resonance stabilization they possess, and the\nsteric hindrance they experience in an attempt to understand these differences in reactivity. There\nare few types of reactions for which chemists are successful in explaining all examples with\ngeneral concepts such as these, and polymerization reactions are no exception. Even for the\nspeci\ufb01c case of free\u2014radical copolymerization, we shall see that reactivity involves the interplay\nof all these considerations.\nTo achieve any sort of pattern in configuration among successive repeat units in a polymer\nchain, the tendency toward random addition must be overcome. Although temperature effects are\npertinent here\u2014remember that high temperature is the great randomizer\u2014real success in regulat-\ning the pattern of successive addition involves the use of catalysts that \u201cpin down\u201d both the\nmonomer and the growing chain so that their reaction is biased in favor of one mode of addition or\nanother. We shall discuss the Ziegler\u2014Natta catalysts that accomplish this, and shall discover these\nto be complicated systems for which no single mechanism is entirely satisfactory. We shall also\ncompare these to the more recently developed \u201csingle\u2014site\u201d catalysts, which offer great potential\nfor controlling multiple aspects of polymer structure.\nFor both copolymers and stereoregular polymers, experimental methods for characterizing the\nproducts often involve spectroscopy. We shall see that nuclear magnetic resonance (NMR)\nspectroscopy is particularly well suited for the study of tacticity. This method is also used for\nthe analysis of copolymers.\n"]], ["block_2", ["5.1\nIntroduction\n"]], ["block_3", ["Copolymers, Microstructure,\nand Stereoregularity\n"]], ["block_4", ["5\n"]], ["block_5", ["165\n"]]], "page_180": [["block_0", [{"image_0": "180_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In writing Equation 5.2.1 through Equation 5.2.4 we make the customary assumption that the\nkinetic constants are independent of the size of the radical, and we indicate the concentration of all\nradicals ending with the M1 repeat unit, whatever their chain length, by the notation [M10]. This\nformalism therefore assumes that only the nature of the radical chain end in\ufb02uences the rate\nconstant for propagation. We refer to this as the terminal control mechanism. If we wished to\nconsider the effect of the next-to-last repeat unit in the radical, each of these reactions and the\nassociated rate laws would be replaced by two alternatives. Thus Reaction (5.A) becomes\n"]], ["block_2", ["and Equation 5.2.1 becomes\n"]], ["block_3", ["Each of these reactions is characterized by a propagation constant, which is labeled by a two-digit\nsubscript: the first number identifies the terminal repeat unit in the growing radical and the second\nidentifies the added monomer. The rate laws governing these four reactions are:\n"]], ["block_4", ["The polymerization mechanism continues to include initiation, termination, and propagation steps,\nand we ignore transfer reactions for simplicity. This time, however, there are four distinctly\ndifferent propagation reactions:\n"]], ["block_5", ["We begin our discussion of copolymers by considering the free-radical polymerization of a mixture\nof two monomers, M1 and M2. This is already a narrow view of the entire \ufb01eld of copolymers, since\nmore than two repeat units can be present in copolymers and, in addition, mechanisms other than free-\nradical chain growth can be responsible for copolymer formation. The essential features ofthe problem\nare introduced by this simpler special case, and so we shall restrict our attention to this system.\n"]], ["block_6", ["In spite of the assortment of things discussed in this chapter, there are also related topics that\ncould be included but which are not owing to space limitations. We do not discuss copolymers\nformed by the step\u2014growth mechanism, for example, or the use of Ziegler\u2014Natta catalysts to\nregulate geometrical isomerism in, say, butadiene polymerization. Some other important omissions\nare noted in passing in the body of the chapter.\n"]], ["block_7", ["5.2.1\nRate Laws\n"]], ["block_8", ["5.2\nC0polymer Composition\n"]], ["block_9", ["166\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_10", ["12,2, knmzonmz]\n(5.2.4)\n"]], ["block_11", ["Ran \n(5-2-1)\n"]], ["block_12", ["Rp.12 k12[M1\u00b0][M2]\n(5.2.2)\n"]], ["block_13", ["R1321 k21iM2\u00b0liM1l\n(52.3)\n"]], ["block_14", ["\u2014M2\u00b0 + M2L \u2014-M2M2o\n(5.D)\n"]], ["block_15", ["\u2014M1- + M1 L \u2014M1Mlo\n(5A)\n_M,. + M2 A. \u2014M1M2-\n(5.3)\n"]], ["block_16", ["\u2014M2. + M1 L) \u2014M2M1o\n(5.C)\n"]], ["block_17", ["Rp.211 k211[M2M1\u00b0][M1]\n(5.2.6)\n"]], ["block_18", ["Rp,111 kllliMlMl\u2018HMll\n(5.2.5)\n"]], ["block_19", ["km\n*MlMl\u2019 + M] -\u2014-\u2014-\u2014> \u2014M1M1M1\u00b0\n(5.13)\n"]], ["block_20", ["\u2014M2M1- + M1 J\ufb02\u2014v \u2014M2M1M1-\n(5F)\n"]]], "page_181": [["block_0", [{"image_0": "181_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where the is considered. \nourselves to the only the terminal unit determines behavior, \nin which the penultimate effect is important are well known.\nThe magnitudes of the various k values in Equation 5.2.1 through Equation 5.2.4 describe the\nintrinsic of addition, and the ks \nthe different species determine the rates at which the four kinds of additions occur. It is the\nproportion of different steps that determines the composition of the copolymer produced.\nMonomer M1 is converted to polymer by Reaction (5A) and Reaction (5C); therefore the rate\nat which this occurs is the sum of Rl and R1921:\n"]], ["block_2", ["Likewise, Reaction (5B) convert M2 to polymer, and the rate at \noccurs is the sum of RP,\u201d and R922:\n"]], ["block_3", ["We saw in Chapter 3 that the stationary-state approximation is applicable to free-radical\nhomOpolymerizations, and the same is true of copolymerizations. Of course, it takes a brief time\nfor the stationary-state radical concentration to be reached, but this period is insignificant com-\npared to the total duration of a polymerization reaction. If the total concentration of radicals is\nconstant, this means that the rate of crossover between the different types of terminal units is also\nequal, or that Rp.l2 =Rp.213\n"]], ["block_4", ["Copolymer Composition\n167\n"]], ["block_5", ["The ratio of Equation 5.2.7 and Equation 5.2.8 gives the relative rates of the two monomer\nadditions and, hence, the ratio of the two kinds of repeat units in the copolymer:\n"]], ["block_6", ["or\n"]], ["block_7", ["Although there are a total of four different rate constants for propagation, Equation \nthat the relationship between the relative amounts of the two monomers incorporated \npolymer and the composition of the monomer feedstock involves only two \nof these constants. Accordingly, we simplify the notation by defining reactivity ratios:\n"]], ["block_8", ["Combining Equation 5.2.9 and Equation 5.2.11 yields one form of the important \ncomposition equation or copolymerization equation:\n"]], ["block_9", ["and\n"]], ["block_10", ["diMli/dl\u2018 _ k11[M1\u00b0][M1]+ k21[M2\u00b0][M1]\nd[M2]/d[\n\u2014\nk12[M1\u00b0][M2] + k22[M2\u00b0][M2]\n(5.2.9)\n"]], ["block_11", ["k12[M1\u00b0][M2] k21[M2'][M1]\n(5.2.10)\n"]], ["block_12", ["r1 = ki\n(5.2.13)\nklz\n"]], ["block_13", ["\u2014i:\u2019:l\u2014] k11[M1\u00b0][M1]+ k21[M2\u00b0][M1]\n(5.2.7)\n"]], ["block_14", ["\u2014 \ndt\n= k12[M1'][M2] + k22[M2\u00b0][M2]\n(5.2.8)\n"]], ["block_15", ["d[M1]/dr_ [M1] (kll/k12)[M1] + [M2]\ndil/df\n_\n[M2] (kzz/k21)[M2] + [M1]\n(5.2.12)\n"]], ["block_16", ["r, \n(5.2.14)\n"]], ["block_17", ["[My] _ kzliMll\n__\n5.2.11\n[My]\nklziMz]\n(\n)\n"]]], "page_182": [["block_0", [{"image_0": "182_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["2.\nIf r1 r2, the copolymer and the feed mixture have the same composition at f1 0.5. In this\ncase Equation 5.2.18 becomes F1 :(r + 1)/2(r + 1) 0.5.\n3.\nIf r1 =r2, with both values less than unity, the copolymer is richer in component 1 than the\nfeed mixture for f1 < 0.5, and richer in component 2 than the feed mixture for f1 > 0.5.\n4.\nIf r1 r2, with both values greater than unity, an S-shaped curve passing through the point\n(0.5, 0.5) would also result, but in this case re\ufb02ected across the 45\u00b0 line compared to item (3).\n5.\nIf r1 74 r2, with both values less than unity, the copolymer starts out richer in monomer 1 than\nthe feed mixture and then crosses the 45\u00b0 line, and is richer in component 2 beyond this\n"]], ["block_2", ["This equation relates the composition of the copolymer formed to the instantaneous composition of\nthe feedstock and to the reactivity ratios r1 and r2 that characterize the specific system.\nFigure 5.1 shows a plot of F1 versusf1\u2014\u2014the mole fractions of monomer 1 in the copolymer and\nin the mixture, respectively\u2014for several values of the reactivity ratios. Inspection of Figure 5.1\nbrings out the following points:\n"]], ["block_3", ["Combining Equation 5.2.15 and Equation 5.2.16 into Equation 5.2.17 yields another form of the\ncopolymer composition equation\n"]], ["block_4", ["As an alternative to Equation 5.2.15, it is convenient to describe the composition of both\nthe polymer and the feedstock in terms of the mole fraction of each monomer. Defining F,- as the\nmole fraction of the ith component in the polymer and f,- as the mole fraction of component i in\nthe monomer solution, we observe that\n"]], ["block_5", ["and\n"]], ["block_6", ["With these substitutions, Equation 5.2.12 becomes\n"]], ["block_7", ["The ratio (d[M1]/dt)/(d[M2]/dt) is the same as the ratio of the numbers of each kind of repeat\nunit in the polymer formed from the solution containing M1 and M2 at concentrations [M1] and\n[M2], respectively. Since the composition of the monomer solution changes as the reaction\nprogresses, Equation 5.2.15 applies to the feedstock as prepared only during the initial stages of\nthe polymerization. Subsequently, the instantaneous concentrations in the prevailing mixture apply\nunless monomer is added continuously to replace that which has reacted and maintain the original\ncomposition of the feedstock. We shall assume that it is the initial product formed that we describe\nwhen we use Equation 5.2.15 so as to remove uncertainty as to the monomer concentrations.\n"]], ["block_8", ["5.2.2\nComposition versus Feedstock\n"]], ["block_9", ["168\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_10", ["1.\nIf r1 r2 1, the copolymer and the feed mixture have the same composition at all times. In\nthis case Equation 5.2.18 becomes\n"]], ["block_11", ["[M1]\nf1 \n\u2014_[M1]+[M2]\n(5.2.17)\n"]], ["block_12", ["_\n_\n(\u201cMd/(If\nF\u2018 F2 d[M1]/dt+d[M2]/dt\n(5.2.16)\n"]], ["block_13", ["d[M]]/dt = [M1] \u201c[111] + [M2] = 1+ r([M1]/[M2]\n(5 15)\nd[M2]/dl\u2018\n[M2] r2[M2] + [M1]\nI + r2[M2]/[M1]\n\u2018\n.\n"]], ["block_14", ["F1 M =fl\n(5.2.19)\n(f1 +12)2\n"]], ["block_15", ["1:\nr1f12 +f1f2\nI'1f12'l'2f1f2+r2f22\n"]], ["block_16", ["(5.2.18)\n"]]], "page_183": [["block_0", [{"image_0": "183_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "183_1.png", "coords": [26, 38, 294, 314], "fig_type": "figure"}]], ["block_2", ["Figure 5.1\nMole fraction of component\n1 in the copolymer as a function of feedstock composition for\nvarious reactivity ratios.\n"]], ["block_3", ["There is a parallel between the composition of a copolymer produced from a certain feed and\nthe composition of a vapor in equilibrium with a two-component liquid mixture. The following\nexample illustrates this parallel when the liquid mixture is an ideal solution and the vapor is an\nideal gas.\n"]], ["block_4", ["An ideal gas obeys Dalton\u2019s law; that is, the total pressure is the sum of the partial pressures of the\ncomponents. An ideal solution obeys Raoult\u2019s law; that is, the partial pressure of the ith component\nabove a solution is equal to the mole fraction of that component in the solution times the vapor\npressure of pure component 1'. Use these relationships to relate the mole fraction of component 1 in\nthe equilibrium vapor to its mole fraction in a two\u2014component solution and relate the result to the\nideal case of the copolymer composition equation.\n"]], ["block_5", ["Copolymer Composition\n169\n"]], ["block_6", ["Example 5.1\n"]], ["block_7", ["0.6 \n_\n"]], ["block_8", ["0.8\nr-\nr1: r2 :1\n\u2014\n"]], ["block_9", ["0.4\n\u2014\nF1 0.33. f2 0.57\n_\n"]], ["block_10", ["0.2 \n\u2014\n"]], ["block_11", ["For the case of r1 0.33 and r2 0.67 shown in Figure 5.1, [Md/[M2] equals 0.5 and\nf1 0.33. This mathematical analysis shows that a comparable result is possible with both r1\nand r2 greater than unity, but is not possible for r1 > 1 and r2 < 1.\nWhen r1: l/rz, the copolymer composition curve will be either convex or concave when\nviewed from the F1 axis, depending on whether r1 is greater or less than unity. The further\nremoved from unity r1 is, the farther the composition curve will be displaced from the 45\u00b0 line.\nThis situation where rlrz l is called an idea! copolymerizarion. The example below explores\nthe origin of this terminology.\n"]], ["block_12", ["0\n_|\n1\nl\nI\nI\nl\n1\n[\ni\nI\n1\nI\nl\n1\n1\n1\n"]], ["block_13", ["\u20181\nI\n\u2014l\nI\nI\nI\n1\n1'\n1\nI\nI\nI\nl\nl\nl\n1\n"]], ["block_14", ["crossover point. At the crossover point the copolymer and feed mixture have the same\ncomposition. The monomer ratio at this point is conveniently solved by Equation 5.2.15:\n"]], ["block_15", ["(W)cross_ r1\n(5.2.20)\n"]], ["block_16", ["0\n02\n0.4\n06\n08\n1\n"]], ["block_17", ["[M1]\n_ \n"]], ["block_18", ["._\nr1=10,f2=0.1\n_\n"]], ["block_19", ["\u2014\nr1=r2=0.15\n-\n"]], ["block_20", ["r1=0.1, {2:10\n"]]], "page_184": [["block_0", [{"image_0": "184_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The parameters r1 and r2 are the vehicles by which the nature of the reactants enter the copolymer\ncomposition equation. We shall call these radical reactivity ratios simply reactivity ratios, although\nsimilarly defined ratios also describe copolymerizations that involve ionic intermediates. There are\nseveral important things to note about reactivity ratios:\n"]], ["block_2", ["This is identical to the ideal liquid\u2014vapor equilibrium if r1 is identified with p?/pg.\nThe vapor pressure ratio measures the intrinsic tendency of component 1 to enter the vapor phase\nrelative to component 2. Likewise r1 measures the tendency of M1 to add to M10 relative to M1\nadding to My. In this sense there is a certain parallel, but it is based on My as a reference radical and\nhence appears to be less general than the vapor pressure ratio. Note, however, that n Hr; means\n"]], ["block_3", ["k1 1/1612 kzl/kgg. In this case the ratio of rate constants for monomer 1 relative to monomer 2 is the\nsame regardless of the reference radical examined. This shows the parallelism to be exact.\n"]], ["block_4", ["Because of the analogy with liquid\u2014vapor equilibrium, copolymers for which n 1/5 are said\nto be ideal. For those nonideal cases in which the copolymer and feedstock happen to have the\nsame composition, the reaction is called an azeotropic polymerization. Just as in the case of\nazeotropic distillation, the composition of the reaction mixture does not change as copolymer is\nformed if the composition corresponds to the azeotrope. The proportion of the two monomers at\nthis point is given by Equation 5.2.20.\nIn this section we have seen that the copolymer composition depends to a large extent on the\nfour propagation constants, although it is sufficient to consider these in terms of the two\nreactivity ratios\n1', and r2. In the next section we shall examine these ratios in somewhat\ngreater detail.\n"]], ["block_5", ["We define F1 to be the mole fraction of component\n1 in the vapor phase and f1 to be its mole\nfraction in the liquid solution. Here p1 and p2 are the vapor pressures of components\n1 and 2 in\nequilibrium with an ideal solution, and p? and p3 are the vapor pressures of the two pure liquids. By\nDalton\u2019s law, Ptot=P1+P2 and F1 =p1/pt0t, since these are ideal gases and p is proportional to\nthe number of moles. By Raoult\u2019s law, p1 =f1p\u20181), p2 fn, and pm, =f1p? +fn. Combining the\ntwo gives\n"]], ["block_6", ["Now examine Equation 5.2.18 for the case of r1 1/r2:\n"]], ["block_7", ["1.\nThe single subscript used to label r is the index of the radical.\n2.\nr1 is the ratio of two propagation constants involving radical 1: The ratio always compares the\npropagation constant for the same monomer adding to the radical relative to the propagation\nconstant for the addition of the other monomer. Thus, if r1 > 1, My adds M1 in preference to\nM2; if r1 < 1, M10 adds M2 in preference to M1.\n3.\nAlthough r1 is descriptive of radical Mr, it also depends on the identity of monomer 2; the\npair of parameters r, and r2 are both required to characterize a particular system, and the\nproduct rlrz is used to quantify this by a single parameter.\n4.\nThe reciprocal of a radical reactivity ratio can be used to quantify the reactivity of monomer\nM2 by comparing its rate of addition to radical M10 relative to the rate of M1 adding M10.\n"]], ["block_8", ["5.3\nReactivity Ratios\n"]], ["block_9", ["Solution\n"]], ["block_10", ["170\nCapolymers, Microstructure, and Stereoregularity\n"]], ["block_11", [{"image_1": "184_1.png", "coords": [39, 145, 194, 188], "fig_type": "molecule"}]], ["block_12", ["F1\n"]], ["block_13", ["F \nfir)?\n:\nfn(P?/p3)\n1\nflPti'l'fZPg\nf1(P?/Pg)+f2\n"]], ["block_14", ["=\nwe +1315\n=r1f1(r1f1+f2):\nma\nr1f12+2flf2+(l/rl)f22\n(r1f1+f22)\nr1f1+f2\n"]]], "page_185": [["block_0", [{"image_0": "185_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 5.1\nValues of Reactivity Ratios r1 and r2 and the Product rlrz for a Few Copolymers at 60\u00b0C\n"]], ["block_2", ["The reactivity copolymerization system are the fundamental parameters in terms of\nwhich the system is described. Since the copolymer composition equation relates the compositions\nof the product itis clear that values of r can be evaluated from experimental data\nin which the corresponding compositions are measured. We shall consider this evaluation procedure\nin Section be this approach is not as free of ambiguity as might be\ndesired. For now we shall simply assume that we know the desired r values for a system; in fact,\nextensive tabulations of such values exist. An especially convenient source of this information is\nthe Polymer Handbook [1]. Table 5.] lists some typical r values at 60\u00b0C.\nAlthough Table 5.1 is rather arbitrarily assembled, note that it contains no system for which r]\nand r2 are both greater than unity. Indeed, such systems are very rare. We can understand this by\nrecognizing that, at least in the extreme case of very large r\u2019s, these monomers would tend to\nsimultaneously homopolymerize. Because of this preference toward homopolymerization, any\ncopolymer that does form in systems with r1 and r2 both greater than unity will be a block-type\npolymer with very long sequences of a single repeat unit. Since such systems are only infrequently\nencountered, we shall not consider them further.\nTable 5.1 also lists the product rlrg for the systems included. These products typically lie in the\nrange between zero and unity, and it is instructive to consider the character of the copolymer\nproduced toward each of these extremes.\nIn the extreme case where r1r2:0 because both r1 and r2 equal zero, the copolymer adds\nmonomers with perfect alternation. This is apparent from the definition of r, which compares the\naddition of the same monomer to the other monomer for a particular radical. If both r\u2019s are zero,\nthere is no tendency for a radical to add a monomer of the same kind as the growing end,\n"]], ["block_3", ["Source: Data from Young, L]. in Polymer Handbook, 3rd ed., Brandrup, J. and Immergut, E.H. (Eds), Wiley, New York,\n1989.\n"]], ["block_4", ["Acrylonitrile\nMethyl vinyl ketone\n0.61\n1.78\n1.09\nMethyl methacrylate\n0.13\n1.16\n0.15\nor-Methyl styrene\n0.04\n0.20\n0.008\nVinyl acetate\n4.05\n0.061\n0.25\nMethyl methacrylate\nStyrene\n0.46\n0.52\n0.24\nMethacrylic acid\n1.18\n0.63\n0.74\nVinyl acetate\n20\n0.015\n0.30\nVinylidene chloride\n2.53\n0.24\n0.61\nStyrene\nVinyl acetate\n55\n0.01\n0.55\nVinyl chloride\n17\n0.02\n0.34\nVinylidene chloride\n1.85\n0.085\n0.16\n2-Vinyl pyridine\n0.55\n1.14\n0.63\nVinyl acetate\nl-Butene\n2.0\n0.34\n0.68\nIsobutylene\n2.15\n0.31\n0.67\nVinyl chloride\n0.23\n1.68\n0.39\nVinylidene chloride\n0.05\n6.7\n0.34\n"]], ["block_5", ["Rd]\nDd;\nr1\nr2\nrlrz\n"]], ["block_6", ["5.\nAs the ratio of two rate constants, a radical reactivity ratio follows the Arrhenius equation with\nan apparent activation energy equal to the difference in the activation energies for the\nindividual constants. Thus for r1, Efpp E1351 El\u201d; 12. Since the activation energies for pI\u2018Opa-\ngation are not large to begin with, their difference is even smaller. Accordingly, the tempera-\nture dependence of r is relatively small.\n"]], ["block_7", ["5.3.1\nEffects of r Values\n"]], ["block_8", ["Reactivity \n171\n"]], ["block_9", [{"image_1": "185_1.png", "coords": [121, 451, 456, 640], "fig_type": "figure"}]], ["block_10", [{"image_2": "185_2.png", "coords": [136, 457, 418, 635], "fig_type": "table"}]]], "page_186": [["block_0", [{"image_0": "186_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Returning to the data of Table 5.1, it is apparent that there is a good deal of variability among\nthe r values displayed by various systems. We have already seen the effect this produces on the overall\ncopolymer composition; we shall return to this matter ofmicrostructures in Section 5 .5. First, however,\nlet us consider the obvious question. What factors in the molecular structure of the two monomers\ngovern the kinetics of the different addition steps? This question is considered in the following\nsections; for now we look for a way to systematize the data as the first step toward an answer.\n"]], ["block_2", ["whichever species is the terminal unit. When only one of the r\u2019s is zero, say r1, then alternation\noccurs whenever the radical ends with an M1. unit. There is thus a tendency toward alternation in\nthis case, although it is less pronounced than in the case where both r\u2019s are zero. Accordingly, we\n\ufb01nd increasing tendency toward alternation as r1 \u2014> 0 and r2 \u2014> 0, or, more succinctly, as the\nproduct rlrz \u2014> 0.\nOn the other end of the commonly encountered range we find the product rlrz \u2014> 1. As noted\nabove, this limit corresponds to ideal copolymerization and means the two monomers have the\nsame relative tendency to add to both radicals. Thus if r1 \u2014> 10, monomer 1 is 10 times more likely\nto add to M10 than monomer 2. At the same time r2 0.1, which also means that monomer 1 is 10\ntimes more likely to add to My. In this case the radicals exert the same in\ufb02uence, so the monomers\nadd at random in the proportion governed by the specific values of the r\u2019s.\nRecognition of these differences in behavior points out an important limitation on the copolymer\ncomposition equation. The equation describes the overall composition of the copolymer, but gives\nno information whatsoever about the distribution of the different kinds of repeat units within the\npolymer. While the overall composition is an important property of the copolymer, the detailed\nmicrostructural arrangement is also a significant feature of the molecule. It is possible for\ncopolymers with the same overall composition to have very different properties because of the\ndifferences in microstructure. Reviewing the three categories presented in Chapter 1, we see\nthe following:\n"]], ["block_3", ["2.\nRandom structures are promoted by rlrz \u2014\u2014> 1:\n"]], ["block_4", ["4.\nAlternating copolymers, while relatively rare, are characterized by combining the properties of\nthe two monomers along with structural regularity. We will see in Chapter 13 that a very high\ndegree of regularity\u2014extending all the way to stereoregularity in the configuration of the\nrepeat units\u2014is required for crystallinity to develop in polymers.\n5.\nRandom copolymers tend to average the properties of the constituent monomers in proportion\nto the relative abundance of the two comonomers.\n6.\nBlock copolymers are closer to blends of homopolymers in properties, but without the latter\u2019s\ntendency to undergo phase separation. Diblock copolymers can be used as surfactants to bind\nimmiscible homopolymer blends together and thus improve their mechanical properties. Block\ncopolymers are generally prepared by sequential addition of monomers to living polymers,\nrather than by depending on the improbable rlrg > 1 criterion in monomers, as was discussed\nin Chapter 4.\n"]], ["block_5", ["3.\n\u201cBlocky\u201d structures are promoted by rlrz > 1:\n"]], ["block_6", ["172\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_7", ["1.\nAlternating structures are promoted by r;\n\u2014\u2014> O and r2 \u2014> 0:\n"]], ["block_8", ["Each of these polymers has a 50:50 proportion of the two components, but the products\nprobably differ in properties. As examples of such differences, we note the following:\n"]], ["block_9", ["M1MZMIMZMIMZMIMZMIMZMIMZMIMZMIMZMIMQMIMZ\n"]], ["block_10", ["MlMgMgMlMlMnM1M2M1M2M1M2M1M2M2M1M1M2\n"]], ["block_11", ["M1MiMiMiMiMiMIMlMlM1M2M2M2M2M2M2M2M2M2M2\n"]]], "page_187": [["block_0", [{"image_0": "187_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We noted above that the rlrzcan used locatea copolymer along \nalternating and random Itmeans of product that some \n"]], ["block_2", ["has been \n"]], ["block_3", ["3.\nThe spacing of the lines is such that each monomer along the base serves as a label for a row of\ndiamonds.\n4.\nEach diamond marks the intersection of two such rows and therefore corresponds to two\ncomonomers.\n"]], ["block_4", ["5.3.2\nRelation of Reactivity Ratios to Chemical Structure\n"]], ["block_5", ["supplemented \n"]], ["block_6", ["Reactivity Ratios\n173\n"]], ["block_7", ["6.\nIndividual monomers have been arranged in such a way as to achieve to the greatest extent\npossible the values of rlrz that approach zero toward the apex of the triangle and values of rlrz\nthat approach unity toward the base of the triangle.\n"]], ["block_8", ["5.\nThe rlrz product for the various systems is the number entered in each diamond.\n"]], ["block_9", ["1,\nVarious monomers are listed along the base of the triangle.\nThe triangle is subdivided into an array of diamonds by lines drawn parallel to the two sides of\nthe triangle.\n"]], ["block_10", ["Figure 5.2\nThe product r1r2 for copolymers whose components define the \nappear. The value marked* is determined in Example 5.4; other values are from Ref. [1].\n"]], ["block_11", ["Increasing\ntendency\n0-00017\n.\ntoward\nalternation\n0.00240ee\n\u2019:\u00b03\u00b0:\u00b0e\n\u201d\u00b0Ai%\u00a3\nALLA\n"]], ["block_12", ["<1)\nas\n(:5\nE\nH\non\na:\n=\na:\n2\u2018\n\u2014\nCD\n03\nE\na\nE\ng\n\u00a32\nCC\u201d\n0\n\u20181\u2019\n91\n3\nS\n'0\n_L\nc\n7\u201c\n0\n_cu\n(DE\n_EU\n.9\n0\nCD\na:\n.9>~\n>~\n0\n>335\nrag.\nE5\n>5\n5\nSP,\n0.\n\u2014c\n+-\nE\u2018\n+-.._\n#0\n>~.:\n+-\n3\u2019\u201d:\nH\nCD\n(Una\n.93\n(.3\nC13\";\n0q\n.Eo\nCUE\n.20\nCD\na\nE\nC]\n<\nE\nE\n>\nE\n>\n"]], ["block_13", [{"image_1": "187_1.png", "coords": [49, 279, 373, 586], "fig_type": "figure"}]], ["block_14", ["0.006\n"]], ["block_15", ["a.)\n_L\n"]], ["block_16", ["a?\n9n\n"]], ["block_17", ["Increasing\ntendency\ntoward\nrandomness\n"]]], "page_188": [["block_0", [{"image_0": "188_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "188_1.png", "coords": [25, 422, 287, 476], "fig_type": "molecule"}]], ["block_2", [{"image_2": "188_2.png", "coords": [31, 423, 275, 474], "fig_type": "figure"}]], ["block_3", ["Before proceeding with a discussion of this display, it is important to acknowledge that the\ncriteria for monomer placement can be met only in part. For one thing, there are combinations for\nwhich data are not readily available. Incidentally, not all of the rlrz values in Figure 5.2 were\nmeasured at the same temperature, but, as noted above, temperature effects are expected to be\nrelatively unimportant. Also, there are outright exceptions to the pattern sought: Generalizations\nabout chemical reactions always seem to be plagued by these. In spite of some reversals of ranking,\nthe predominant trend moving upward from the base along any row of diamonds is a decrease in\nrlrz values.\nFrom the geometry of this triangular display, it follows immediately\u2014if one overlooks\nthe exceptions\u2014that the more widely separated a pair of comonomers are in Figure 5.2, the\ngreater is their tendency toward randomness. We recognize a parallel here to the notion\nthat widely separated elements in the periodic table will produce more polar bonds than those\nwhich are closer together, and vice versa. This is a purely empirical and qualitative trend. The next\norder of business is to seek an explanation for its origin in terms of molecular structure. If we focus\nattention on the electron-withdrawing or electron-donating attributes of the substituents on the\ndouble bond, we find that the substituents of monomers that are located toward the right-hand\ncorner of the triangle in Figure 5.2 are recognized as electron donors. Likewise, the substituents\nin monomers located toward the left-hand corner of the triangle are electron acceptors. The\ndemarcation between the two regions of behavior is indicated in Figure 5.2 by reversing the direction\nof the lettering at this point. Pushing this point of view somewhat further, we conclude that the\nsequence acetoxy < phenyl < vinyl is the order of increase in electron-donating tendency. Chlor-\no < carbonyl < nitrile is the order of increase in electron-withdrawing tendency. The positions of\ndiethyl fumarate and vinylidene chloride relative to their monosubstituted analogs indicates\nthat \u201cmore is better\u201d with respect to these substituent effects. The location of methyl methacrylate\nrelative to methyl acrylate also indicates additivity, this time with partial compensation of\nopposing effects.\nThe reactivity ratios are kinetic in origin, and therefore re\ufb02ect the mechanism or, more\nspeci\ufb01cally, the transition state of a reaction. The transition state for the addition of a vinyl\nmonomer to a growing radical involves the formation of a partial bond between the two species,\nwith a corresponding reduction of the double-bond character of the vinyl group in the monomer:\n"]], ["block_4", ["The contribution of this polar structure to the bonding lowers the energy of the transition state. This\nmay be viewed as a lower activation energy for the addition step and thus a factor that promotes\nthis particular reaction. The effect is clearly larger the greater the difference in the donor\u2014acceptor\nproperties of X and Y. The transition state for the successive addition of the same monomer\n(whether X or Y substituted) is Structure (5.11). This involves a more uniform distribution of\n"]], ["block_5", ["If substituent X is an electron donor and Y an electron acceptor, then the partial bond in the\ntransition state is stabilized by a resonance form (5.1), which attributes a certain polarity to the\nemerging bond:\n"]], ["block_6", ["174\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_7", [{"image_3": "188_3.png", "coords": [38, 427, 223, 464], "fig_type": "molecule"}]], ["block_8", [{"image_4": "188_4.png", "coords": [40, 508, 201, 595], "fig_type": "figure"}]], ["block_9", ["1\n\u201dN.\n+ ///\\Y \u2014..\nWE\n"]], ["block_10", ["5\u2018 3\"\ni\ni\nas\n\u201dMS\nx\nv\nx\nx\n"]], ["block_11", [{"image_5": "188_5.png", "coords": [64, 527, 197, 576], "fig_type": "figure"}]], ["block_12", ["\u2014--\nProduct\n(5G)\nx \nY\nX\n"]], ["block_13", ["[5.I]\n[5-H]\n"]]], "page_189": [["block_0", [{"image_0": "189_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 5.2\nValues of the Cross-Propagation Constants kn (L mol'1 3*) for Four Monomer\u2014Radical\nCombinations\n"]], ["block_2", ["Styrene\n145\n49,000\n14,000\n230,000\nAcrylonitrile\n435\n1,960\n2,510\n46,000\nMethyl acrylate\n203\n1,310\n2,090\n23,000\nVinyl acetate\n2.9\n230\n230\n2,300\n"]], ["block_3", ["Source: From Brandrup, J. and Immergut, EH. (Eds), Polymer Handbook, 3rd ed., Wiley, New York, 1989.\n"]], ["block_4", ["a substituent, whether it is in a monomer or a radical. Conjugation allows greater electron\ndelocalization, which, in turn, lowers the energy of the system that possesses this feature.\nComparison of the range of km along rows and columns in Table 5.2 suggests that resonance\nstabilization produces a bigger effect in the radical than in the monomer. After all, the right- and\nleft\u2014hand columns in Table 5.2 (various radicals) differ by factors of 100\u20141000, whereas the top\n"]], ["block_5", ["Monomer\nStyrene\nAcrylonitrile\nMethyl acrylate\nVinyl acetate\n"]], ["block_6", ["Table 5.2 lists a few cross-propagation constants calculated by Equation 5.4.1. Far more extensive\ntabulations than this have been prepared by correlating copolymerization and homopolymerization\ndata for additional systems.\nExamination of Table 5.2 shows that the general order of increasing radical activity is\nstyrene < acrylonitrile < methyl acrylate < vinyl acetate. An additional observation is that any\none of these species shows the reverse order of reactivity for the corresponding monomers. As\nmonomers, the order of reactivity in Table 5.2 is styrene > acrylonitrile > methyl acrylate > vinyl\nacetate. These and similar rankings based on more extensive comparisons are summarized in terms\nof substituents in Table 5.3.\nAn important pattern to recognize among the substituents listed in Table 5.3 is this: Those that\nhave a double bond conjugated with the double bond in the olefin are the species that are more\nstable as radicals and more reactive as monomers. The inverse relationship between the stability of\nmonomers and radicals arises precisely because monomers gain (or lose) stability by converting to\nthe radical: The greater the gain (or loss), the greater (or less) the incentive for the monomer\nto react. It is important to realize that the ability to form conjugated structures is associated with\n"]], ["block_7", ["The tendency toward alternation is not the only pattern in terms of which copolymerization can be\ndiscussed. of radicals and monomers may also be examined as a source of insight\ninto copolymer formation. The reactivity of radical 1 copolymerizing with monomer 2 is measured\nby the rate constant km. The absolute value of this constant can be determined from copolymer-\nization data (r1) and studies yielding absolute homopolymerization constants (kn):\n"]], ["block_8", ["charge and thus lacks the stabilizing effect of the polar\nresonance of addition is that for alternation,\nat least X Y sufficiently different.\nAlthough we use the term resonance in describing the effect of polarity in stabilizing the\ntransition state in alternating copolymers, the emphasis of the foregoing is definitely on polarity\nrather that resonance plays an important role in free-\nradical polarity effects are ignored. In Section 5.4 we examine some\nevidence for this and consider the origin of this behavior.\n"]], ["block_9", ["5.4\nResonance and Reactivity\n"]], ["block_10", ["Resonance and Reactivity\n175\n"]], ["block_11", ["k12 \n(5.4.1)\nr1\n"]], ["block_12", ["Radical\n"]]], "page_190": [["block_0", [{"image_0": "190_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "190_1.png", "coords": [25, 70, 97, 343], "fig_type": "figure"}]], ["block_2", [{"image_2": "190_2.png", "coords": [27, 79, 314, 343], "fig_type": "figure"}]], ["block_3", ["Table 5.3\nList of Substituents Ranked in Terms of Their Effects on\nMonomer and Radical Reactivity\n"]], ["block_4", ["and bottom rows (various monomers) differ only by the factors of 50-100. In order to examine this\neffect in more detail, consider the addition reaction of monomer M to a reactant radical R0 to form\na product radical Po. What distinguishes these species is the presence or absence of resonance\nstabilization (subscript rs). If the latter is operative, we must also consider which species benefit\nfrom its presence. There are four possibilities:\n"]], ["block_5", ["O\n0\nIncreasing\nJLOH\n)LO/R\nIncreasing\nmonomer\nradical\nactivity\nactivity\n"]], ["block_6", ["176\nCopolymers, Microstructure. and Stereoregularity\n"]], ["block_7", ["I.\nUnstabilized monomer converts stabilized radical to unstabilized radical:\n"]], ["block_8", ["This reaction suffers none of the reduction in resonance stabilization that is present in Reaction\n(SH) and Reaction (5.1). It is energetically more favored than both of these, but not as much as\nthe reaction in which. . ..\nStabilized monomer converts unstabilized radical to stabilized radical:\n"]], ["block_9", ["There is an overall loss of resonance stabilization in this reaction. Since it is a radical which\nsuffers the loss, the effect is larger than in the reaction in which. . ..\nStabilized monomer converts stabilized radical to another stabilized radical:\n"]], ["block_10", ["Here too there is an overall loss of resonance stabilization, but it is monomer stabilization\nwhich is lost, and this is energetically less costly than Reaction (5.H).\nUnstabilized monomer converts unstabilized radical to another unstabilized radical:\nR+MeR\nan\n"]], ["block_11", ["Rrs' + Mrs Prs\u2018\n(SI)\n"]], ["block_12", ["R\u00b0 + Mrs Prs'\n(5K)\n"]], ["block_13", ["Rrs- + M \u2014> P.\n(S.H)\n"]], ["block_14", ["it o\nm\n"]], ["block_15", [",0\nFt\nY\n-R\nO\n"]], ["block_16", ["\u2014 CEN\nAO/R\n"]], ["block_17", ["/O\\R\n\u2014H\n"]], ["block_18", ["- Cl\n"]], ["block_19", ["ll\n"]]], "page_191": [["block_0", [{"image_0": "191_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "191_1.png", "coords": [28, 444, 250, 521], "fig_type": "figure"}]], ["block_2", ["Note that this inquiry into copolymer propagation rates also increases our understanding of the\ndifferences in free-radical homopolymerization rates. Recall that in Chapter 3 a discussion of this\naspect of homopolymerization was deferred until copolymerization was introduced. The trends\nunder consideration enable us to make some sense out of the rate constants for propagation in free-\nradical homopolymerization as well. For example, in Table 3.4 we see that kp values at 60\u00b0C for\nvinyl acetate and styrene are 2300 and 165 L mol'1 3\u2014], respectively. The relative magnitude of\nthese constants can be understood in terms of the sequence above.\nResonance stabilization energies are generally assessed from thermodynamic data. If we define\n"]], ["block_3", ["Radical\nStyrene\nStyrene\nVinyl acetate\nVinyl acetate\n+\n<\n+\n<\n+\n<\n+\nMonomer\nVinyl acetate\nStyrene\nVinyl acetate\nStyrene\n"]], ["block_4", ["31 to be the resonance stabilization energy of species 1', then the heat of formation of that species\nwill be less by an amount 31 than for an otherwise equivalent molecule without resonance.\nLikewise, the change in enthalpy AH for a reaction that is in\ufb02uenced by resonance effects is\nless by an amount As (A is the usual difference: products minus reactants) that the AH for a\nreaction which is otherwise identical except for resonance effects:\n"]], ["block_5", ["Thus if we consider the homopolymerization of ethylene (no resonance possibilities),\n"]], ["block_6", [{"image_2": "191_2.png", "coords": [34, 632, 278, 680], "fig_type": "molecule"}]], ["block_7", ["Systems from Table 5.2 which correspond to these situations are the following:\n"]], ["block_8", ["Resonance and Reactivity\n177\n"]], ["block_9", ["we find a value of A3 19 kJ mol\u2018l, according to Equation 5.4.2. Reaction (5.M) is a specific\nexample of the general Reaction (5.1), and the negative value of A3 in this example indicates the\noverall loss of resonance stabilization, which is characteristic of Reaction (5.1).\nAlthough it is not universally true that the activation energies of reactions parallel their heats of\nreaction, this is approximately true for the kind of addition reaction we are discussing. Accord-\ningly, we can estimate E* aAH, with a an appropriate proportionality constant. If we consider\nthe difference between two activation energies by combining this idea with Equation 5.4.2, the\ncontribution of the nonstabilized reference reaction drops out of Equation 5.4.2 and we obtain\n"]], ["block_10", ["as a reference reaction, and compare it with the homopolymerization of styrene (resonance effects\n"]], ["block_11", ["present),\n"]], ["block_12", ["AHrs AHnO ,8 A3\n(5.4.2)\n"]], ["block_13", ["Eii\u2014 Biz a[\u2014A311 0236312)]\n= \u2018(SPF 8R1. _SMI) + (8P2. _BRI. 8M2)\n(5.4.3)\n"]], ["block_14", ["AHno ,, \u2014\u201488.7 kJ mol-1\n"]], ["block_15", ["AH,, \u201469.9 kJ mol\u201d1\n"]], ["block_16", ["\u2014CH2\u2014CH2\u00b0 + CH2=CH2 -+ \u201cCH2CH2CH2CH2\u2018\n(5.L)\n"]], ["block_17", ["This reaction converts the less effective resonance stabilization of a monomer to a more effective\nform of radical stabilization. This is the most favorable of the four reaction possibilities.\n"]], ["block_18", ["Rrso +M< Rrso +Mrs < R- +M< Ro +Mrs\n"]], ["block_19", [{"image_3": "191_3.png", "coords": [58, 141, 315, 194], "fig_type": "molecule"}]], ["block_20", ["In summary, we can rank these reactions in terms of their propagation constants as follows:\n"]], ["block_21", [".H\n+\n\\\n\u2014+\n.H\n"]]], "page_192": [["block_0", [{"image_0": "192_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "192_1.png", "coords": [34, 122, 233, 154], "fig_type": "molecule"}]], ["block_2", ["1.\nDefine styrene to be monomer 1 and vinyl acetate to be monomer 2.\nThe difference in resonance stabilization energy 3p;\n\u2014\u2014 3M. >1, since styrene is resonance\nstabilized and the effect is larger for the radical than the monomer.\n3.\nThe difference 8132\u2018 3M2 \u00a7 0, since neither the radical nor the monomer of vinyl acetate\nshows appreciable stabilization.\n4.\nTherefore, according to Equation 5.4.4 and Equation 5.4.5, n > 1 while r2 < 1.\n5.\nExperimental values for this system are r1 55 and r2: 0.01.\n"]], ["block_3", ["We might be hard-pressed to estimate the individual resonance stabilization energies in\nEquation 5.4.4 and Equation 5.4.5, but the quantitative application of these ideas is not difficult.\nConsider once again the styrene-vinyl acetate system:\n"]], ["block_4", ["In writing the second version of this relation, the proportionality constant has been set equal to\nunity as a simpli\ufb01cation. Note that the resonance stabilization energy of the reference radical Rlo\nalso cancels out of this expression.\nThe temperature dependence of the reactivity ratio r1 also involves the E\u2019\ufb01\u2014 E\u2019f\u2018z difference\nthrough the Arrhenius equation; hence\n"]], ["block_5", ["1.\nThe reactivity ratios are proportional to the product of two exponentials.\n2.\nEach exponential involves the difference between the resonance stabilization energy of the\nradical and monomer of a particular species.\n3.\nA positive exponent is associated with the same species\nas identifies the r (i.e., for\n"]], ["block_6", ["An analogous expression can be written for r2:\nM)\n:.(_8PI'_\u2018\n8W\n5 4 5\nr2 oc exp(\nRT\nexp(\nRT\n)\n(\n.\n. )\n"]], ["block_7", ["1.\nIf resonance effects alone are considered, it is possible to make some sense of the ranking of\nvarious propagation constants.\n2.\nIn this case only random microstructure is predicted.\n3.\nIf polarity effects alone are considered, it is possible to make some sense out of the tendency\ntoward alternation.\n4.\nIn this case homOpolymerization is unexplained.\n"]], ["block_8", ["Although this approach does correctly rank the parameters r1 and r2 for the styrene\u2014vinyl acetate\nsystem, this conclusion was already reached qualitatively above using the same concepts and\nwithout any mathematical manipulations. One point that the quantitative derivation makes clear is\nthat explanations of copolymer behavior based exclusively on resonance concepts fail to describe\nthe full picture. All that we need to do is examine the product rlrg as given by Equation 5.4.4 and\nEquation 5.4.5, and the shortcoming becomes apparent. According to these relationships, the\nproduct rlrz always equals unity, yet we saw in the last section that experimental rlrg values\ngenerally lie between zero and unity. We also saw that polarity effects could be invoked to\nrationalize the rlrz product.\nThe situation may be summarized as follows:\n"]], ["block_9", ["178\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_10", ["According to this formalism, the following applies:\n"]], ["block_11", ["6P: \n3,4,)\n\u2014(8P2\u2018 3M2)\n5 4 4\nr1 oc exp(\u2014\u2014\u2014\u2014\u2014RT\nexp(\nRT\n)\n(\n.\n. )\n"]], ["block_12", ["11, M1\n\u2014-> Pp), whereas the negative exponent is associated with the other species (for\nr], M2 \u2014-> 132' ).\n"]]], "page_193": [["block_0", [{"image_0": "193_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "193_1.png", "coords": [33, 295, 238, 445], "fig_type": "figure"}]], ["block_2", ["It is apparent from the size of the conjugated system here that numerous resonance possibilities\nexist in this species in both the radical and the molecular form. Styrene also has resonance\nstructures in both forms. On the principle that these effects are larger for radicals than monomers,\nwe conclude that the difference 8}\u201d 8M > 0 for both hemin and styrene. On the principle that\ngreater resonance effects result from greater delocalization, we expect the difference to be larger\nfor hemin than for styrene. According to Equation 5.4.4, r1 oc el\u2018t\u2018rg\"\u2018\u201de\u20183\"\u201c\u00a3lller > 1. According to\nEquation 5.4.5, r2 oc esm\u2018i\u2018lh\u2018ire\u2014l\u2018t\u2018rgcr < 1. Experimentally, the values for these parameters turn out to\nbe r1: 65 and r2 0.18.\n"]], ["block_3", ["as these and resonance are both operative, and, if these still fall short of\nexplaining all observations, there is another old standby to fall back on: steric effects. Resonance,\npolarity, and considerations are all believed to play an important role in copolymerization\nchemistry of organic chemistry. Things are obviously simplified if only\none of these is considered, but it must be remembered that doing this necessarily reveals only one\nfacet of are times, particularly before launching an experimental\ninvestigation ofwhen some guidelines are very useful. The following example\nillustrates this point.\n"]], ["block_4", ["It is proposed to polymerize the vinyl group of the hemin molecule with other vinyl comonomers to\nprepare model compounds to be used in hemoglobin research. Considering hemin and styrene to be\nspecies 1 and 2, respectively, use the resonance concept to rank the reactivity ratios of r, and r2.\n"]], ["block_5", ["In Section 5.3 we noted that variations in the product rlrg led to differences in the polymer\nmicrostructure, even when the overall compositions of two systems are the same. In this section we\nshall take a closer look at this variation, using the approach best suited for this kind of detail,\nstatistics.\n"]], ["block_6", ["Hemin is the complex between protoporphyrin and iron in the +3 oxidation state. Iron is in the +2\nstate in the heme of hemoglobin. The molecule has the following structure:\n"]], ["block_7", ["A Closer Look at Microstructure\n179\n"]], ["block_8", ["The way out of this apparent dilemma is easily stated, although not easily acted upon. It is not\nadequate these approaches for the explanation of something as complicated\n"]], ["block_9", ["Solution\n"]], ["block_10", ["5.5\nA Closer Look at Microstructure\n"]], ["block_11", ["Example 5.2\n"]], ["block_12", [{"image_2": "193_2.png", "coords": [36, 300, 208, 411], "fig_type": "molecule"}]], ["block_13", [{"image_3": "193_3.png", "coords": [46, 333, 194, 393], "fig_type": "molecule"}]], ["block_14", ["OH\n"]], ["block_15", ["CH2\n"]]], "page_194": [["block_0", [{"image_0": "194_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Note that p11 + p12 p22 \u2014|\u2014 p21 1. In writing these expressions we make the assumption that only\nthe terminal unit of the radical in\ufb02uences the addition of the next monomer. This same assumption\nwas made in deriving the copolymer composition equation. We shall have more to say below about\nthis particular assumption.\nNext let us consider the probability of \ufb01nding a sequence of repeat units in a copolymer, which\nis exactly 12 units of M1 in length. This may be represented as M2(M1),,M2. Working from left to\nright in this sequence, we note the following:\n"]], ["block_2", ["For the various possible combinations in a copolymer, Equation 5.5.2 becomes\n"]], ["block_3", ["This equation can also be written in terms of the propagation rates of the different types of addition\nsteps which generate the sequences:\n"]], ["block_4", ["1.\nIf the addition of monomer M1 to a radical ending with M2 occurs L times in a sample, then\nthere will be a total of L sequences, of unspecified length, of M1 units in the sample.\n2.\nIf v 1 consecutive M1 monomers add to radicals capped by M1 units, the total number of\nsuch sequences is expressed in terms of p11 to be Lp'ff\u2018.\n3.\nIf the sequence contains exactly 12 units of type M], then the next step must be the addition of\nan M2 unit. The probability of such an addition is given by p12, and the number of sequences is\nLpiilplz-\n"]], ["block_5", ["The similarity of this derivation to those in Section 2.4 and Section 3.7 should be apparent.\nSubstitution of the probabilities given by Equation 5.5.3 and Equation 5.5.4 leads to\n"]], ["block_6", ["A similar result can be written for d)\u201c, where n denotes the length of a sequence of M2 units. These\nexpressions give the fraction of sequences of specified length in terms of the reactivity ratios of the\n"]], ["block_7", ["Suppose we define as p,3,- the probability that a unit of type i is followed in the polymer by a unit of\ntype j, where both i and j can be either 1 or 2. Since an 1' unit must be followed by either an i or aj,\nthe fraction of ij sequences (pairs) out of all possible sequences (pairs) de\ufb01nes pg:\n"]], ["block_8", ["Since L equals the total number of M1 sequences of any length, the fraction of sequences of length\n"]], ["block_9", ["180\nCopolymers, Microstructure. and Stereoregularity\n"]], ["block_10", ["5.5.1\nSequence Distributions\n"]], ["block_11", ["I), (by, is given by\n"]], ["block_12", ["k\nM \nM\nM\nP11=k\n\u201d[\n1H\n1]\n=\n\u201cl\n1]\n(5.5.3)\n11[M1'][M1] + k12[M1'][M2]\nr1[M1] + [M2]\n[M2]\n: \n(5.5.4\n\u201c2\nr1 [M1] + [M2]\n)\n"]], ["block_13", ["k\nM \nM\nM\np22 k\n22[\n2 H\n2]\n2\nrd\n2]\n(5.5.5)\n22[M2'][M2] + k21iM2'HM1]\nVziMzi -|- [M1]\nM\n[)2]\n[\n'1\n(5.5.6)\n: r2[M2] + [M1]\n"]], ["block_14", ["Ra\u201c\n_\nke\u2018li'l [Mr]\n1.. \n_\n(5.5.2\np}\nRij + RH\nkili'] [M1] + kiz\u2018lMi'HMil\n)\n"]], ["block_15", ["Number of if sequences\ni\u201c \n..\n..\n5.5.1\np 1\nNumber of 1] sequences + Number of :1 sequences\n(\n)\n"]], ["block_16", ["9b,, pi\u2019flpnz\n(5.5.7)\n"]], ["block_17", ["\u201cb\" ([MriihiliMzii\ufb02 (\ufb01lls/IT)\n(5'58)\n"]]], "page_195": [["block_0", [{"image_0": "195_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "195_1.png", "coords": [13, 365, 277, 636], "fig_type": "figure"}]], ["block_2", [{"image_2": "195_2.png", "coords": [30, 300, 188, 347], "fig_type": "molecule"}]], ["block_3", ["copolymer system and the composition of the feedstock. Figure 5.3 illustrates by means of a bar\ngraph how (1),, varies with v for two polymer systems prepared from equimolar solutions of\nmonomers. The shaded bars in Figure 5.3 describe the system for which \n"]], ["block_4", ["A Closer Look at Microstructure\n181\n"]], ["block_5", ["unshaded \nTable 5.4 shows the effect of variations in the composition of the feedstock for the system\nrlrz :1. The following observations can be made concerning Figure 5.3 and Table 5.4:\n"]], ["block_6", ["4.\nTable 5.4 also shows that increasing the percentage of M 1 in the monomer solution \ufb02attens\nand broadens the distribution of sequence lengths. Similar results are observed for lower\nvalues of rlrg, but the broadening is less pronounced when the tendency toward alternation\nis high.\n"]], ["block_7", ["2.\nFigure 5.3 shows that for rlrg 0.03, about 85% of the M1 units are sandwiched between two\nMg\u2019s. We have already concluded that low values of the rlrz product indicate a tendency\ntoward alternation.\n"]], ["block_8", ["3.\nFigure 5.3 also shows that the proportion of alternating M1 units decreases, and the fraction of\nlonger sequences increases, as rlrg increases. The 50 mol% entry in Table 5.4 shows that the\ndistribution of sequence lengths gets \ufb02atter and broader for 7'e 1, the ideal case.\n"]], ["block_9", ["Next we consider the average value of a sequence length of M1, 17. Combining Equation 1.7.7\nand Equation 5.5.7 gives\n"]], ["block_10", ["Simplifying this result involves the same infinite series that we examined in connection with\nEquation 2.4.5; therefore we can write immediately\n"]], ["block_11", [{"image_3": "195_3.png", "coords": [35, 458, 272, 601], "fig_type": "figure"}]], ["block_12", ["0/0\n_\n"]], ["block_13", ["451, 1968.)\n"]], ["block_14", ["1.\nIn all situations, the fraction (by decreases with increasing :2.\n"]], ["block_15", ["feedstocks with r1r2:0.03 (shaded) and r1r2:0.30 (unshaded). (Data from Tosi, C., \n"]], ["block_16", ["Figure 5.3\nFraction of sequences of the indicated length for copolymers \n"]], ["block_17", ["100\n\u2018I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nl\nl\nI\n"]], ["block_18", ["80\ne\n_\n"]], ["block_19", ["60\n\u2014\n..\n"]], ["block_20", ["40\n\u2014\nJ\n"]], ["block_21", ["20\n\u2014\n\u2014\n"]], ["block_22", ["0\n"]], ["block_23", ["17:\n(5.5.9)\n"]], ["block_24", ["\"\nD\n['11'2 0.30\n\"\n"]], ["block_25", ["..\nI\nr1r2 0.03\n"]], ["block_26", ["_\n"]], ["block_27", ["2:1 12(1),, : 2:,\n"]], ["block_28", ["2:195\u201d\n2:1Pif1l\u201912\n"]], ["block_29", ["1\n2\n3\n4\n5\n"]], ["block_30", [{"image_4": "195_4.png", "coords": [144, 416, 235, 462], "fig_type": "figure"}]], ["block_31", ["V1\n"]], ["block_32", ["I\n1\n1\n1\nI\nl W\nI\nI\n"]]], "page_196": [["block_0", [{"image_0": "196_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "196_1.png", "coords": [17, 77, 451, 225], "fig_type": "figure"}]], ["block_2", ["1\n90\n30\n70\n60\n50\n40\n30\n20\n10\n2\n9\n16\n21\n24\n25\n24\n21\n16\n9\n3\n0.9\n3.2\n6.3\n9.6\n12.5\n14.4\n14.7\n12.3\n3.1\n4\n0.09\n0.64\n1.39\n3.34\n6.25\n3.64\n10.3\n10.2\n7.29\n5\n0.13\n0.57\n1.54\n3.13\n5.13\n7.20\n3.19\n6.56\n6\n0.17\n0.62\n1.56\n3.11\n5.04\n6.55\n5.90\n7\n0.05\n0.25\n0.73\n1.37\n3.53\n5.24\n5.31\n3\n0.10\n0.39\n1.12\n2.47\n4.19\n4.73\n9\n0.04\n0.20\n0.67\n1.73\n3.36\n4.30\n10\n0.10\n0.40\n1.21\n2.63\n3.37\n11\n0.05\n0.24\n0.35\n2.15\n3.59\n12\n0.14\n0.59\n1.72\n3.23\n"]], ["block_3", ["The following example demonstrates the use of some of these relationships pertaining to\nmicrostructure.\n"]], ["block_4", ["The hemoglobin molecule contains four heme units. It is proposed to synthesize a hemin (molecule\n1)\u2014styrene (molecule 2) copolymer such that 17 = 4 in an attempt to test some theory concerning\nhemoglobin. As noted in Example 5.2, r1: 65 and r2 0.18 for this system. What should be the\nproportion of monomers to obtain this average hemin sequence length? What is the average styrene\nsequence length at this composition? Does this system seem like a suitable model if the four hemin\nclusters are to be treated as isolated from one another in the theory being tested? Also evaluate (1),.\nfor several v bracketing 17 to get an idea of the distribution of these values.\n"]], ["block_5", ["Table 5.4\nPercentage of Sequences of Length v for Copolymers Prepared from Different Feedstocks f1\nwith rlrz 1\n"]], ["block_6", ["Use Equation 5.5.11 to evaluate [Md/[M2] for r1 65 and 17 4:\n"]], ["block_7", ["A value of 11 is obtained by similar operations:\n"]], ["block_8", ["By combining Equation 5.5.4 and Equation 5.5.10, we obtain\n"]], ["block_9", ["Solution\n"]], ["block_10", ["Example 5.3\n"]], ["block_11", ["182\nCopolymers. Microstructure. and Stereoregularity\n"]], ["block_12", ["14f,\n0.1\n0.2\n0.3\n0.4\n0.5\n0.6\n0.7\n0.3\n0.9\n"]], ["block_13", ["it =1+r2%\n(5.5.12)\n"]], ["block_14", [",\u2014, =\n1\n= _1_\n(5.5.10)\n1 P11\nP12\n"]], ["block_15", ["17=1+r1-\u2014\u2014-\u2014\u2014\n(5.5.11)\n"]], ["block_16", ["[M1]\n17 1\n4 1\n[M2]\n=\n=\n= 0.046\nd\n\u2014 = 21.7\n[Mg]\nr1\n65\nan\n[M1]\n"]], ["block_17", [{"image_2": "196_2.png", "coords": [80, 91, 404, 223], "fig_type": "table"}]]], "page_197": [["block_0", [{"image_0": "197_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "197_1.png", "coords": [21, 233, 284, 287], "fig_type": "molecule"}]], ["block_2", [{"image_2": "197_2.png", "coords": [22, 145, 235, 226], "fig_type": "molecule"}]], ["block_3", ["A Closer Look at Microstructure\n183\n"]], ["block_4", ["Use this ratio 5.5.12 to i1:\n"]], ["block_5", ["The number of styrene units in an average sequence is a little larger than the length of the average\nhemin sequence. It is not unreasonable to describe the hemin clusters as isolated, on the average, in\nthis molecule. The product rlrz 11.7 in this system, which also indicates a tendency toward block\nformation. Use Equation with [Md/[M2] 0.046 and the r1 and r2 values to evaluate\n"]], ["block_6", ["For the systems represented in Figure 5.3 and the equimolar case in Table 5.4, the average\nlengths are 17 1.173 for r1r2= 0.03, 17 = 1.548 for rlrg :0.30, and 17 2.000 for rlrg = 1.0.\nEquation 5.5.1 1 and Equation 5.5.12 suggest a second method for the experimental determination\nof reactivity ratios, in addition to the copolymer composition equation. If the average sequence\nlength can be determined for a feedstock of known composition, then r1 and r2 can be evaluated. We\nshall return to this possibility in the next section. In anticipation of applying this idea, let us review\nthe assumptions and limitation to which Equation 5.5.11 and Equation 5.5.12 are subject:\n"]], ["block_7", ["We have suggested earlier that both the copolymer composition equation and the average sequence\nlength offer possibilities for experimental evaluation of the reactivity ratios. Note that in so doing\nwe are finding parameters which fit experimental results to the predictions of a model. Nothing\nabout this tests the model itself. It could be argued that obtaining the same values of r1 and r2 from\nthe fitting of composition and microstructure data would validate the model. It is not likely,\nhowever, that both types of data would be available and of sufficient quality to make this\nunambiguous. We shall examine the experimental side of this in the next section.\nStatistical considerations make it possible to test the assumption of independent Let\n"]], ["block_8", ["as the conversion of monomers to polymer progresses. As in Section 5.2, we assume that either\nthe initial conditions apply (little change has taken place) or that monomers are continuously\nbeing added (replacement of reacted monomer).\n2.\nThe kinetic analysis described by Equation 5.5.3 and Equation 5.5.4 assumes that no repeat\nunit in the radical other than the terminal unit in\ufb02uences the addition. The penultimate unit in\nthe radical as well as those still further from the growing end are assumed to have no effect.\n"]], ["block_9", ["Solving for several values of V, we conclude that the distribution of sequence length is quite broad:\n"]], ["block_10", ["(1),,\n0.251\n0.188\n0.140\n0.105\n0.079\n0.059\n"]], ["block_11", ["us approach this topic by considering an easier problem: coin tossing. Under conditions where two\nevents are purely random\u2014as in tossing a fair coin\u2014the probability of a specific sequence of\n"]], ["block_12", ["3.\nItem (2) requires that each event in the addition process be independent of all others. We have\nconsistently assumed this throughout this chapter, beginning with the copolymer composition\nequation. Until now we have said nothing about testing this assumption. Consideration of\ncopolymer sequence length offers this possibility.\n"]], ["block_13", [":2\n1\n2\n3\n4\n5\n6\n"]], ["block_14", ["5.5.2\nTerminal and Penultimate Models\n"]], ["block_15", ["1.\nThe instantaneous monomer concentration must be used. Except at the azeotrope, this changes\n"]], ["block_16", ["_\n[M2l\n=1+r\u2014=1+0.1821.7\n=49\n"]], ["block_17", ["= (%)V\u20141 <\ufb01) : (0.0749)\u201d\"1(0.251)\n"]]], "page_198": [["block_0", [{"image_0": "198_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We recall that the fraction of times a particular outcome occurs is used to estimate probabilities.\nTherefore we could evaluate pm; by counting the number of times NH the first toss yielded a head\nand the number of times NHH two tosses yielded a head followed by a head and write\n"]], ["block_2", ["If the events are not independent, provision must be made for this, so we de\ufb01ne a quantity called\nthe conditional probability. For the probability of a head given the prior event of a head, this is\nwritten pH/H, where the first quantity in the subscript is the event under consideration and that\nfollowing the slash mark is the prior condition. Thus pry/H is the probability of a tail following a\nhead. If the events are independent, ppm, :pH; if not, then pm; must be evaluated as a separate\nquantity. If the coin being tossed were biased, that is, if successive events are not independent,\nEquation 5.5.13 would become\n"]], ["block_3", ["This procedure is readily extended to three tosses. For a fair coin the probability of three heads is\nthe cube of the probability of tossing a single head:\n"]], ["block_4", ["If we were testing whether a coin were biased or not, we would use ideas like these as the basis\nfor a test. We could count, for example, HHH and HH sequences and divide them according to\nEquation 5.5.19. If pmHH 75 pH, we would be suspicious.\nA similar logic can be applied to copolymers. The story is a bit more complicated to tell, so we\nonly outline the method. If penultimate effects operate, then the probabilities p11, p12, etc., defined\nin Equation 5.5.3 through Equation 5.5.6 should be replaced by conditional probabilities. As a\nmatter of fact, the kind of conditional probabilities needed must be based on the two preceding\nevents. Thus Reaction (5.E) and Reaction (5.E) are two of the appropriate reactions, and the\ncorresponding probabilities are p1,\u201d and plm. Rather than work out all of the probabilities in\ndetail, we summarize the penultimate model as follows:\n"]], ["block_5", ["outcomes is given by the product of the probabilities of the individual events. The probability of\ntossing a head followed by a head~indicated HH\u2014is given by\n"]], ["block_6", ["If the coin is biased, conditional probabilities must be introduced:\n"]], ["block_7", ["Using Equation 5.5.15 to eliminate mm from the last result gives\n"]], ["block_8", ["01'\n"]], ["block_9", ["1.\nA total of eight different reactions are involved, since each reaction like Reaction (5A) is\nreplaced by a pair of reactions like Reaction (5.E) and Reaction (5.E).\n2.\nThere are eight different rate laws and rate constants associated with these reactions. Equation\n5.2.1, for example, is replaced by Equation 5.2.5 and Equation 5.2.6.\n"]], ["block_10", ["184\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_11", ["PHHH PH/HHpH/HPH\n(5.5.17)\n"]], ["block_12", ["PHH\nPHHH PH/HH (p\u2014H)PH\n(5.5.18)\n"]], ["block_13", ["PHH PHPH\n(5.5.13)\n"]], ["block_14", ["PHH PH/HPH\n(55-14)\n"]], ["block_15", ["_PHH _NHH\nPH/H \nW\nNH\n(55.15)\n"]], ["block_16", ["PHHH PHPHPH\n(5.5.16)\n"]], ["block_17", ["PH/HH \n(5.5.19)\nPHH\nNHH\n"]]], "page_199": [["block_0", [{"image_0": "199_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["those of four units, tetrads; those of five units, pentads; and so on. Next we examine the ratio\nNMIM/NMl and NMIMIMI /NM1M1. If these are the same, then the mechanism is Shown to have\nterminal control; if not, it may be penultimate control. To prove the penultimate model it would\nalso be necessary to count the number of M1 tetrads. If the tetrad\u2014triad ratio were the same as the\ntriad\u2014dyad ratio, the penultimate model is established.\nThis situation can be generalized. If the ratios do not become constant until the ratio of pentads\nto tetrads is considered, then the unit before the next to last\u2014called the antepenultimate unit\u2014\nplays a role in the addition. This situation has been observed, for example, for propylene oxide\u2014\nmaleic anhydride copolymers. The foregoing discussion has been conducted in terms of M]\nsequences. Additional relationships of the sort we have been considering also exist for dyads,\ntriads, and so forth, of different types of specific composition. Thus an ability to investigate\nmicrostructure experimentally allows some rather subtle mechanistic effects to be studied. In the\nnext section we shall see how such information is obtained.\n"]], ["block_2", ["As we have already seen, the reactivity ratios of a particular copolymer system determine both the\ncomposition and microstructure of the polymer. Thus it is important to have reliable values for\nthese parameters. At the same time it suggests that experimental studies of composition and\nmicrostructure can be used to evaluate the various r\u2019s.\n"]], ["block_3", ["Evaluation of reactivity ratios from the copolymer composition equation requires only composition\ndata\u2014~that is, relatively straightforward analytical chemistry\u2014and has been the method most widely\nused to evaluate r1 and r2. As noted in the last section, this method assumes terminal control and\n"]], ["block_4", ["quences, of any definite composition, of two units are called dyads; those of three units, triads;\n"]], ["block_5", ["These observations suggest how the terminal mechanism can be proved to apply to a copoly-\nmerization reaction if experiments exist which permit the number of sequences of a particular\nlength to be determined. If this is possible, we should count the number of Ml\u2019s (this is given by\nthe copolymer composition) and the number of MIMI and MIMIMI sequences. Speci\ufb01ed se-\n"]], ["block_6", ["seeks the best fit of the data to that model. It offers no means for testing the model, and as we shall see,\nis subject to enough uncertainty to make even self-consistency difficult to achieve. Microstructure\n"]], ["block_7", ["6.\nEquation 5.5.4 shows that p11 is constant for a particular copolymer if the terminal model\napplies; therefore the ratio NM.M./NM.\nalso equals this constant. Equation 5.5.20 shows\nthat P1111 is constant for a particular copolymer if the penultimate model applies; therefore\nthe ratio NM]M|M|/NM1M1 also equals this constant, but the ratio NMIMI /NM. does not have the\nsame value.\n"]], ["block_8", ["5.\nThe probability p11 can be written as the ratio NMIMl/NMI using Equation 5.5.15. This is\nreplaced by pm], which is given by the ratio NMM]Ml/NM,Ml according to Equation 5.5.19.\n"]], ["block_9", ["3.\nEight rate constants are clustered in four ratios, which define new reactivity ratios. Thus r1 as\ndefined in Equation 5.2.13 is replaced by rf (cm/km and rf\u2019 \ufb01rm/((212 whereas r; is\nreplaced by r5 k222/k221 and r\ufb01\u2019 km/km.\n4.\nThe probability p11 as given by Equation 5.5.3 is replaced by the conditional probability pl,1 1,\nwhich is defined as\n"]], ["block_10", ["5.6\nCopolymer Composition and Microstructure: Experimental Aspects\n"]], ["block_11", ["5.6.1\nEvaluating Reactivity Ratios from Composition Data\n"]], ["block_12", ["Copolymer Composition and Microstructure: Experimental Aspects\n185\n"]], ["block_13", ["=\nk111[M1M1'][M1]\n_\nr\ufb02Mll/[Mg]\npV1\u2018\nknitMe-HMI] + k112[M1M1o][M2]_1+ r{[M1]/[M2]\n(5.5.20)\n"]], ["block_14", ["There are eight of these conditional probabilities, each associated with the reaction described\nin item (1).\n"]]], "page_200": [["block_0", [{"image_0": "200_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["studies, by contrast, offer both a means to evaluate the reactivity ratios and also to test the\nmodel. The capability to investigate this level of structural detail was virtually nonexistent\nuntil the advent of modem instrumentation, and even now is limited to sequences of rather\nmodest length.\nIn this section we shall use the evaluation of reactivity ratios\nas the unifying theme;\nthe experimental methods constitute the new material introduced. The copolymer composition\nEquation 5.2.18 relates the r\u2019s to the mole fractions of the monomers in the feedstock and in\nthe copolymer. To use the equation to evaluate r1 and r2, the composition of a copolymer\nresulting from a feedstock of known composition must be measured. The composition of the\nfeedstock itself must also be known, but we assume this poses no problems. The copolymer\nspecimen must be obtained by pr0per sampling procedures and puri\ufb01ed of extraneous materials.\nRemember that monomers, initiators, and possibly solvents and soluble catalysts are involved\nin these reactions also, even though we have been focusing attention on the copolymer alone.\nThe proportions of the two kinds of repeat unit in the copolymer are then determined by\neither chemical or physical methods. Elemental analysis is a widely used chemical method, but\nspectroscopic analysis (UV\u2014visible, IR, NMR, and mass spectrometry) for functional groups is\ncommonly employed.\nSince the copolymer equation involves both r1 and r; as unknowns, at least two polymers\nprepared from different feedstocks must be analyzed. It is preferable to use more than this\nminimum number of observations, and it is helpful to rearrange the copolymer composition\nequation into a linear form so that simple graphical methods can be employed to evaluate the\nr\u2019s. Several ways to linearize the equation exist:\n"]], ["block_2", ["Each of these forms weigh the errors in various data points differently, so some may be more\nsuitable than others, depending on the precision of the data. Ideally all should yield the same\nvalues of the reactivity ratios. The following example illustrates the use of Equation 5.6.1 to\nevaluate r1 and r2.\n"]], ["block_3", ["2.\nIn terms of ratios rather than fractions, Equation 5.6.1 may be written as\nW(\ufb02_1)=rlw_r,\n(5.6.2)\nill/\u201d2\n\u201d2\n\u201d1/712\n"]], ["block_4", ["where n1 refers to the number of repeat units in the polymer. This expression is also of\nthe form y=mx+b if x=1\n)_\u2019 2 \u2014r2\u2014 _|_ r1\n(5.6.3)\nx\nx\n"]], ["block_10", ["1\nx:_y+2\nGeo\nr1\nr1\n"]], ["block_11", ["y\nrz)\u2019\nV1\n"]], ["block_12", ["This is the equation of a straight line, so r1 and r; can be evaluated from the slope and intercept\nof an appropriate plot.\n"]], ["block_13", [{"image_2": "200_2.png", "coords": [54, 428, 244, 457], "fig_type": "molecule"}]], ["block_14", ["1\n1\n"]]], "page_201": [["block_0", [{"image_0": "201_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "201_1.png", "coords": [22, 78, 358, 164], "fig_type": "figure"}]], ["block_2", [{"image_2": "201_2.png", "coords": [25, 463, 397, 649], "fig_type": "figure"}]], ["block_3", ["f1\nF1\nf1\nF1\n"]], ["block_4", ["ff(F1\u2014\u20141)/F1(1\u2014f1)2\n\u20140.0083\n20.0143\n\u20140.0316\n\u20140.0486\n\u20140.0832\n\u20140.1315\n\u20140.2792\n\u20140.7349\n"]], ["block_5", ["f1(1\u20142F1)/F1(1-f1)\n0.0217\n\u20140.1036\n\u20140.2087\n\u20140.3832\n\u20140.6127\n\u20140.9668\n\u20142.8102\n46.4061\n"]], ["block_6", ["Table \nValues of F1 as a Function for f1 for the Methyl Acrylate (MO\u2014Vinyl\nChloride \n"]], ["block_7", ["Note: These 5.4.\n"]], ["block_8", ["Source: Data from Chapin, and Fordyce, R., J. Am. Chem. Soc, 70, 538, 1948.\n"]], ["block_9", ["We calculate the variables to be used as ordinate and abscissa for the data in Table 5.5 using\nEquation 5.6.1:\n"]], ["block_10", ["0.075\n0.441\n0.421\n0.864\n0.154\n0.699\n0.521\n0.900\n"]], ["block_11", ["0.237\n0.753\n0.744\n0.968\n0.326\n0.828\n0.867\n0.983\n"]], ["block_12", ["The data in Table 5.5 list the mole fraction of methyl acrylate in the feedstock and in the\ncopolymer for the methyl acrylate (MO\u2014vinyl chloride (M2) system. Use Equation 5.6.1 as the\nbasis for the graphical determination of the reactivity ratios, which describe this system.\n"]], ["block_13", ["Least-square analysis of these values gives a slope r1 8.929 and an intercept r22 0.053.\nFigure 5.4b shows these data plotted according to Equation 5.6.1. The line is drawn with the\nleast\u2014squares slope and intercept. The last point on the left in Figure 5.4b, which this line passes\nthrough, corresponds to F1 0.983 andf1 0.867. Because the functional form plotted involves the\nsmall differences F 1 l and l f1, this point is also subject to the largest error. This illustrates\nthe value of having alternate methods for analyzing the data. The authors of this research carried\n"]], ["block_14", ["Copolymer Composition and Microstructure: Experimental Aspects\n187\n"]], ["block_15", ["the various methods were r1 9.616 i 0.603 and r2 0.0853 i 0.0239. The standard deviations\n"]], ["block_16", ["Figure 5.4\n(a) Mole fraction of methyl acrylate in copolymers with vinyl chloride as of feedstock\ncomposition, and (b) Finemann\u2014Ross plot to extract reactivity ratios, as described in Example 5.4.\n"]], ["block_17", ["Solution\n"]], ["block_18", ["out several different analyses of the same data; the values they obtained for r1 and r2 averaged over\n"]], ["block_19", ["Example 5.4\n"]], ["block_20", ["0'8:\u2014\n.\n_\n.25..\n_\u2018\nF1\n3\n\u00b0\n-\nE\nE\n0.7}\n.\n\u2014_\nZ\u2014\n'\n\u2014;\n"]], ["block_21", ["(a)\n\u20191\n(b)\n"]], ["block_22", ["0.9:\u2014\n.\n\u2014i\n05\n\u20182\n:\n\u00b0\n\u201c\nL\n\u2014;\n"]], ["block_23", [{"image_3": "201_3.png", "coords": [47, 470, 217, 628], "fig_type": "figure"}]], ["block_24", ["E\n\u2018\n\u201441\n\u2014f\n0'67\n\u201c\nE\n:\n"]], ["block_25", [":\nI\n=_\n_i\n0'5:\n:\n-6:-\nr]\n\u2014\no\n\u2014\n:\n2\n04\u2018liilltlillllllllrrr'l\n 'lllilllllllllllll'i \n0\n0.2\n0.4\n0.6\n0.8\n1\n-0-8\n\u20140-6\n\u20140.4\n\u20140.2\n0\n"]], ["block_26", ["171T\n[II\nill\nlrl\nIII\nII\n[II\nIII\nIII\nI_\n:\nI\nI\nI\n.1.\nr\n"]], ["block_27", [{"image_4": "201_4.png", "coords": [166, 485, 383, 633], "fig_type": "figure"}]], ["block_28", ["I\nI\n"]]], "page_202": [["block_0", [{"image_0": "202_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In the second version of this equation c is the speed of light and )t the wavelength of the radiation.\n3.\nThe more widely separated two states are in energy, the shorter the wavelength of the radiation\nabsorbed. Transitions between electronic states have higher energies, and correspond to UV\u2014\nvisible wavelengths, whereas vibrational quantum states are more closely spaced and are\ninduced by IR radiation.\n4.\nDifferent light-absorbing groups, called chromophores, absorb characteristic wavelengths,\nopening the possibility of qualitative analysis based on the location of an absorption peak.\n5.\nIf there is no band overlap in\na spectrum, the absorbance at a characteristic\nwave\u2014\nlength is proportional to the concentration of chromophores present. This is the basis of\nquantitative analysis using spectra. With band overlap, things are more complicated but\nstill possible.\n6.\nThe proportionality between the concentration of chromophores and the measured absorbance\nis given by Beer\u2019s law (recall the discussion in Section 3.3.4):\nA=wc\n658\n"]], ["block_2", ["The difference is positive for absorbed energy.\n2.\nThe energy absorbed is proportional to the frequency of the radiation via Planck\u2019s constant\n(h=6c3x1044te:\n"]], ["block_3", ["In spite of the compounding of errors to which it is subject, the foregoing method was the\nbest procedure for measuring reactivity ratios until the analysis of microstructure became feasible.\nLet us now consider this development. Most of the experimental information concerning\ncopolymer microstructure has been obtained by modern instrumental methods. Techniques such\nas UV-visible, IR, NMR, and mass spectroscopy have all been used to good advantage in\nthis type of research. Advances in instrumentation have made these physical methods particularly\nsuitable to answer the question we pose: With what frequency do particular sequences of repeat\nunits occur in a c0polymer? The choice of the best method for answering this question is governed\nby the specific nature of the system under investigation. Few general principles exist beyond\nthe importance of analyzing a representative sample of suitable purity. Our approach is to consider\nsome specific examples. In view of the diversity of physical methods available and the\nnumber of copolymer combinations which exist, a few samples barely touch the subject. They\nwill suffice to illustrate the concepts involved, however. The simpler question\u2014What is the mole\nfraction of each repeat unit in the polymer sample?\u2014can usually be answered via the same\ninstrumental techniques.\nSpectroscopic techniques based on the absorption of UV, visible, or IR radiation depend\non the excitation from one quantum state to another. References in physical or analytical chemistry\nshould be consulted for additional details, but a brief summary will be sufficient for our purposes:\n"]], ["block_4", ["of about 6% and 28% in r1 and r2 analyzedfrom the same data indicate the hazards of this method\nfor determining r values.\n"]], ["block_5", ["5.6.2\nSpectroscopic Techniques\n"]], ["block_6", ["1.\nThe excitation energy AE re\ufb02ects the separation between the final (subscript f) and initial\n(subscript i) quantum states:\nAEz\ufb02\u2014Ei\n65$\n"]], ["block_7", ["188\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_8", ["where A is the (dimensionless) absorbance, b is the sample thickness, 0 is the chromophore\nconcentration, and e is the absorptivity. Usually quantitative measurements are facilitated by\n"]], ["block_9", ["AE=hp=h\u00a7\nGan\n"]]], "page_203": [["block_0", [{"image_0": "203_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1.\nNuclei with an odd number of protons plus neutrons\u2014especially 1H and 13C\u2014possess magnetic\nmoments and show two quantum states (spin up and spin down) in a strong magnetic field.\n2.\nIf energy of the proper frequency is supplied, a transition between these quantum states occurs\nwith the absorption of an amount of energy equal to the separation of the states, just as in UV\u2014-\nvisible and IR absorption. For NMR the frequency of the absorbed radiation lies in the radio\nfrequency range and depends on the local magnetic field at the atom in question.\n3.\nElectrons in a molecule also have magnetic moments and set up secondary magnetic fields,\nwhich partly screen each atom from the applied field. Thus atoms in different chemical\nenvironments display resonance at slightly different magnetic fields.\n4.\nThe displacement 5 of individual resonances from that of a standard is small, and is measured\nin parts per million (ppm) relative to the applied field. These so-called chemical shifts are\ncharacteristic of a proton or carbon in a specific environment.\n5.\nThe interaction between nuclei splits resonances into multiple peaks, the number and relative\nintensity of which also assist in qualitative identification of the proton responsible for the\nabsorption. Proton splitting is most commonly caused by the interaction of protons on adjacent\ncarbons with the proton of interest. If there are m equivalent hydrogens on an adjacent carbon,\nthe proton of interest produces 771 + 1 peaks by this coupling.\n6.\nMore distant coupling is revealed in high magnetic fields. Unresolved fine structures in a field\nof one strength may be resolved at higher field where more subtle long-range influences can be\nprobed. The use of NMR spectroscopy to characterize copolymer microstructure takes advan-\ntage of this last ability to discern environmental effects that extend over the length of several\nrepeat units. This capability is extremely valuable in analyzing the stereoregularity of a\npolymer, and we shall have more to say about it in that context in Section 5.7.\n7.\nIn NMR spectroscopy the absorptivitiesare, in essence, all the same, so that the integrated area\nunder a peak is directly proportional to the number of nuclei of that type in the sample. Thus if\ndifferent repeat units have identifiably different peaks, as is almost always the case, the\nrelative abundance of each type can be extracted by peak integration without any additional\ncalibration.\n"]], ["block_2", ["NMR spectroscopy is especially useful for microstructure studies, because of the sensitivity to\nthe chemical environment of a particular nucleus. We shall consider its application to copolymers\nnow, and to questions of stereoregularity in Section 5.7. NMR has become such an important\ntechnique (actually a family of techniques) in organic chemistry that contemporary textbooks in\nthe subject discuss its principles quite thoroughly, as do texts in physical and analytical chemistry,\nso here also we note only a few pertinent highlights:\n"]], ["block_3", ["7 ,\nFor copolymers, or any other mixture of chromophores, the measured absorbance is given by\nthe sum of individual Beer\u2019s law terms:\n"]], ["block_4", ["Copolymer Composition and Microstructure: Experimental Aspects\n189\n"]], ["block_5", ["140.2) 810.2)i + 820.2)2902\n(5.610)\n"]], ["block_6", ["14011) 81011l1+ 82(A1)b02\n"]], ["block_7", ["A 81bC1+ 82bC2 -|- 831563 -|- \n(5.69)\n"]], ["block_8", ["These relations amount to a system of two equations with two unknowns, c1 and C2, which can\nbe solved in a straightforward manner.\n"]], ["block_9", ["Recalling that .9 depends on the chromophore and on the wavelength, measurements at\ndifferent wavelengths can be used to extract the concentrations of each component. For a\ncopolymer with two monomers, at least two wavelengths would be needed, and ideally they\nshould be chosen to such that if .91 is large at M, then 82 is large at A2.\n"]], ["block_10", ["calibration with standards of known concentrations, so that a, b, and various other instrumental\nparameters need not be determined individually.\n"]]], "page_204": [["block_0", [{"image_0": "204_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "204_1.png", "coords": [26, 419, 318, 623], "fig_type": "figure"}]], ["block_2", ["It is these conjugated double bonds that are the chromophores of interest in this system. What\nmakes this particularly useful is the fact that the absorption maximum for this chromophore is\ndisplaced to longer wavelengths the more conjugated bonds there are in a sequence. Qualita-\ntively, this can be understood in terms of a one-dimensional particle in a box model for which the\nenergy level Spacing is inversely proportional to the square of the length of the box. In this case\nthe latter increases with the length of the conjugated polyene system. This in turn depends on the\nnumber of consecutive butadiene repeat units in the copolymer. For an isolated butadiene\nmolecule dehalogenation produces one pair of conjugated double bonds; two adjacent butadienes,\nfour conjugated double bonds; three adjacent butadienes, six conjugated double bonds; and so on.\nSequences of these increasing lengths are expected to absorb at progressively longer wave-\nlengths.\nFigure 5.5 shows the appropriate portion of the spectrum for a copolymer prepared from a\nfeedstock for which f1 0.153. It turns out that each polyene produces a set of three bands: the\n"]], ["block_3", ["As suggested in the foregoing, the analysis for overall composition in a copolymer sample is by now a\nrelatively straightforward affair. The analysis for sequence distribution, however, is not. The primary\ndifficulty is that the energy of a particular transition, be it electronic, vibrational, or nuclear, is\ndetermined primarily by the immediate chromophore of interest, and only weakly in\ufb02uenced by\nchemically bonded neighbors. NMR offers the most promise in this respect, especially with the advent\nof higher magnetic \ufb01elds; this feature can provide sufficient resolution to detect the in\ufb02uence of repeat\nunits up to about five monomers down the chain. Nevertheless, there are cases where UV\u2014visible\nspectroscopy can help. An elegant example is the copolymer of styrene (molecule 1) and 1-chloro-1,3-\nbutadiene (molecule 2). These molecules quantitatively degrade with the loss of HCl upon heating in\nbase solution. This restores 1,3-unsaturation to the butadiene repeat unit:\n"]], ["block_4", ["Figure 5.5\nUltraviolet\u2014visible spectrum of dehydrohalogenated copolymers of styrene\u2014l-chloro-1,3-buta\u2014\ndiene. (Redrawn from Winston, A. and Wichacheewa, P., Macromolecules, 6, 200, 1973. With permission.)\n"]], ["block_5", ["5.6.3\nSequence Distribution: Experimental Determination\n"]], ["block_6", ["190\nCopolymers. Microstructure, and Stereoregularity\n"]], ["block_7", ["\u00e9\u2019\nTetrad\nTriad\nDyad\n"]], ["block_8", ["6 L-\n8g\nPentad\nEo\n4 -\n"]], ["block_9", [{"image_2": "204_2.png", "coords": [36, 191, 272, 268], "fig_type": "figure"}]], ["block_10", [{"image_3": "204_3.png", "coords": [39, 196, 256, 263], "fig_type": "molecule"}]], ["block_11", ["2 L\n"]], ["block_12", ["0\nL\nl\n_1\n_L\n1\n550\n500\n450\n400\n350\n300\n"]], ["block_13", ["3\n.\u2014\n"]], ["block_14", ["CI /\n\u2014nHC|\n\\ \\\nn\nm\nn\nm\n(5.N)\n"]], ["block_15", [{"image_4": "204_4.png", "coords": [108, 440, 304, 548], "fig_type": "figure"}]], ["block_16", ["Wavelength (nm)\n"]]], "page_205": [["block_0", [{"image_0": "205_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Mole fraction of dyads\nf1\n1 1\n12\n22\n"]], ["block_2", [{"image_1": "205_1.png", "coords": [34, 500, 162, 629], "fig_type": "figure"}]], ["block_3", ["MQMIMZ, MQMZMQ, MzMl, and MIMZMI. These can be divided into two groups of three,\ndepending on the identity ofcentral unit. Thus the center of a triad can be bracketed by two\n"]], ["block_4", ["same resonance feature in the copolymer of methyl methacrylate (M1) and acrylonitrile (M2).\nFigure 5.6 shows that 60 MHz spectrum several of these copolymers in the neighborhood of the\nmethoxy resonance. Three resonance peaks rather than one are observed. Figure 5.6 also lists the\nmethyl methacrylate content of each of these polymers. As the methyl methacrylate content\ndecreases, the peak on the right decreases and the left increases. We therefore identify the peak\non the right-hand peak with the MIMIMI sequence, the left-hand peak with MlMz, and the\npeak in the center with MIMIMQ. The MIMIM, peak occurs at the same location as in the methyl\nmethacrylate homopolymer.\nThe areas under the three peaks give the relative proportions of three sequences. In the\nfollowing example we consider some results on dyad sequences determined by comparable\nprocedures in vinylidene chloride\u2014isobutylene copolymers.\n"]], ["block_5", ["0.584\n0.68\n0.29\n\u2014\n"]], ["block_6", ["observed \ncompounds and the location and relative intensities of the peaks have been shown be \n"]], ["block_7", ["dent of these features have been identified, \ninterpreted \nunits and compared sequences of various lengths. Further consideration \n"]], ["block_8", ["monomers by of each. In \ncentral repeat unit in a differentand a characteristic proton in that repeat unit \n"]], ["block_9", ["different location,that environment. As a specific example, consider\nmethoxy group in poly(methyl methacrylate). The hydrogens in the group are magnetically\nequivalent and hence produce a single resonance at 8 3.74 ppm. Now suppose we look for the\n"]], ["block_10", ["367, and \nbands overlaps that of the triad and is not resolved. Likewise, only one band (at 473 nm) is\n"]], ["block_11", ["dyad is identified with the peaks at )1 298, 312, and 327 nm; the triad with the peaks at )t 347,\n"]], ["block_12", ["The mole fractions of various dyads in the vinylidine chloride (M1)\u2014\u2014isobutylene (M2) system\nwere determined,r by NMR spectroscopy. A selection of the values obtained are listed below, as\nwell as the compositions of the feedstocks from which the copolymers were prepared; assuming\nterminal control, evaluate r1 from each of the first three sets of data, and r2 from each of the last\nthree.\n"]], ["block_13", ["system is left for Problem through Problem 5 at the end of the chapter.\nWe now illustrate the application of NMR to gather 00polymer sequence information. Suppose\nwe consider the various triads of repeat units. There are six possibilities: MlMlMl, MIMIMQ,\n"]], ["block_14", ["0.505\n0.61\n0.36\n-\u2014\u2014\n"]], ["block_15", ["0.471\n0.59\n0.38\n\u2014\u2014\n"]], ["block_16", ["0.130\n\u2014\n0.67\n0.08\n0.121\n\u2014\n0.66\n0.10\n0.083\n\u2014\n0.64\n0.17\n"]], ["block_17", ["Copolymer Composition and Microstructure: Experimental Aspects\n191\n"]], ["block_18", ["tJ.B. Kinsinger, T. Fischer, and CW. Wilson, Polym. Lett., 5, 285 (1967).\n"]], ["block_19", ["Example 5.5\n"]]], "page_206": [["block_0", [{"image_0": "206_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "206_1.png", "coords": [10, 238, 271, 442], "fig_type": "figure"}]], ["block_2", [{"image_2": "206_2.png", "coords": [21, 37, 259, 466], "fig_type": "figure"}]], ["block_3", ["Equation 5.5.3 and Equation 5.5.5 provide the method for evaluating the r\u2019s from the data given.\nWe recognize that a 12 dyad can come about from\n1 adding to 2 as well as from 2 adding to 1;\ntherefore we use half the number of 12 dyads as a measure of the number of additions of monomer\n2 to chain end 1. Accordingly, by Equation 5.5.1,\n"]], ["block_4", ["Figure 5.6\nChemical shift (from hexamethyldisiloxane) for acrylonitrile\u2014methyl methacrylate copolymers\nof the indicated methyl methacrylate (M1) content. Methoxyl resonances are labeled as to the triad source.\n(From Chujo, R., Ubara, H., and Nishioka, A., Polym. J., 3, 670, 1972. With permission.)\n"]], ["block_5", ["Since [M1]/[M2] :f1/(1_f1)a Equation 5.5.2 can be written\n"]], ["block_6", ["Solution\n"]], ["block_7", ["192\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_8", ["N11\n2N11\n21s2\n=\nz\nand\n2\u2014\u2014\nN11+(1/2)N12\n2N11+N12\np22\n2N22 +N12\nP11\n"]], ["block_9", ["2\nrilfl/(l\u2014fill\nand\np\n=\nr2l(1\u201cf1)/f1l\n1+r1[f1/(1\u2014f1)]\n2\u20192\n1+r2l(1*~f1)/f:l\nP11\n"]], ["block_10", ["l\n|\nL\n|\n4.00\n3.90\n3.80\n3.70\n"]], ["block_11", [{"image_3": "206_3.png", "coords": [97, 65, 237, 277], "fig_type": "figure"}]], ["block_12", ["Chemical shift (ppm)\n"]], ["block_13", ["M1M1M2\n"]], ["block_14", ["33%\n"]], ["block_15", ["24%\n"]], ["block_16", ["50%\n"]]], "page_207": [["block_0", [{"image_0": "207_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["f1\nr1\nfl\n\u20192\n"]], ["block_2", [{"image_1": "207_1.png", "coords": [34, 528, 251, 636], "fig_type": "figure"}]], ["block_3", ["From\nFrom\n"]], ["block_4", ["Characterizing Stereoregularity\n193\n"]], ["block_5", ["2N1}\n=\nF|[f1/(1\u2014f1)]\n2s\n=\nr2[(1\u2014f1)/fl]\n2N11+N12\n1+?1lf1/(1-f1)]\n2N22+N12\n1+r2[(1-fI)/f11\n"]], ["block_6", ["0.584\n3.33\n0.130\n0.036\n"]], ["block_7", ["0.471\n3.48\n0.083\n0.048\nAverage\n3.38\nAverage\n0.042\n"]], ["block_8", ["We introduced the concept of Stereoregularity in Section 1.6. Figure 1.3 illustrates \nsyndiotactic, and atactic structures ofvinyl polymer in which \nfully extended chain lie, on the same side, alternating sides, \nto the backbone. It is important to appreciate the fact that these different \ncon\ufb01gurations\u2014have their origin in the bonding of the polymer, and no \nbonds-whanges in conformation-dwill convert one structure into another.\nOur discussion of stereoregularity in this chapter is primarily \nmonosubstituted ethylene repeat units. We shall represent these by\n"]], ["block_9", ["the penultimate model describes this better than the terminal model, \ncomings of the latter are not evident in the example. Problem 6 and Problem 7 at the of\nchapter also refer to this system.\n"]], ["block_10", ["0.505\n3.32\n0.121\n0.042\n"]], ["block_11", ["Particularly when rthis method for evaluating small \ngraphical analysis of composition data (compare Example 5.4 and Figure 5.4).\n"]], ["block_12", ["last example were able of tetrads of different composition \nvinylidene chloride\u2014isobutylene copolymer. Based on the longer sequences, they concluded \n"]], ["block_13", ["arising from still the authors of \n"]], ["block_14", ["these symbols, the isotactic, syndiotactic, and atactic structures \n"]], ["block_15", ["Monosubstituted ethylene\n"]], ["block_16", ["In this representation the X indicates the substituent; other bonds involve \nformalism also applies to 1,1-disubstituted ethylenes in which the substituents \n"]], ["block_17", ["5.7\nCharacterizing Stereoregularity\n"]], ["block_18", ["by Structure (5.111) through Structure (5V), reSpectively:\n"]], ["block_19", [{"image_2": "207_2.png", "coords": [39, 526, 162, 619], "fig_type": "figure"}]], ["block_20", ["+\ufb02xv\n"]], ["block_21", ["By making measurements at higher magnetic \ufb01elds, it is possible to resolve spectral features\n"]], ["block_22", ["l\nI\n"]], ["block_23", ["X\nX\n"]], ["block_24", ["X\nX\nX\nl\nI\n(5.111)\n"]], ["block_25", ["X\nX\n"]], ["block_26", ["1\nl\n"]], ["block_27", ["I\nX\n|\n"]], ["block_28", ["|\n|\n(5.1V)\n"]], ["block_29", ["i\n"]], ["block_30", ["l\nI\n(5V)\n"]]], "page_208": [["block_0", [{"image_0": "208_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The carbon atoms carrying the substituents are not truly asymmetric, since the two chain\nsections\u2014while generally of different length\u2014are locally the same on either side of any carbon\natom, except near the ends of the chain. As usual, we ignore any uniqueness associated with\nchain ends.\nThere are several topics pertaining to stereoregularity that we shall not cover to simplify the\npresentation:\n"]], ["block_2", ["The successive repeat units in Structure (5.III) through Structure (5.V) are of two different\nkinds. If they were labeled M1 and M2, we would find that, as far as microstructure is concerned,\nisotactic polymers are formally the same as homopolymers, syndiotactic polymers are formally the\nsame as alternating copolymers, and atactic polymers are formally the same as random copoly-\nmers. The analog of block copolymers, stereoblock polymers, also exist. Instead of using M1 and\nM2 to differentiate between the two kinds of repeat units, we shall use the letters D and L as we did\nin Chapter 1.\nThe statistical nature of polymers and polymerization reactions has been illustrated at many\npoints throughout this volume. It continues to be important in the discussion of stereoregularity.\nThus it is generally more accurate to describe a polymer as, say, predominantly isotactic rather\nthan perfectly isotactic. More quantitatively, we need to be able to describe a polymer in terms of\nthe percentage of isotactic, syndiotactic, and atactic sequences.\nCertain bulk properties of polymers also re\ufb02ect differences in stereoregularity. We will see in\nChapter 13 that crystallinity is virtually impossible unless a high degree of stereoregularity is\n"]], ["block_3", ["1.\nStereoregular copolymers. We shall restrict our discussion to stereoregular homopolymers.\nComplications arising from other types of isomerism. Positional and geometrical isomerism,\nalso described in Section 1.6, will be excluded for simplicity. In actual polymers these are not\nalways so easily ignored.\n3.\nPolymerization\nof\n1,2-disubstituted\nethylenes.\nSince\nthese\nintroduce\ntwo\ndifferent\n\u201casymmetric\u201d carbons into the polymer backbone (second substituent Y), they have the\npotential to display ditacticity. Our attention to these is limited to the illustration of some\nterminology, which is derived from carbohydrate nomenclature (Structure (5.VI) through\nStructure (5.IX)).\n"]], ["block_4", ["194\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_5", [{"image_1": "208_1.png", "coords": [42, 201, 196, 510], "fig_type": "figure"}]], ["block_6", [{"image_2": "208_2.png", "coords": [46, 307, 195, 356], "fig_type": "figure"}]], ["block_7", [{"image_3": "208_3.png", "coords": [48, 355, 187, 476], "fig_type": "figure"}]], ["block_8", [{"image_4": "208_4.png", "coords": [52, 369, 189, 411], "fig_type": "figure"}]], ["block_9", ["Y\nX\nY\nX\nJ\nI\nI\nI\nI\n|\nI\n1x\nI I n\nI I I\nI\n6-\n>\n"]], ["block_10", ["Y\nX\nY\nX\nY\nX\nY\nX\n"]], ["block_11", ["X\nX\nX\nl\nI\nI\nI\nl\nI\n|\nI\n|\n|\nI\nI\nI\n|\nI\n|\n(5.VIII)\nY\nY\nY\nY\n"]], ["block_12", ["Y\nX\nY\nX\nI\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_13", ["I\n|\nI\nI\n|\n|\n|\n|\n(5.VII)\nY\nX\nY\nX\n"]], ["block_14", ["I\n|\n|\nI\nl\nl\nI\n|\n(5 \nI\nI\nI\nI\n|\nI\nl\nl\n\u00b0\n)\n"]], ["block_15", ["Th reo-di-syndiotactic\n"]], ["block_16", ["Erythro-di-syndiotactic\n"]], ["block_17", ["Erythro-di-isotactic\n"]], ["block_18", ["Threo-di-isotactic\n"]]], "page_209": [["block_0", [{"image_0": "209_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The main conclusion we wish to draw from this line of development is that the difference between\nEi* and E3\u2018 could vary widely, depending on the nature of the active center. If the active center in a\npolymerization is a free radical unencumbered by interaction with any surrounding species, we\n"]], ["block_2", ["and, since the concentration dependences are identical, the relative rate of the two processes is\ngiven by the ratio of the rate constants. This same ratio also gives the relative number of dyads\nhaving the same or different configurations:\n"]], ["block_3", ["The Arrhenius equation enables us to expand on this still further:\n"]], ["block_4", ["1.\nThese are addition polymerizations in which chain growth is propagated through an active\ncenter. The latter could be a free radical or an ion; we shall see that a coordinated intermediate\nis the more usual case.\n2.\nThe active\u2014center chain end is open to front or rear attack in general; hence the configuration\nof a repeat unit is not fixed until the next unit attaches to the growing chain.\n3.\nThe reactivity of a growing chain is, as usual, assumed to be independent of chain length. In\nrepresenting this schematically, as either DM* or LM*, the M* indicates the terminal active\ncenter, and the D or L, the penultimate units of \ufb01xed con\ufb01guration. From a kinetic point of\nView, we ignore what lies further back along the chain.\n4.\nAs in Chapter 3 and Chapter 4, the monomer is represented by M.\n"]], ["block_5", ["What is significant about these reactions is that only two possibilities exist: addition with the\nsame configuration (D \u2014> DD or L 6 LL) or addition with the opposite configuration (D ~e> DL or\nL \u2014> LD). We shall designate these isotactic (subscript i) or syndiotactic (subscript s) additions,\nrespectively, and shall define the rate constants for the two steps k, and k5. Therefore the rates of\nisotactic and syndiotactic propagation become\n"]], ["block_6", ["and\n"]], ["block_7", ["gross, bulk properties provide qualitative evidence for differences in stereoregularity, but, as with\ncopolymers, itdetail that quantitatively characterizes the tacticity of a\npolymer. We shall examine the statistics of this situation in the next section, and the application\nof NMR in Section \nThe analogy between stereoregular polymers and copolymers can be extended still further. We\ncan write chemical equations for propagation reactions leading to products that differ in config-\nuration along with the rate laws. We do this without specifying anything\u2014\"at least for\nnow\u2014about the There are several things that need to be defined to do this:\n"]], ["block_8", ["present in a polymer. Since crystallinity plays such an important part in determining the mecha-\nnical properties itself in these other behaviors also. These\n"]], ["block_9", ["Characterizing \n195\n"]], ["block_10", ["RN _ k, _\nNumber dyads with same configuration\n(5 3)\nRP,S \nkS \nNumber dyads with different configurations\n'\n'\n"]], ["block_11", ["Rai k, [M*] [M]\n(5.7.1)\n"]], ["block_12", ["RP. ks[M*l[Ml\n(5.7.2)\n"]], ["block_13", ["ISO dyads\na \ufb02ewwf\u2014E\ufb02/RT\n(5.7.4)\nSyndio dyads\nF\nAS\n"]], ["block_14", ["DDM*\nLLM*\n\u2014\u2014DM* + M/\nor\n\u2014LM* + M\u201d\n(5.0)\n"]], ["block_15", ["With these definitions in mind, we can write\n"]], ["block_16", ["DLM*\n\\ LDMi\"\n"]]], "page_210": [["block_0", [{"image_0": "210_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Since it is unlikely that a polymer will possess perfect stereoregularity, it is desirable to assess\nthis pr0perty quantitatively, both to describe the polymer and to evaluate the effectiveness of\nvarious catalysts in this regard. In discussing tacticity in terms of microstructure, it has\nbecome conventional to designate a dyad as meso if the repeat units have the same configuration,\nand as racemic if the configuration is reversed. The terminology is derived from the stereochem\u2014\nistry of small molecules; its basis is seen by focusing attention on the methylene group in the\nbackbone of the vinyl polymer. This methylene lies in a plane of symmetry in the isotactic\nmolecule [5.X],\n"]], ["block_2", ["ki/k,s 0.63 at 60\u00b0C. The preference for syndiotactic addition is greater than this (i.e., Ef\u2018 E3 is\nlarger) in some systems, apparently because there is less repulsion between substituents when they\nare staggered in the transition state. In all cases, whatever difference in activation energies exists\nmanifests itself in product composition to a greater extent at low temperatures. At high temper-\natures small differences in E* value are leveled out by the high average thermal energy available.\nThe foregoing remarks refer explicitly to free-radical polymerizations. If the active center is\nsome kind of associated species\u2014an ion pair or a coordination complex\u2014then predictions based\non unencumbered intermediates are irrelevant. It turns out that the Ziegler\u2014Natta catalysts\u2014which\nwon their discoverers the Nobel Prize\u2014apparently operate in this way. The active center of the\nchain coordinates with the catalyst in such a way as to block one mode of addition. High levels of\nstereoregularity are achieved in this case. Although these substances also initiate the polymeriza-\ntion, the term catalyst is especially appropriate in the present context, since the activation energy\nfor one mode of addition is dramatically altered relative to the other by these materials. We shall\ndiscuss the chemical makeup of Ziegler\u2014Natta catalysts and some ideas about how they work in\nSection 5.10. For now it is sufficient to recognize that these catalysts introduce a real bias into\nEquation 5.7.4 and thereby favor one pattern of addition.\nIn the next section we take up the statistical description of various possible sequences.\n"]], ["block_3", ["would expect Ef\" E;k to be small. Experiment confirms this expectation; for vinyl chloride it is on\nthe order of 1.3 kJ mol\". Thus at the temperatures usually encountered in free-radical polymer-\nizations (ca. 60\u00b0C), the exponential in Equation 5.7.4 is small and the proportions of isotactic\nand syndiotactic dyads are roughly equal. This is the case for poly(vinyl chloride), for which\n"]], ["block_4", ["and thereby defines a meso (subscript m) structure as far as the dyad is concerned. Considering\nonly the dyad, we see that these two methylene protons are in different environments. Therefore\neach will show a different chemical shift in an NMR spectrum. In addition, each proton splits the\nresonance of the other into a doublet, so a quartet of peaks appears in the spectrum. Still\nconsidering only the dyad, we see that the methylene in a syndiotactic grouping [5.XI] contains\ntwo protons in identical environments:\nX i\n+0~|~\n(5.XI)\n.L x\n"]], ["block_5", ["These protons show a single chemical shift in the NMR spectrum. This is called a racemic\n(subscript r) structure, since it contains equal amounts of D and L character. In the next section\nwe shall discuss the NMR spectra of stereoregular polymers in more detail.\n"]], ["block_6", ["196\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_7", ["5.8\nA Statistical Description of Stereoregularity\n"]], ["block_8", ["X\nIII\nX\n+9+\n(5.X)\n"]], ["block_9", ["H\n"]]], "page_211": [["block_0", [{"image_0": "211_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "211_1.png", "coords": [20, 358, 301, 489], "fig_type": "figure"}]], ["block_2", [{"image_2": "211_2.png", "coords": [33, 320, 284, 575], "fig_type": "figure"}]], ["block_3", ["A Statistical Description of\n197\n"]], ["block_4", ["modes, respectively, then pm + pr possibilities. \nthe analog of pi,- for copolyrners; hence, by analogy with Equation 5.5.1, this equals the fraction of\nisotactic kinetic approach of the last \n"]], ["block_5", ["the rate rates of iso and syndio \n"]], ["block_6", ["This expression is the equivalent of Equation 5.5.2 for copolymers.\nThe system of notation we have defined can readily be extended to sequences of greater length.\nTable 5.6 illustrates how either In or r dyads can be bracketed by two additional repeat units to\nform a tetrad. Each of the outer units is either m or r with respect to the unit it is attached to, so the\nmeso dyad generates three tetrads. Note that the tetrads mmr and rmm are equivalent and are not\ndistinguished. A similar set of tetrads is generated from the r dyad.\nThe same system of notation can be extended further by focusing attention on the backbone\nsubstituents rather than on the methylenes. Consider bracketing a center substituent with a pair of\nmonomers in which the substituents have either the same or opposite configurations as the central\nsubstituent. Thus the probabilities of the resulting triads are obtained from the probabilities of the\nrespective m or r additions. The following possibilities exist:\n"]], ["block_7", ["Table 5.6\nThe Splitting of M680 and Racemic Dyads into Six Tetrads\n"]], ["block_8", ["3\nI\nA\n"]], ["block_9", ["Pm \nki +ks\n(5.8.1)\n"]], ["block_10", ["If we define pm and pr as the probability of addition occurring in the meso and racemic\n"]], ["block_11", ["X\nXX\nX\n"]], ["block_12", ["r\nlmlrlr'\nX\nx\n"]], ["block_13", [{"image_3": "211_3.png", "coords": [72, 492, 323, 688], "fig_type": "figure"}]], ["block_14", [{"image_4": "211_4.png", "coords": [108, 528, 295, 653], "fig_type": "figure"}]], ["block_15", [{"image_5": "211_5.png", "coords": [112, 334, 296, 694], "fig_type": "figure"}]], ["block_16", ["XX \nlrlmlrl\nX\nx\nXX\nr \n|m|r|m|\nXX\n"]], ["block_17", ["3..\u2014\n3\u2014\u2014\n"]], ["block_18", ["X\nX\n"]]], "page_212": [["block_0", [{"image_0": "212_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "212_1.png", "coords": [22, 588, 269, 660], "fig_type": "figure"}]], ["block_2", ["These triads can also be bracketed by two more units to generate 10 different pentads following\nthe pattem established in Table 5.6. It is left for the reader to verify this number by generating the\nvarious structures.\nThe probabilities of the various dyad, triad, and other sequences that we have examined have all\nbeen described by a single probability parameter pm. When we used the same kind of statistics for\ncopolymers, we called the situation one of terminal control. We are considering similar statistics\nhere, but the idea that the stereochemistry is controlled by the terminal unit is inappropriate. The\nactive center of the chain end governs the chemistry of the addition, but not the stereochemistry.\nEquation 5.7.1 and Equation 5.7.2 merely state that an addition must be of one kind or another, but\nthat the rates are not necessarily identical.\nA mechanism in which the stereochemistry of the growing chain does exert an influence on the\naddition might exist, but at least two repeat units in the chain are required to define any such\nstereochemistry. Therefore this possibility is equivalent to the penultimate mechanism in copoly-\nmers. In this case the addition would be described in terms of conditional probabilities, just as\nEquation 5.5.20 does for copolymers. Thus the probability of an isotactic triad controlled by the\nstereochemistry of the growing chain would be represented by the reaction.\n"]], ["block_3", ["and described by the probability\n"]], ["block_4", ["2.\nA syndiotactic triad (5.XIII) is generated by two successive racemic additions:\n"]], ["block_5", ["3.\nA heterotactic triad (5.XIV) is generated by mr and rm sequences of additions:\n"]], ["block_6", ["198\nCopolymers, Microstructure. and Stereoregularity\n"]], ["block_7", ["1.\nAn isotactic triad (5.XII) is generated by two successive meso additions:\n"]], ["block_8", [{"image_2": "212_2.png", "coords": [44, 181, 144, 225], "fig_type": "figure"}]], ["block_9", ["Pcontrol Pu/m\n(So-85)\n"]], ["block_10", ["X\nX\nXIXIX\nq_l_t.MH 'I'I'\n(5.1:)\nm\nImlml\n"]], ["block_11", [{"image_3": "212_3.png", "coords": [51, 73, 149, 119], "fig_type": "figure"}]], ["block_12", ["p, p3,,\n(5.3.2)\n"]], ["block_13", ["[\u20193 =(1_pm)2\n(5.8.3)\n"]], ["block_14", ["Ph 2pm(1\u2014pm)\n(5.8.4)\n"]], ["block_15", ["The probability of the isotactic triad is\n"]], ["block_16", ["The probability of the syndiotactic triad is given by pf, which becomes\n"]], ["block_17", ["The probability of a heterotactic (subscript h) triad is\n"]], ["block_18", ["The factor 2 arises because this particular sequence can be generated in two different orders.\n"]], ["block_19", ["I\nl\n|\nI\nI\n(5.x1n)\n"]], ["block_20", ["X\nJ\n"]], ["block_21", ["i\nl\ni\n|\n1\n(5.xn)\n"]], ["block_22", ["m\nm\n"]], ["block_23", ["X\n"]], ["block_24", ["X\n"]], ["block_25", [{"image_4": "212_4.png", "coords": [97, 594, 253, 635], "fig_type": "figure"}]], ["block_26", ["1\nX\n"]], ["block_27", ["(5 .XIV)\n"]]], "page_213": [["block_0", [{"image_0": "213_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "213_1.png", "coords": [22, 395, 274, 644], "fig_type": "figure"}]], ["block_2", ["Evaluate Equation 5.8.2 through Equation 5.8.4 for pm between zero and unity; these results are\nplotted in Figure 5.7.\n"]], ["block_3", ["Use zero-order Markov statistics to evaluate the probability of isotactic, syndiotactic, and\nheterotactic triads for the series of pm values spaced at intervals of 0.1. Plot and comment on\nthe results.\n"]], ["block_4", ["is exerted such situations are known, we shall limit our discussion to the\nsimple case where the single probability pm is sufficient to describe the various additions. The\nlatter, Markov (or Bernoulli) statistics to avoid the vocabu-\nlary of terminal control. The case where the addition is in\ufb02uenced by whether the last linkage in\nthe chain is m or r is followa first-order Markov process.\nThe number of m or r linkages in an \u201crt-ad\u201d is n\u2014~l. Thus dyads are characterized by a single\nlinkage (either In or r), triads by two linkages (either mm, mr, or rr), and so forth. The m and r\nnotation thus reduces by l order of the description from what is obtained when the repeat units\nthemselves are described. reason the terminal control mechanism for copolymers is a first-\norder Markov process and model is a second\u2014order Markov process. Note that\nthe compound probabilities which describe the probability of an n-ad in terms of pm are also of\norder n\u2014l. In the following example we calculate the probability of various triads on the basis\nof zero\u2014order Markov statistics.\n"]], ["block_5", ["ptriad\n\"\n-\n"]], ["block_6", ["A Statistical \n199\n"]], ["block_7", ["where the conditional is the probability of an m addition, given the fact ofa prior\nm addition. As with copolymers, triads must be considered in order to test whether the simple\nstatistics tobe examined to whether stereochemical \n"]], ["block_8", ["Figure 5.7\nFractions of iso, syndio, and hetero triads as a function of pm, calculated assuming zero\u2014order\nMarkov statistics in Example 5.6.\n"]], ["block_9", ["Example 5.6\n"]], ["block_10", ["Solution\n"]], ["block_11", ["0.8 \nps\nPi\n\u2014\n"]], ["block_12", ["0.5 \n\u2014\n"]], ["block_13", ["l'\n._\n0.4 F\n\u2014\n"]], ["block_14", ["0.2 \n._\n"]], ["block_15", ["0\n"]], ["block_16", ["1\nI\nI\nI\nI\nI\nI\nl\nI\nI\nI\nl\nI\nI\nI\nI\nI\n"]], ["block_17", ["0\n0.2\n0.4\n0 6\n0 8\n1\npm\n"]], ["block_18", ["_\nPh\n-\n"]], ["block_19", ["F\n"]], ["block_20", ["I\nI\nI\nI\nl\nl\nl\n1\nl\ni\nl\nl\nI\nl\nI\n"]], ["block_21", ["'\u2014\n"]]], "page_214": [["block_0", [{"image_0": "214_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "214_1.png", "coords": [26, 35, 196, 214], "fig_type": "figure"}]], ["block_2", ["0\n0\n0\n0\n0.1\n0.01\n0.81\n0.18\n0.2\n0.04\n0.64\n0.32\n0.3\n0.09\n0.49\n0.42\n0.4\n0.16\n0.36\n0.48\n0.5\n0.25\n0.25\n0.50\n0.6\n0.36\n0.16\n0.48\n0.7\n0.49\n0.09\n0.42\n0.8\n0.64\n0.04\n0.32\n0.9\n0.81\n0.01\n0.18\n1.0\n1.0\n0\n0\n"]], ["block_3", ["200\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_4", ["1.\nThe probabilities give the fractions of the three different types of triads in the polymer.\n2.\nIf the fractions of triads could be measured, they either would or would not lie on a single\nvertical line in Figure 5.7. If they did occur at a single value ofpm, this would not only give the\nvalue of pm (which could be obtained from the fraction of one kind of triad), but would also\nprove the statistics assumed. If the fractions were not consistent with a single pm value, higher-\norder Markov statistics are indicated.\n3.\nThe fraction of isotactic sequences increases as pm increases, as required by the definition of\nthese quantities.\n4.\nThe fraction of syndiotactic sequences increases as Pm -\u2014> 0, which corresponds to pr -\u2014> 1.\n5.\nThe fraction of heterotactic triads is a maximum at Pm Pr 0.5 and drops to zero at either\nextreme.\n6.\nFor an atactic polymer the proportions of isotactic, syndiotactic, and heterotactic triads are\n0.25 20.251050.\n"]], ["block_5", ["pm\npi]\n\u201c\u2014m\n2pm(1_pm)\n"]], ["block_6", ["The following observations can be made from these calculations:\n"]], ["block_7", ["To investigate the triads by NMR, the resonances associated with the chain substituent are\nexamined, since Structure (5.XII) through Structure (5.XIV) show that it is these that experience\ndifferent environments in the various triads. If dyad information is sufficient, the resonances of the\nmethylenes in the chain backbone are measured. Structure (5X) and Structure (5.XI) show that\nthese serve as probes of the environment in dyads. In the next section we shall examine in more\ndetail how this type of NMR data is interpreted.\n"]], ["block_8", ["It is not the purpose of this book to discuss in detail the contributions of NMR spectroscopy to\nthe determination of molecular structure. This is a specialized field in itself and a great deal\nhas been written on the subject. In this section we shall consider only the application of NMR\nto the elucidation of stereoregularity in polymers. Numerous other applications of this powerful\ntechnique have also been made in polymer chemistry, including the study of positional and\ngeometrical isomerism (Section\n1.6) and copolymers (Section 5.7). We shall also make no\nattempt to compare the NMR spectra of various different polymers; instead, we shall examine\nprimarily the NMR spectra of different poly(methyl methacrylate) preparations to illustrate the\ncapabilities of the method using the first system that was investigated by this technique as\nthe example.\nFigure 5.8 shows the 60 MHz spectra of poly(methy1 methacrylate) prepared with different\ncatalysts so that predominantly isotactic, syndiotactic, and atactic products are formed. The three\nspectra in Figure 5.8 are identified in terms of this predominant character. It is apparent that the\n"]], ["block_9", ["5.9\nAssessing Stereoregularity by Nuclear Magnetic Resonance\n"]]], "page_215": [["block_0", [{"image_0": "215_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "215_1.png", "coords": [30, 36, 281, 343], "fig_type": "figure"}]], ["block_2", ["1.\nHydrogens of the methylene group in the backbone of the poly(methy1 methacrylate) produce\na single peak in a racemic dyad, as illustrated by Structure (5.XIII).\n2.\nThe same group of hydrogens in a meso dyad (5.X) produces a quartet of peaks: two different\nchemical shifts, each split into two by the two hydrogens in the methylene.\n3.\nThe peaks centered at 5 1.84 ppm\u2014a singlet in the syndiotactic and a quartet in the isotactic\npolymers\u2014are thus identified with these protons. This provides an unambiguous identification\nof the predominant stereoregularity of these samples.\n4.\nThe features that occur near 5 1.0 ppm are associated with the protons of the (rt-methyl\ngroup. The location of this peak depends on the configurations of the nearest neighbors.\n5.\nWorking from the methylene assignments, we see that the peak at 5 1.22 ppm in the isotactic\npolymer arises from the methyl in the center of an isotactic triad, the peak at 5 0.87 ppm\nfrom a syndiotactic triad, and the peak at 5 1.02 ppm from a homotactic triad.\n6.\nThe peak at 5 3.5 ppm is due to the methoxy group.\n"]], ["block_3", ["Figure 5.8\nNuclear magnetic resonance spectra of three poly(methyl methacrylate) samples. Curves are\nlabeled according to the predominant tacticity of samples. (From McCall, D.W. and Slichter, W.P., in Newer\nMethods of Polymer Characterization, Ke, B. (Ed), Interscience, New York, 1964. With permission.)\n"]], ["block_4", ["Once these assignments are made, the areas under the various peaks can be measured to\ndetermine the various fractions:\n"]], ["block_5", ["spectra are quite different, especially in the range of 5 values between\n1 and 2 ppm. Since the\natactic polymer has the least regular structure, we concentrate on the other two to make the\nassignment of the spectral features to the various protons.\nSeveral observations from the last section provide the basis of interpreting these spectra:\n"]], ["block_6", ["Assessing \n201\n"]], ["block_7", ["1.\nThe area under the methylene peaks is proportional to the dyad concentration: The singlet\ngives the racemic dyads and the quartet gives the meso dyads.\n"]], ["block_8", ["1\n1\nl\n5\n4\n3\n2\n5 (ppm)\n"]], ["block_9", [{"image_2": "215_2.png", "coords": [53, 50, 270, 227], "fig_type": "figure"}]], ["block_10", ["Syndiotactic\n"]], ["block_11", ["lsotactic\n"]], ["block_12", ["Atactic\n"]], ["block_13", ["er-\nOl-\n"]]], "page_216": [["block_0", [{"image_0": "216_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Source: Data from McCall, D.W. and Slichter, WP. in Newer Methods of Potymer Characterization, Ke, B. (Ed),\nInterscience, New York, 1964.\n"]], ["block_2", ["Table 5.7\nThe Fractions of Meso and Racemic Dyads and Iso, Syndio, and Hetero Triads\nfor the Data in Figure 5.8\n"]], ["block_3", ["Atactic\n0.22\n0.78\n0.07\n0.55\n0.38\nSyndiotactic\n0.17\n0.83\n0.04\n0.68\n0.28\nIsotactic\n0.87\n0. 13\n0.89\n0.04\n0.07\n"]], ["block_4", ["Sample\nMeso\nRacemic\nIso\nSyndio\nHetero\n"]], ["block_5", ["1.\nThe number of isotactic sequences containing ni iso repeat units is Nni.\n2.\nThe number of syndiotactic sequences containing as syndio repeat units is \ufb01lms.\n"]], ["block_6", ["The spectra shown in Figure 5.8 were early attempts at this kind of experiment, and the\nmeasurement of peak areas in this case was a rather subjective affair. We shall continue with an\nanalysis of these spectra, even though improved instrumentation has resulted in greatly enhanced\nspectra. One development that has produced better resolution is the use of higher magnetic fields.\nAs the magnetic field increases, the chemical shifts for the various features are displaced propor-\ntionately. The splitting caused by spin\u2014spin coupling, on the other hand, is unaffected. This can\nproduce a considerable sharpening of the NMR spectrum. Other procedures such as spin decoup-\nling, isotopic substitution, computerized stripping of superimposed spectra, and 13C-NMR also\noffer methods for identifying and quantifying NMR spectra.\nTable 5.7 lists the estimated fractions of dyads of types m and r and the fractions of triads of\ntypes i, s, and h from Figure 5.8. These fractions represent the area under a specific peak (or four\npeaks in the case of the meso dyads) divided by the total area under all of the peaks in either the\ndyad or triad category. As expected for the sample labeled isotactic, 89% of the triads are of type i\nand 87% of the dyads are of type m. Likewise, in the sample labeled syndiotactic, 68% of the triads\nare s and 83% of the dyads are r.\nThe sample labeled atactic in Figure 5.8 was prepared by a free\u2014radical mechanism and is\nexpected to follow zero-order Markov statistics. As a test for this, we examine Figure 5.7 to see\nwhether the values of 19,, p3, and ph, which are given by the fractions in Table 5.7, agree with a\nsingle set of Pm values. When this is done, it is apparent that these proportions are consistent with\nthis type of statistics within experimental error and that pm \u201c=V 0.25 for poly(methyl methacrylate).\nUnder the conditions of this polymerization, the free-radical mechanism is biased in favor of\nsyndiotactic additions over isotactic additions by about 3:1, according to Equation 5.8.1. Presum-\nably this is due to steric effects involving the two substituents on the oc-carbon.\nWith this kind of information it is not difficult to evaluate the average lengths of isotactic\nand syndiotactic sequences in a polymer. As a step toward this objective, we define the\nfollowing:\n"]], ["block_7", ["3.\nSince isotactic and syndiotactic sequences must alternate, it follows that:\n"]], ["block_8", ["2.\nThe area under one of the methyl peaks is proportional to the concentration of the correspond-\ning triad.\n3.\nIt is apparent that it is not particularly easy to determine the exact areas of these features when\nthe various contributions occur together to any significant extent. This is clear from the atactic\nspectrum, in which slight shoulders on both the methylene and methyl peaks are the only\nevidence of meso methylenes and iso methyls.\n"]], ["block_9", ["202\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_10", [{"image_1": "216_1.png", "coords": [46, 489, 133, 523], "fig_type": "molecule"}]], ["block_11", ["EN!\u201c :a\n(5.9.1)\n"]], ["block_12", ["Dyads\nTriads\n"]]], "page_217": [["block_0", [{"image_0": "217_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4,\nThe number of iso triads in a sequence of n, iso repeat units is ni\u2014 l, and the number of syndio\ntriads in a sequence of ns syndio repeat units is rig\u20141. We can verify these relationships by\nexamining a specific chain segment:\n"]], ["block_2", ["Assessing \n203\n"]], ["block_3", ["5.\nThe number of racemic dyads in a sequence is the same as the number of syndiotactic units (is.\nThe number of meso dyads in a sequence is the same as the number of iso units In. These can\nalso be verified from structure above.\n"]], ["block_4", ["Use the dyad and triad fractions in Table 5.7 to calculate the average lengths \nsyndiotactic sequences for the polymers of Figure 5.8. Comment on the results.\n"]], ["block_5", ["Example 5.7\n"]], ["block_6", [{"image_1": "217_1.png", "coords": [39, 222, 252, 276], "fig_type": "molecule"}]], ["block_7", [{"image_2": "217_2.png", "coords": [46, 411, 154, 455], "fig_type": "molecule"}]], ["block_8", ["In this example both the iso and syndio sequences consist of eight repeat units, with seven\ntriads in each. The repeat unit marked is counted as part of each type of triad, but is itself the\ncenter of a hetero triad.\n"]], ["block_9", ["In this equation the summations are over all values of n of the specified type. Also remember\nthat the V\u2019s and n\u2019s in this discussion (with subscript i or s) are de\ufb01ned differently from the 12\u2019s\nand n\u2019s defined earlier in the chapter for 00polymers. Using Equation 5.9.1 and remembering\nthe definition of an average provided by Equation 1.7.7, we see that Equation 5.9.2 becomes\n"]], ["block_10", ["where the overbar indicates the average length of the indicated sequence.\n"]], ["block_11", ["With these definitions in mind, we can immediately write expressions for the ratio of the\ntotal number of iso triads, vi, to the total number of syndio triads, vs:\n"]], ["block_12", [" 2 _ \n(5 9 2)\nVs\u2014Za(ns_1)mZN\u201d5(nS)\u2014ZNHS\n.\nI\n"]], ["block_13", ["5 \n(5.9.3)\n"]], ["block_14", ["Equation 5.9.3 and Equation 5.9.4 can be solved simultaneously for E, and rig in terms of the\ntotal number of dyads and triads:\n"]], ["block_15", ["\u2014DDLDLDLDLD*DDDDDDDDL-\n"]], ["block_16", ["A similar result can be written for the ratio of the total number (12) of dyads of the two types\n(m and I), using item (5) above:\n"]], ["block_17", ["and\n"]], ["block_18", ["Use of these relationships is illustrated in the following example.\n"]], ["block_19", ["_ _\n1 Vi/Vs\nHi \n"]], ["block_20", ["a, \n1 \nVi/VS\n(5.9.6)\n(Vim/VI) (Vi/Vs)\n"]], ["block_21", ["VS\n\ufb01s\ufb02l\n"]], ["block_22", ["pm\n2mm # \ufb01r\n(5.9.4)\n:7.\n\u2014 22mm.) \u201c\n(is\n"]], ["block_23", ["1\u2014(Vi/Vs)(Vr/Vm)\n(59.5)\n"]]], "page_218": [["block_0", [{"image_0": "218_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "218_1.png", "coords": [30, 472, 349, 601], "fig_type": "figure"}]], ["block_2", ["Figure 5.9\n13C-NMR assignments for polypropylene. (From Bruce, MD. and Waymouth, R.M., Macro-\nmolecules, 31, 2707, 1998. With permission.)\n"]], ["block_3", ["Thus the fractions in Table 5.7 can be substituted for the 12\u2019s in Equation 5.9.3 and Equation 5.9.4.\nThe values of \ufb01i and fl, so calculated for the three polymers are:\n"]], ["block_4", ["We conclude this section via Figure 5.9, which introduces the use of l3C-NMR obtained at 100\nMHz for the analysis of stereoregularity in polypropylene. This spectrum shows the carbons\non the pendant methyl groups for an atactic polymer. Individual peaks are resolved for all the\npossible pentad sequences. Polypropylene also serves as an excellent starting point for the next\nsection, in which we examine some of the catalysts that are able to control stereoregularity in\nsuch polymers.\n"]], ["block_5", ["This analysis adds nothing new to the picture already presented by the dyad and triad probabilities.\nIt is somewhat easier to visualize an average sequence, however, although it must be remembered\nthat the latter implies nothing about the distribution of sequence lengths.\n"]], ["block_6", ["Atactic\n1 .59\n5.64\nSyndiotactic\n1.32\n6.45\nIsotactic\n9.14\n1.37\n"]], ["block_7", ["Since the total numbers of dyads and triads always occur as ratios in Equation 5.9.3 and Equation\n5.9.4, both the numerators and denominators of these ratios can be divided by the total number of\ndyads or triads to convert these total numbers into fractions, i.e.,\n"]], ["block_8", ["204\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_9", ["Solution\n"]], ["block_10", [{"image_2": "218_2.png", "coords": [39, 440, 189, 607], "fig_type": "figure"}]], ["block_11", ["I\nl\nI\nI\nI\nI\nI\nF\nI\nI\nI\nI\nl\nI\nI\nl\nl\u2014r\nI\u2014l\nI\u2014I\n1\u2014l\n1\nI_|\n22\n21.5\n21\n20.5\n20\n19.5\n"]], ["block_12", [{"image_3": "218_3.png", "coords": [46, 363, 341, 637], "fig_type": "figure"}]], ["block_13", ["Vi/Vs (Vi/Vtot)/(Vs/Vtot) Pi/Ps\n"]], ["block_14", ["\u201di\n\u201ds\n"]], ["block_15", [{"image_4": "218_4.png", "coords": [112, 393, 294, 588], "fig_type": "figure"}]], ["block_16", [{"image_5": "218_5.png", "coords": [118, 379, 251, 418], "fig_type": "molecule"}]]], "page_219": [["block_0", [{"image_0": "219_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Since the coordination almost certainly involves the transition metal atom, there is a resemblance\nhere to anionic polymerization. The coordination is an important aspect of the present picture,\nsince it is this feature that allows the catalyst to serve as a template for stereoregulation.\nThe assortment of combinations of components is not the only variable to consider in describing\nZiegler\u2014Natta catalysts. Some other variables include the following:\n"]], ["block_2", ["Among other possibilities in these reactions, these free radicals can initiate ordinary free-radical\npolymerization. The Ziegler\u2014Natta systems are thus seen to encompass several mechanisms for the\ninitiation of polymerization. Neither ionic nor free-radical mechanisms account for stereoregular-\nity, however, so we must look further for the mechanism whereby the Ziegler\u2014Natta systems\nproduce this interesting effect.\nThe stereoregulating capability of Ziegler\u2014Natta catalysts is believed to depend on a coordin-\nation mechanism in which both the growing polymer chain and the monomer coordinate with the\ncatalyst. The addition then occurs by insertion of the monomer between the growing chain and the\ncatalysts by a concerted mechanism (5.XV):\n"]], ["block_3", ["In this discussion we consider Ziegler\u2014Natta catalysts and their role in achieving stereoregularity.\nThis is a somewhat of the situation, since there are other catalysts\u2014such as phenyl\nmagnesium bromide, a Grignard reagent\u2014which can produce stereoregularity; the Ziegler\u2014Natta\ncatalysts polymers\u2014unbranched polyethylene to name one\u2014which lack\nstereoregularity. catalysts are historically the most widely used and best\u2014\nunderstood stereoregulating systems, so the loss of generality in this approach is not of great\nconsequence.\nThe fundamental Ziegler\u2014Natta recipe consists of two components: a halide or other compound\nof a transition metal from among the group IVB to VIIIB elements, and an organometallic\ncompound of metal from groups IA to IIIA. Some of the transition metal\ncompounds studied include TiCl4, TiCl3, VC14, VC13, ZrCl4, Cl'Cl3, MoCls, and CuCl. Represen-\ntative organometallics include (C2H5)3A1, (C2H5)2Mg, C4H9Li, and (C2H5)22n. These are only a\nfew of the possible compounds, so the number of combinations is very large.\nThe individual components of the Zieglerm-Natta system can separately account for the initiation\nof some forms of polymerization reactions, but not for the fact of stereoregularity. For example,\nbutyl lithium can initiate anionic polymerization (see Section 4.3) and TiCl4 can initiate cationic\npolymerization (see Section 4.5). In combination, still another mechanism for polymerization,\ncoordination polymerization, is indicated. When the two components of the Ziegler\u2014Natta system\nare present together, complicated exchange reactions are possible. Often the catalyst must \u201cage\u201d to\nattain maximum effectiveness; presumably this allows these exchange reactions to occur. Some\npossible exchange equilibria are\n"]], ["block_4", ["The organotitanium halide can then be reduced to TiCl3:\n"]], ["block_5", ["Ziegler-Natta \n205\n"]], ["block_6", ["5.10\nZiegler\u2014Natta Catalysts\n"]], ["block_7", ["H\nH\nX\n_(:3_\\/*\u2018/\\H2\n(5.XV)\n"]], ["block_8", ["2A16 e [A1(C2H5>2]+[A1(C2H5>4]\u2018\nTiCl4 + [A1(C2H5)2]+ 4:) C2H5TiC13 + [A1(C2H5)Cl]+\n(5.Q)\n"]], ["block_9", ["C2H5TiC13 \u2014> TiCl3 + Cs'\n(5.R)\n"]], ["block_10", [{"image_1": "219_1.png", "coords": [50, 549, 105, 595], "fig_type": "molecule"}]], ["block_11", ["X\nCat\n"]]], "page_220": [["block_0", [{"image_0": "220_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1.\nCatalyst solubility. Polymerization systems may consist of one or two phases. Titanium-based\ncatalysts are the most common of the heterogeneous systems; vanadium-based catalysts are\nthe most common homogeneous systems. Since the catalyst functions as a template for the\nformation of a stereoregular product, it follows that the more extreme orienting effect of\nsolid surface (i.e., heterogeneous catalysts) is required for those monomers that interact\nonly weakly with the catalyst. The latter are nonpolar monomers. Polar monomers interact\nmore strongly with catalysts, and dissolved catalysts are able to exert sufficient control for\nStereoregularity.\n2.\nCrystal structure of solids. The a-crystal form of TiCl3 is an excellent catalyst and has been\ninvestigated extensively. In this particular crystal form of TiCl3, the titanium ions are located\nin an octahedral environment of chloride ions. It is believed that the stereoactive titanium ions\nin this crystal are located at the edges of the crystal, where chloride ion vacancies in the\ncoordination sphere allow coordination with the monomer molecules.\n3.\nTacticity of products. Most solid catalysts produce isotactic products. This is probably because\nof the highly orienting effect of the solid surface, as noted in item (1). The preferred isotactic\nconfiguration produced at these surfaces is largely governed by steric and electrostatic\ninteractions between the monomer and the ligands of the transition metal. Syndiotacticity is\nmostly produced by soluble catalysts. Syndiotactic polymerizations are carried out at low\ntemperatures, and even the catalyst must be prepared at low temperatures; otherwise specifi\u2014\ncity is lost. With polar monomers syndiotacticity is also promoted by polar reaction media.\nApparently the polar solvent molecules compete with monomer for coordination sites, and\nthus indicate more loosely coordinated reactive species.\n4.\nRate of polymerization. The rate of polymerization for homogeneous systems closely resem-\nbles anionic polymerization. For heterogeneous systems the concentration of alkylated tran-\nsition metal sites on the surface appears in the rate law. The latter depends on the particle size\nof the solid catalyst and may be complicated by sites of various degrees of activity. There is\nsometimes an inverse relationship between the degree of Stereoregularity produced by a\ncatalyst and the rate at which polymerization occurs.\n"]], ["block_2", ["The catalysts under consideration both initiate the polymerization and regulate the polymer\nformed. There is general agreement that the mechanism by which these materials exert their\nregulatory role involves coordination of monomer with the transition metal atom, but proposed\ndetails beyond this are almost as numerous and specific as the catalysts themselves. We shall return\nto a description of two specific mechanisms below. The general picture postulates an interaction\nbetween monomer and catalyst such that a complex is formed between the qr electrons of the olefin\nand the d orbitals of the transition metal. Figure 5.10 shows that the overlap between the \ufb01lled\norbitals of the monomer can overlap with vacant dx2_y2 orbitals of the metal. Alternatively, hybrid\norbitals may be involved on the metal. There is a precedent for such bonding in simple model\ncompounds. It is known, for example, that Pt2+ complexes with ethylene by forming a dsp2\nhybridnqr sigma bond and a dp hybrid~rr* pi bond. A crucial consideration in the coordination is\nmaximizing the overlap of the orbitals involved. Titanium(III) ions seem ideally suited for this\n"]], ["block_3", ["206\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_4", ["(a)\n(b)\n"]], ["block_5", [{"image_1": "220_1.png", "coords": [35, 557, 341, 648], "fig_type": "figure"}]], ["block_6", ["Figure 5.10\nPossible orbital overlaps between a transition metal and an olefin.\n"]], ["block_7", [{"image_2": "220_2.png", "coords": [43, 563, 150, 638], "fig_type": "molecule"}]], ["block_8", ["dx2 y2\n"]], ["block_9", ["a\nc\nG>\nl\n1\u201d\n\u20ac09\n050/\"Q M .\\n*\n\\6\nC (3/\n"]], ["block_10", [{"image_3": "220_3.png", "coords": [168, 556, 319, 658], "fig_type": "figure"}]], ["block_11", [{"image_4": "220_4.png", "coords": [180, 593, 339, 650], "fig_type": "molecule"}]]], "page_221": [["block_0", [{"image_0": "221_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["function; higher effective nuclear charge on the metal results in less spatial extension of d orbitals\nand diminished overlap.\nMany mechanisms have been proposed that elaborate on this picture. These are often so specific\nthat they cannot be generalized beyond the systems for which they are proposed. Two schemes that\ndo allow some generalization are presented here. Although they share certain common features,\nthese mechanisms are distinguished by the fact that one\u2014the monometallic model\u2014does not\ninclude any participation by the representative metal in the mechanism. The second\u2014the bimetallic\nmodel\u2014does assume the involvement of both metals in the mechanism.\nThe monometallic mechanism is illustrated by Figure 5.11. It involves the monomer coordin-\nating with an alkylated titanium atom. The insertion of the monomer into the titanium\u2014carbon bond\npropagates the chain. As shown in Figure 5.11 this shifts the vacancy\u2014represented by the\nsquare\u2014in the coordination sphere of the titanium to a different site. Syndiotactic regulation\noccurs if the next addition takes place via this newly created vacancy. In this case the monomer and\nthe growing chain occupy alternating coordination sites in successive steps. For the more common\nisotactic growth the polymer chain must migrate back to its original position.\nThe bimetallic mechanism is illustrated in Figure 5.12; the bimetallic active center is the\ndistinguishing feature of this mechanism. The precise distribution of halides and alkyls is not\nspelled out because of the exchanges described by Reaction (5Q). An alkyl bridge is assumed\nbased on observations of other organometallic compounds. The coordination of the olefin with\nthe titanium is followed by insertion of the monomer into the bridge to propagate the reaction.\nAt present it is not possible to determine which of these mechanisms or their variations most\naccurately represents the behavior of Ziegler\u2014Natta catalysts. In view of the number of variables in\n"]], ["block_2", [",r\u2019\nu\n\u201c1,0\nWCIH\nMe\n\"To\n:,Cl\nMe\nMe\n,0\nol-,Ti--EI \nC'7Tui\"'\u201c=\n\u2014-\u2014--\nCI-,Ti---~CH2 \u2014-\nC|7Ti\n\u2014\u00bb-\nCI-,Ti--n\n0'83.\nC'CI\nH20\n0'63.\n0' ('3.\n0'33\n"]], ["block_3", ["Ziegler\u2014Natta \n207\n"]], ["block_4", ["Polypropylene polymerized with triethyl aluminum and titanium trichloride \ncontain various kinds of chain ends. Both terminal vinylidene unsaturation and aluminum-bound\nchain ends have been identified. Propose two termination reactions to account \ntions. Do the termination reactions allow any discrimination between the monometallic \nbimetallic propagation mechanisms?\n"]], ["block_5", ["these catalyzed polymerizations, both mechanisms may be valid, each for different specific \nIn the following example the termination step of coordination polymerizations is considered.\n"]], ["block_6", ["Figure 5.11\nMonometallic mechanism. The square indicates a vacant ligand site.\n"]], ["block_7", ["HTMe\ng\nH+\nMe\nH\nMe\nMew/g\nI\n\u2019,.C\"'2\n3515+ H202\nH2C/\nEC\"\n9\n/Ti:~\n:\u2018AI/\n\u2014\u20141-\u00bb\n/Ti\\\n:Al/\ni/\n\u2014\u2014\u2014I-\u2014 \\ ., x I\u201c-\nI\n\u201d\u201c- ,\u2019\n\\\nI\n\u2018\n, \n\\\nTl\u201c\n\u2019Al\n/TI\u00ab.\n:AI\nR\nR\n/\n\u201cFl\u2019\n\\\nTR\u201d\n"]], ["block_8", ["Example 5.8\n"]], ["block_9", ["Figure 5.12\nThe bimetallic mechanism.\n"]], ["block_10", [{"image_1": "221_1.png", "coords": [85, 564, 458, 665], "fig_type": "figure"}]], ["block_11", [{"image_2": "221_2.png", "coords": [98, 63, 440, 121], "fig_type": "molecule"}]], ["block_12", [{"image_3": "221_3.png", "coords": [139, 589, 339, 645], "fig_type": "molecule"}]], ["block_13", [{"image_4": "221_4.png", "coords": [205, 577, 448, 650], "fig_type": "molecule"}]], ["block_14", [{"image_5": "221_5.png", "coords": [207, 58, 334, 123], "fig_type": "molecule"}]], ["block_15", ["Me\n"]], ["block_16", [{"image_6": "221_6.png", "coords": [288, 57, 430, 124], "fig_type": "molecule"}]]], "page_222": [["block_0", [{"image_0": "222_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "222_1.png", "coords": [25, 212, 258, 265], "fig_type": "figure"}]], ["block_2", ["These reactions appear equally feasible for titanium in either the monometallic or bimetallic\nintermediate. Thus they account for the different types of end groups in the polymer, but do not\ndifferentiate between propagation intermediates. In the commercial process for the production of\npolypropylene by Ziegler\u2014Natta catalysts, hydrogen is added to terminate the reaction, so neither\nof these reactions is pertinent in this case.\n"]], ["block_3", ["The discussion in the preceding section indicates that Ziegler\u2014Natta catalysts represent a rather\ncomplicated subject. This complexity is often re\ufb02ected in the structure of the polymers produced.\nFor example, the different ways that the two metal centers may or may not interact during an\naddition step suggest that there are, in fact, multiple catalytic sites active in a given polymerization.\nThis can lead to sites with greatly different propagation rates, different stereoselectivity, and\ndifferent propensities to incorporate any comonomers present. The net result is that polymer\nmaterials produced by Ziegler\u2014Natta catalysts, eSpecially under commercial conditions, tend to\nbe highly heterogeneous at the molecular level. A broad strategy to overcome this limitation is\nbased on the concept of a single-site catalyst, i.e., one that has a single, well-defined catalytic\ngeometry that can control the desired aspect of prOpagation. In this section, we brie\ufb02y consider\nsome examples of such catalysts for stereochemical control in the polymerization of a-olefins. We\nbegin with a little more consideration of catalysis in general.\nThe majority of catalysts in commercial use are heterogeneous. In this usage, the term\nheterogeneous means that the phase of the catalyst (e.g., solid) is distinct from that of the reagents\nand products (usually gases and liquids). When the catalyst is a relatively small molecule, it is\nretained in the solid phase by immobilization on some kind of inert, robust support. The reaction of\ninterest therefore takes place at the solid\u2014liquid or solid\u2014gas interface. The fact of immobilization\ncan itself contribute to the multiple site nature of heterogeneous catalysts, for example by exposing\ndifferent faces of the catalytically active metal center, by restricting accessibility of reagents to\ncatalyst particles deep within a porous support, and by presenting a distribution of different cluster\nsizes of catalytic particles. Given these disadvantages, one might ask why heterogeneous catalysis\nis the norm. The answer is simple: It is much easier to separate (and possibly regenerate)\nheterogeneous catalysts from products and unreacted reagents. Note that if the activity of a catalyst\nis sufficiently high (in terms of grams polymer produced per gram catalyst employed), then\n"]], ["block_4", ["The transfer of a tertiary hydrogen between the polymer chain and a monomer can account for the\nvinylidene group in the polymer:\n"]], ["block_5", ["A reaction analogous to the alkylation step of Reaction (4.Q) can account for the association of an\naluminum species with chain ends:\n"]], ["block_6", ["208\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_7", ["5.11\nSingle-Site Catalysts\n"]], ["block_8", ["Solution\n"]], ["block_9", [{"image_2": "222_2.png", "coords": [37, 213, 234, 261], "fig_type": "molecule"}]], ["block_10", ["Ti/W\u2018J\n+ MeVAIVMe \u2014\u2014\u2014\u2014)- Me\\./\nH\nMe\n:W\nMeH\n"]], ["block_11", ["+\n"]], ["block_12", [{"image_3": "222_3.png", "coords": [156, 210, 356, 273], "fig_type": "molecule"}]], ["block_13", [{"image_4": "222_4.png", "coords": [197, 224, 352, 259], "fig_type": "molecule"}]], ["block_14", ["Me\n./\\/Me\n(5.81\u00bb)\nMe\nT'\n+ ACHQ\n"]], ["block_15", ["-->+\n(5.8a)\n"]], ["block_16", ["HMe+\nIVMe \u2014:-- \ufb02AK/AIVMe + TiAMe\n"]]], "page_223": [["block_0", [{"image_0": "223_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["as chloride.\nThe choice of metal, ligands, and design of the overall constraining geometry provide a rich\npalette from which catalysts may be designed. In general, the stereoregulation of monomer\naddition can be achieved through one of two modes. Under chain\u2014end control, the addition of a\nmonomer is influenced mostly by the configuration of the previous repeat unit, which is reminis-\ncent of the terminal model of copolymerization. To appreciate how this can happen, it is important\nto realize that the growing polymer remains bound to the metal center during the addition step.\nAlternatively, under Site control the ligand set may be chosen to provide a chiral confining\nenvironment, which exerts a dominant in\ufb02uence on the stereochemistry of addition. The symmetry\nof the catalyst is often strongly correlated with the mode and effect of stereocontrol. This is\nsummarized in Figure 5.13. Catalysts with a plane of symmetry, or CS, tend to produce syndiotactic\npolymers under site control, but either iso\u2014 or syndiotactic polymers under chain-end control.\nWhen the symmetry is C2, i.e., identical after rotation by 180D about a single axis, the addition is\n"]], ["block_2", ["However, this representation is not complete. Just as Ziegler\u2014Natta catalysts always involve a\nmixture of at least two active ingredients, single\u2014site catalysts involve another component. The\nmost common is a partially hydrolyzed trimethyl aluminum species\u2014methylaluminoxane (MAO).\nThe active center is more properly denoted [LnMP]+[X]_, where the metal site is cationic by virtue\nof being coordinatively unsaturated, and the counterion contains MAO and a displaced ligand, such\n"]], ["block_3", ["Figure 5.13\nRole of catalyst symmetry in stereocontrol. The open square represents the unsaturated site for\nmonomer addition, and the Cp rings are represented by the pendant lines. A catalyst of type (a) is isospecific\nand (b) is syndiospeci\ufb01c, when under site control; (c) and (d) can be either 130- or syndiospecific, under chain-\nend control. (From Coates, G.W., Chem. Rev., 100, 1223, 2000.)\n"]], ["block_4", ["separation and recovery may not be necessary. In contrast, single-site polymeriza-\ntion catalysts are usually they are molecularly dispersed within the reaction\nmedium. This situation leads to better defined products, and is much more amenable to detailed\nstudies of mechanism. strategies for immobilizing such catalysts are available,\nmaking them also of potential commercial interest.\nMost single-site catalysts have the general formula [LnMP], where L,., represents a set of\nligands, M is the active center, and P is the growing polymer. Furthermore, a common\nmotif is for two of the ligands to contain cyclopentadienyl (Cp) rings, which may themselves be\ncovalently linked or bridged. The example shown below (5.XVI) was one of the first such\nmetallocene systems and produces highly isotactic polypropylene.\n"]], ["block_5", ["I\nl\nI\n(a)\nP\n(b)\nP\n(C)\nP\n(d)\nP\n"]], ["block_6", ["Single-Site Catalysts\n209\n"]], ["block_7", ["C \nVol)\nV\nC 0\np\nDEM\nmaul/l\nDEM\np\nElm-M\np\n"]], ["block_8", ["Cp\nCD\nCp\nCp\nCp\nCp\n"]], ["block_9", [{"image_1": "223_1.png", "coords": [115, 533, 360, 603], "fig_type": "molecule"}]], ["block_10", [{"image_2": "223_2.png", "coords": [261, 535, 351, 599], "fig_type": "molecule"}]], ["block_11", ["(5.xv1)\n"]]], "page_224": [["block_0", [{"image_0": "224_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "224_1.png", "coords": [4, 51, 297, 126], "fig_type": "figure"}]], ["block_2", [{"image_2": "224_2.png", "coords": [19, 42, 427, 151], "fig_type": "figure"}]], ["block_3", [{"image_3": "224_3.png", "coords": [28, 62, 203, 116], "fig_type": "figure"}]], ["block_4", ["As indicated by the double arrows, the catalyst actually oscillates between two isomeric structures.\nThe structure on the left is chiral with C2 symmetry, and gives isotactic polypropylene (note that\nthe chloride ligands are not in the plane of the page). The structure on the right, however, is achiral,\n"]], ["block_5", ["iSOSpecific under site control. When a further mirror plane exists, in C2,, symmetry, chain-end\ncontrol leads to either iso- or syndiospecific addition.\nWe now illustrate these phenomena with two particular catalysts and a cartoon sketch of\nthe mechanism of monomer insertion. The monomer in question is polypropylene, the commer-\ncially most important stereogenic polyolefin and the most studied model system. However, it\nshould be noted that the \ufb02exibility of design for single-site catalysts offers the possibility of\nmore tolerance toward monomer polarity or functionality than in the Ziegler\u2014Natta analogs,\nthereby enabling stereocontrol of many different monomers or comonomers. The catalyst\n(5.XVI) has C2 symmetry and is isospecific under site control. The mechanism is illustrated in\nFigure 5.14, where for simplicity the Cp-containing ligands are represented by horizontal lines.\nThe polymer chain is bound to the metal through the unsubstituted backbone carbon and the\norientation of the incoming monomer is in\ufb02uenced by the location of the Cp ligand. In the\ntransition state the unsubstituted carbon of the new monomer coordinates with the metal and will\nbecome the new terminal carbon of the growing chain. A key role is thought to be played by a\nso-called \u201ca-agostic\u201d interaction between the metal and the hydrogen on the terminal carbon of\nthe polymer chain, which stabilizes the particular geometry of the transition state. After the\nincorporation of the monomer, the polymer chain (or a last few repeat units thereof) has\n\u201c\ufb02ipped\u201d to the other side of the metal center, in a process which is often compared to the\naction of a windshield wiper.\nIn contrast, the following zirconocene (5.XVII) is syndiospecific, consistent with its C,\nsymmetry. The mechanism is analogous to that illustrated in Figure 5.14, except that the inversion\nof the position of the bulky ligand inverts the preferred orientation of the incoming monomer.\n"]], ["block_6", ["The range of possibilities afforded by this class of catalysts is vast. As one last example,\nconsider the following zirconocene (5.XVIII), developed by Coates and Waymouth [2]:\nm=6\u201d\nZrCI2\nZrCI2\n(5.XVIII)\n"]], ["block_7", ["210\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_8", ["Figure 5.14\nProposed mechanism of iSOSpecific polymerization of polypropylene. (From Coates, G.W.,\nChem. Rev, 100, 1223, 2000.)\n"]], ["block_9", ["c\nH /\nM~.\nH\nH H\nH\nH\nH\nH/L\u201c/H M\nH\nHH/\u00e9\nW\u201c\nJk\nH\nH\n|\u2014\n\u2019\nMe\nMe\nMe HMe\nH\nMe\u2014H\n"]], ["block_10", [{"image_4": "224_4.png", "coords": [37, 535, 314, 618], "fig_type": "figure"}]], ["block_11", [{"image_5": "224_5.png", "coords": [37, 450, 110, 502], "fig_type": "molecule"}]], ["block_12", [{"image_6": "224_6.png", "coords": [41, 575, 146, 617], "fig_type": "molecule"}]], ["block_13", [{"image_7": "224_7.png", "coords": [170, 56, 289, 120], "fig_type": "figure"}]], ["block_14", [{"image_8": "224_8.png", "coords": [191, 52, 429, 123], "fig_type": "figure"}]], ["block_15", [{"image_9": "224_9.png", "coords": [277, 46, 423, 138], "fig_type": "molecule"}]], ["block_16", [{"image_10": "224_10.png", "coords": [287, 75, 411, 116], "fig_type": "molecule"}]], ["block_17", ["(5.xvn)\n"]]], "page_225": [["block_0", [{"image_0": "225_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This chapter has covered a broad range of issues relating to the structure of polymer chains at the\nlevel of a few repeat units. The two main topics have been copolymerization and stereoregularity.\nThese topics share many features in common, including (i) the importance of the relative reactivity\nof a growing chain end to addition of a particular monomer, or a monomer in a particular\nconfiguration; (ii) the use of statistics in describing composition, average sequence lengths, and\nsequence length distribution; (iii) the central role of spectroscopic methods, and especially NMR,\nin characterizing structural details.\n"]], ["block_2", ["and actually leads to random stereochemistry, i.e., atactic polypropylene. Now, consider the\ninteresting situation the of monomer insertion is more rapid than the rate of exchange\nbetween factor of 20. suchthe \n\u201cstereoblock copolymer,\u201d with alternating sequences of isotactic and atactic polypropylene, where\nthe average sequence length would be about 20. Such a polymer has some very appealing\nproperties. The isotactic blocks can crystallize, as will be discussed in detail in Chapter 13,\nwhereas the atactic blocks cannot. The result is that for temperatures above the glass transition\nof the atactic block (about \u201410\u00b0C, see Chapter 12) but below the melting temperature of the\nstereoregular block (about 140\u00b0C) the material acts as a crosslinked elastomer (see Chapter 10).\nThe crystallites tie the different molecules together, imparting mechanical strength, but the atactic\nblocks can be stretched appreciably without breaking, like a rubbery material. The mechanical\nresponse is sensitive to the relative lengths of the two blocks, which can be tuned through\nmonomer concentration and polymerization temperature. The result is an appealing situation in\nwhich an inexpensive monomer can be used to produce a variety of different products by\nstraightforward modi\ufb01cation to the reaction conditions.\n"]], ["block_3", ["1.\nThe key parameters in copolymerization are the reactivity ratios, which in\ufb02uence the relative\nrates at which a given radical will add the same monomer versus a comonomer. Thus a given\nreactivity ratio is specific to a particular pair of monomers, and copolymerization of two\nmonomer system requires specification of two reactivity ratios.\n2.\nThe copolymerization equation relates the mole fraction of monomers in polymer to the\ncomposition of the feedstock via the reactivity ratios. Different classes of behavior may be\nassigned based on the product of the reactivity ratios, including an \u201cideal\u201d copolymerization\nwhen the two reactivity ratios are reciprocals of one another.\n3.\nThe relative magnitudes of reactivity ratios can be understood, at least qualitatively, by\nconsidering the contributions of resonance stabilization, polarity differences, and possible\nsteric effects.\n4.\nStatistical considerations give predictions for the average sequence length and sequence \ndistributions in a copolymer on the basis of reactivity ratios and feedstock composition.\nHowever, the probability of adding a given monomer to a growing chain end may be determined\n"]], ["block_4", ["by the last, the last plus next-to-last, or even the last, next-to-last \nmers added. These mechanisms are referred to as terminal, penultimate, and \ncontrol, respectively.\n5.\nStereoregularity may be viewed as a subset of copolymerization, in of\nmonomer with an asymmetric center may follow the same stereochemistry previous\nrepeat unit, thereby forming a meso dyad, or by the opposite stereochemistry, \nracemic dyad. Isotactic, syndiotactic, and atactic polymers thus correspond to \nmeso dyads, predominantly racemic dyads, or random mixtures of the two, respectively.\n"]], ["block_5", ["Chapter Summary\n211\n"]], ["block_6", ["6.\nCopolymer sequence lengths (dyads, triads, tetrads, etc.) can be determined by \nThese in turn may be used to discriminate among terminal, penultimate, and antepenultimate\n"]], ["block_7", ["5.12\nChapter Summary\n"]]], "page_226": [["block_0", [{"image_0": "226_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "226_1.png", "coords": [34, 527, 226, 615], "fig_type": "figure"}]], ["block_2", ["control mechanisms. Similarly NMR gives access to stereochemical information, being\nsensitive to sequences of meso dyads, racemic dyads, and even longer sequences.\n7.\nStereoregularity is obtained by coordination polymerization in the presence of particular\ncatalysts. The most commonly used systems for the polymerization of a-olefins are referred\nto as Ziegler\u2014Natta catalysts, a class which actually spans a large variety of particular\ncompounds. The mechanisms of action of these catalysts are typically rather complicated.\nMore recently there have been rapid advances in the development of single\u2014site catalysts,\nwhich are usually based on metallocenes: a metal center coordinated to one or more\ncyclopentadienyl ligands. The terminology refers to the presence of a well-defined catalytic\nsite throughout the polymerization medium, leading to more homogeneous products. These\nsystems are capable of being fine-tuned to regulate a variety of structural features, including\nstereochemistry and comonomer addition.\n"]], ["block_3", ["212\nCopolymers, Microstructure, and Stereoregularity\n"]], ["block_4", ["1.\nWrite structural formulas for maleic anhydride (M1) and stilbene (M2). Neither of these\nmonomers homopolymerize to any significant extent, presumably owing to steric effects.\nThese monomers form a copolymer, however, with r1=r2=0.03.Jf Criticize or defend the\nfollowing proposition: The strong tendency toward alternation in this copolymer suggests that\npolarity effects offset the steric hindrance and permit copolymerization of these monomers.\n2.\nStyrene and methyl methacrylate have been used as comonomers in many investigations of\ncopolymerization. Use the following list of n values for each of these copolymerizing with the\nmonomers listed below to rank the latter with respect to reactivity. To the extent that the data\nallow, suggest where these substituents might be positioned in Table 5.3.\n"]], ["block_5", ["4.\nAdditional data from the research of the last problem yield the following pairs off], F 1 values\n(remember that styrene is component 1 in the styrene\u20141-chloro\u20141,3-butadiene system). Use the\n"]], ["block_6", ["Problems\n"]], ["block_7", ["3.\nAs part of the research described in Figure 5.5, Winston and Wichacheewa measured\nthe weight percentages of carbon and chlorine in copolymers of styrene (molecule 1) and\n1-chloro-1,3-butadiene (molecule 2) prepared from various feedstocks. A portion of their data\nis given below. Use these data to calculate F 1 , the mole fraction of styrene in these copolymers.\n"]], ["block_8", ["TRM. Lewis and ER. Mayo, J. Am. Chem. 506., 70, 1533 (1943).\n"]], ["block_9", ["f1\nPercent C\nPercent Cl\n"]], ["block_10", [{"image_2": "226_2.png", "coords": [51, 358, 326, 464], "fig_type": "figure"}]], ["block_11", ["M2\nStyrene as M1\nMethyl methacrylate as M1\n"]], ["block_12", ["Acrylonitrile\n0.41\n1.35\nAlly] acetate\n90\n23\n"]], ["block_13", ["1 ,2-Dichloropropene-2\n5\n5.5\nMethacrylonitrile\n0.30\n0.67\nVinyl chloride\n17\n12.5\nVinylidene chloride\n1.85\n2.53\n2-Vinyl pyridine\n0.55\n0.395\n"]], ["block_14", ["0.892\n81.80\n10.88\n0.649\n71.34\n20.14\n0.324\n64.95\n27.92\n0.153\n58.69\n34.79\n"]], ["block_15", [{"image_3": "226_3.png", "coords": [144, 380, 304, 457], "fig_type": "figure"}]]], "page_227": [["block_0", [{"image_0": "227_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "227_1.png", "coords": [34, 278, 279, 380], "fig_type": "figure"}]], ["block_2", ["6.\nUse the values determined in Example 5.5 for the vinylidene chloride (M1)\u2014isobutylene (M2)\nsystem to calculate F 1, for various values off1, according to the terminal mechanism. Prepare a\nplot of the results. On the same graph, plot the following experimentally measured values off1\nand F1. Comment on the quality of the fit.\n"]], ["block_3", ["5.\nThe reactivity ratios for the styrene (M1)\u20141-chloro\u20141,3-butadiene (M2) system were found to\nbe r, 0.26 and r2 1.02 by the authors of the research described in the last two problems,\nusing the results of all their measurements. Use these r values and the feed compositions listed\nbelow to calculate the fraction expected in the copolymer of 1-chlorobutadiene sequences of\nlengths v: 2, 3, or 4. From these calculated results, evaluate the ratios Nag/N22 and N2222/\nN222. Copolymers prepared from these feedstocks were dehydrohalogenated to yield the\npolyenes like that whose spectrum is shown in Figure 5.5. The absorbance at the indicated\nwavelengths was measured for 1% solutions of the products after HCl elimination.\n"]], ["block_4", ["7.\nSome additional dyad fractions from the research cited in the last problem are reported at\nintermediate feedstock concentrations (M1 :vinylidene chloride; Mzzisobutylenefr Still\nassuming terminal control, evaluate r, and r2 from these data. Criticize or defend the following\nproposition: The copolymer composition equation does not provide a very sensitive test for\n"]], ["block_5", ["TJ.B. Kinsinger, T. Fischer, and CW. Wilson, Polym. Left, 5, 285 (1967).\n"]], ["block_6", ["Problems\n213\n"]], ["block_7", [{"image_2": "227_2.png", "coords": [41, 483, 187, 570], "fig_type": "figure"}]], ["block_8", ["f1\nA=312nm\nA2367 nm\nA=412nm\n"]], ["block_9", ["f1\nF1\nf1\nF1\n"]], ["block_10", ["f1\nF1\nf1\nF1\n"]], ["block_11", ["0.947\n0.829\n0.448\n0.362\n0.861\n0.688\n0.247\n0.207\n0.698\n0.515\n0.221\n0.200\n0.602\n0.452\n"]], ["block_12", ["form suggested by Equation 5.6.1 to prepare a graph based on these data and evaluate r1\nand r2.\n"]], ["block_13", ["0.548\n0.83\n0.225\n0.66\n0.471\n0.79\n0.206\n0.64\n0.391\n0.74\n0.159\n0.61\n0.318\n0.71\n0.126\n0.58\n0.288\n0.70\n0.083\n0.52\n"]], ["block_14", ["As noted in Section 5.6, these different wavelengths correspond to absorbance by sequences of\ndifferent lengths. Compare the appropriate absorbance ratios with the theoretical sequence\nlength ratios calculated above and comment brie\ufb02y on the results.\n"]], ["block_15", ["0.551\n154\n77\n20\n0.490\n151\n78\n42\n"]], ["block_16", ["0.734\n71\n19\n\u2014\n"]], ["block_17", ["0.829\n74\n13\n\u2014\n"]], ["block_18", ["Absorbance\n"]]], "page_228": [["block_0", [{"image_0": "228_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018J.C. Randall, J. Polym. Sci, Polym. Phys. Ed, 13, 889 (1975).\n"]], ["block_2", ["214\nCopolymers, Microstructure. and Stereoregularity\n"]], ["block_3", ["10.\nRandall,r used 13C-NMR to study the methylene spectrum of polystyrene. In 1,2,4-trichlor-\nobenzene at 120\u00b0C, nine resonances were observed. These were assumed to arise from a\ncombination of tetrads and hexads. Using m and r notation, extend Table 5.6 to include all 20\npossible hexads. Criticize or defend the following proposition: Assuming that none of the\nresonances are obscured by overlap, there is only one way that nine methylene resonances\ncan be produced, namely, by one of the tetrads being split into hexads whereas the remaining\ntetrads remain unsplit.\n"]], ["block_4", ["11.\nIn the research described in the preceding problem, Randall was able to assign the five peaks\nassociated with tetrads in the 13C\u2014NMR spectrum on the basis of their relative intensities,\nassuming zero-order Markov statistics with pm 2: 0.575. The five tetrad intensities and their\nchemical shifts from TMS are as follows:\n"]], ["block_5", ["9.\nA hetero triad occurs at each interface between iso and syndio triads. The total number of\nhetero triads, therefore, equals the total number of sequences of all other types:\n"]], ["block_6", ["8.\nFox and Schnecko carried out the free-radical polymerization of methyl methacrylate\nbetween \u201440\u00b0C and 250\u00b0C. By analysis of the a\u2014methyl peaks in the NMR spectra of the\nproducts, they determined the following values of B, the probability of an isotactic placement\nin the products prepared at different temperatures.\n"]], ["block_7", [{"image_1": "228_1.png", "coords": [47, 88, 223, 205], "fig_type": "figure"}]], ["block_8", [{"image_2": "228_2.png", "coords": [52, 372, 164, 407], "fig_type": "molecule"}]], ["block_9", ["f,\n11\n12\n22\n"]], ["block_10", ["_\nIll-1\n_\n2\np\u201c\nvh+vi+vffzi+\ufb01s\n"]], ["block_11", ["the terminal control mechanism. Dyad fractions are more sensitive, but must be examined\nover a wide range of compositions to provide a valid test.\n"]], ["block_12", ["B\n0.36\n0.33\n0.27\n0.27\n0.24\n0.22\n0.20\n0.18\n0.14\n"]], ["block_13", ["Evaluate Ei\u201c 15;\" by means of an Arrhenius plot of these data using B/(l B) as a measure of\nki/ks. Brie\ufb02y justify this last relationship.\n"]], ["block_14", ["0.418\n0.55\n0.43\n0.03\n0.353\n0.48\n0.49\n0.04\n0.317\n0.44\n0.52\n0.04\n0.247\n0.38\n0.58\n0.04\n0.213\n0.34\n0.62\n0.04\n0. 198\n0.32\n0.64\n0.05\n"]], ["block_15", ["Criticize or defend the following proposition: The sequence DL\u2014 is already two thirds of the\nway to becoming a hetero triad, whereas the sequence DD\u2014 is two thirds of the way toward an\niso triad. This means that the fraction of heterotactic triads is larger when the average length\nof syndio sequences is greater than the average length of iso sequences.\n"]], ["block_16", ["T (\u00b0C)\n250\n150\n100\n95\n60\n30\n0\n\u201420\n\u201440\n"]], ["block_17", ["Use this relationship and Equation 5.9.1 to derive the expression\n"]], ["block_18", ["1\"h :Nni \u20181':a\n"]], ["block_19", ["Mole fraction of dyads\n"]]], "page_229": [["block_0", [{"image_0": "229_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n21 5\n"]], ["block_2", ["1K. Ito and Y. Yamashita, J. Polym. sci. 3A, 2165 (1965).\n"]], ["block_3", ["13.\nThe following are experimental tacticity fractions of polymers prepared from different\nmonomers and with various catalysts. On the basis of Figure 5.7, decide whether these\npreparations are adequately described by a single parameter pm or whether some other type\nof statistical description is required (remember to make some allowance for experimental\nerror). On the basis of these observations, criticize or defend the following proposition:\nRegardless of the monomer used, zero-order Markov statistics apply to all free-radical,\nanionic, and cationic polymerizations, but not to Ziegler\u2014Natta catalyzed systems.\n"]], ["block_4", ["12.\nThe fraction of sequences of the length indicated below have been measured for a copolymer\nsystem at different feed ratios.)r From appropriate ratios of these sequence lengths, what\nconclusions can be drawn concerning terminal versus penultimate control of addition?\n"]], ["block_5", ["14.\nReplacing one of the alkyl groups in R3Al with a halogen increases the stereospecificity of the\nZiegler\u2014Natta catalyst in the order I > Br > C1 > R. Replacement of a second alkyl by halogen\ndecreases specificity. Criticize or defend the following pr0position on of these\nobservations: The observed result of halogen substitution is consistent the effect on \n"]], ["block_6", [{"image_1": "229_1.png", "coords": [45, 245, 264, 328], "fig_type": "figure"}]], ["block_7", [{"image_2": "229_2.png", "coords": [48, 54, 225, 135], "fig_type": "figure"}]], ["block_8", ["45.38\n0.10\n44.94\n0.28\n44.25\n0.13\n43.77\n0.19\n42.84\n0.09\n"]], ["block_9", ["The remaining 21% of the peak area is distributed among the remaining hexad features. Use\nthe value of Pm given to calculate the probabilities of the unSplit tetrads (see Problem 10) and\non this basis assign the features listed above to the appropriate tetrads. Which of the tetrads\nappears to be split into hexads?\n"]], ["block_10", ["Methyl methacrylatea\nThermal\nToluene\n60\n8\n33\n59\nn-Butyl lithium\nToluene\n\u2014\u201478\n78\n16\n6\nn-Butyl lithium\nMethyl isobutyrate\n\u201478\n21\n31\n48\n(Jr-Methyl styreneb\nTiCl4\nToluene\n\u201478\n\u2014\n19\n8 1\nEt3Al/TiCl4\nBenzene\n25\n3\n35\n62\n"]], ["block_11", ["aMethyl methacrylate data from K. Hatada, K. Ota, and H. Yuki, Polym. Left, 5, 225 (1967).\nl\u2019ut-Methyl styrene data from S. Brownstein, S. Bywater, and OJ. Worsfold, Makromol. Chem,\n48, 127 (1961).\n"]], ["block_12", ["n\u2014Butyl lithium\nCyclohexane\n4\n\u2014\u2014\n31\n69\n"]], ["block_13", ["Catalyst\nSolvent\nT (\u00b0C)\n130\nHetero\nSyndio\n"]], ["block_14", ["[Mll/[l\nP(M1)\nP(M1M1)\nP(M1M1M1)\n"]], ["block_15", [{"image_3": "229_3.png", "coords": [57, 470, 378, 560], "fig_type": "figure"}]], ["block_16", ["l3C STMS (ppm)\nRelative area under peak\n"]], ["block_17", ["3\n0.168\n0.0643\n0.0149\n4\n0.189\n0.0563\n0.0161\n9\n0.388\n0.225\n0.107\n19\n0.592\n0.425\n0.278\n"]], ["block_18", [{"image_4": "229_4.png", "coords": [204, 447, 371, 565], "fig_type": "figure"}]], ["block_19", ["Fraction of polymer\n"]]], "page_230": [["block_0", [{"image_0": "230_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "230_1.png", "coords": [32, 142, 358, 265], "fig_type": "figure"}]], ["block_2", ["216\n"]], ["block_3", ["l.\nBrandrup, J. and Immergut, E.H., Eds., Polymer Handbook, 3rd ed., Wiley, New York, 1989.\n2.\nCoates, G.W. and Waymouth, R.M., Science, 267, 217 (1995).\n"]], ["block_4", ["Allcock, HR. and Lampe, F.W., Contemporary Polymer Chemistry, 2nd ed., Prentice-Hall, Englewood Cliffs,\nNJ, 1990.\nBovey, F.W., High Resolution NMR of Macromolecules, Academic Press, New York, 1972.\nCoates, G.W., Precise control of polyole\ufb01n stereochemistry using single\u2014site metal catalysts, Chem. Rev., 100,\n1223 (2000).\nKoenig, J.L., Chemical Microstructure of Polymer Chains, Wiley, New York, 1980.\nNorth, A.M., The Kinetics of Free Radical Polymerization, Pergamon Press, New York, 1966.\nOdian, G., Principles of Polymerization, 4th ed., Wiley, New York, 2004.\nRempp, P. and Merrill, E.W., Polymer Synthesis, 2nd ed., Hiithig & Wepf, Basel, 1991.\n"]], ["block_5", ["References\n"]], ["block_6", ["lFJ. Karol and W.L. Carrick, J. Am. Chem. Soc, 83, 585 (1960).\n"]], ["block_7", ["Further Readings\n"]], ["block_8", ["15.\n"]], ["block_9", ["16.\n"]], ["block_10", ["ease of alkylation produced by substituents of different electronegativity. This evidence thus\nadds credence to the monometallic mechanism, even though the observation involves the\norganometallic.\n"]], ["block_11", ["21](Cs )2\n4.5\nVOC13\n2.4\nZn(n-Bu)2\n4.5\n"]], ["block_12", ["The weight percent propylene in ethylene\u2014propylene copolymers for different Ziegler\u2014Natta\ncatalysts was measured for the initial polymer produced from identical feedstocksfr The\nfollowing results were obtained. Interpret these results in terms of the relative in\ufb02uence of the\ntwo components of the catalyst on the product found.\n"]], ["block_13", ["VCl4, plus\nAl(i-Bu)3, plus\nAl(i-Bu)3\n4.5\nHfCl4\n0.7\nCH3TiCl3\n4.5\nZrCl4\n0.8\n"]], ["block_14", ["Imagine a given single-site catalyst for polypropylene introduced a stereodefect on average\nonce every 10 monomer additions. Furthermore, assume the catalyst was supposed to be\nhighly isospecific. Explain how measurements of triad populations (e.g., mmm, mmr, etc.)\ncould be used to distinguish between chain-end control and site control. (Hint: consider the\nsequences of D and L in the two cases.)\n"]], ["block_15", ["Catalyst\nWeight percent\nCatalyst\nWeight percent\ncomponents\npropylene\ncomponents\npropylene\n"]], ["block_16", ["Copolymers, Microstructure, and Stereoregularity\n"]]], "page_231": [["block_0", [{"image_0": "231_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The remarkable properties of polymers derive from their size. As pointed out in Chapter 1, it is not\nthe high molecular weight per se that gives polymers mechanical strength, \ufb02exibility, elasticity,\netc., but rather their large spatial extent. In this chapter, we will learn how to describe the three-\ndimensional shape of polymers in an average sense, and how the average size of the object in space\nwill depend on molecular weight. We will also explore the equilibrium distribution of sizes.\nTo gain an appreciation of the possibilities, consider a polyethylene molecule with M 280,000\ng/mol (this is a reasonable value for a randomly selected molecule in some commercial grades of\npolyethylene). As the monomer (\u2014CHz\u2014CH2\u2014) molecular weight is 28 g/mol, the degree\nof polymerization, N, is 10,000, and there are 20,000 C\u2014C backbone bonds. Assuming a perfectly\nlinear structure (actually not likely for a commercial polyethylene), the contour length L of this\nmolecule would be roughly 20,000 X 1.5 A 30,000 A or about 3 pm, because 1.5 A is approximately\nthe average length of a C\u2014C bond. This is simply huge. If stretched out to its full extent, L would be\nhalf the size of a red blood cell, and possibly visible under a high-power optical microscope. Some\ncommercial polymers are 10 times bigger than this one, and some DNA molecules have molecular\nweights in excess of 109 g/mol. However, as we will see, it is very rare indeed for a chain to be so\nextended, and the contour length is not usually the most useful measure of size. Now consider the\nopposite extreme, where the same polyethylene molecule collapses into a dense ball or glabule. The\ndensity of bulk polyethylene is about 0.9 g/mL. The volume occupied by this 280,000 g/mol molecule\nwould be (280,000/0.9)/(6 x 1023) mL 520,000 A3, and if we assume it is a sphere, the radius would\nbe ((3/411)volume)1/as 50 A. The range from 50 A at the smallest to 3 pm at the largest covers three\norders of magnitude; it is a remarkable fact that such a mundane molecule could adopt conformations\nwith sizes varying over that range. If the dense sphere were a tennis ball, the chain contour would be the\nlength of a football field.\nPolyethylene, and most carbon chain polymers, is not likely to adopt either of these extreme\nconformations. The reason is easy to see. Select a C\u2014C bond anywhere along the chain; we can\nrepresent the structure as R\u2019CHr\u2014CHZR\u201d. There is rotation about this bond, with three energet-\nically preferred relative orientations of R\u2019 and R\u201d called trans (t), gauche plus (g+), and gauche\nminus (g') (see Figure 6.1a). For the chain of 20,000 bonds, there are three possible conformations\nfor each bond, and therefore 320\u2019000 m 1010\u2019000 possible conformations. This number is effectively\nin\ufb01nite. If our molecule were in a high temperature liquid state, and if we assume it takes 1 ps to\nchange one bond conformation, then the molecule would not even approach sampling all possible\nconformations over the history of the universe. Similarly, it would be highly improbable for it to\neven visit any given conformation twice. We can now see why the chance of being fully extended,\nin the all\u2014trans state, is unlikely to say the least; the probability is about 1 in 1010\u2019000. The dense\nsphere state might be marginally more probable, as there are many sequences of t, g+, and g\u2018 that\nmight produce something close to that, but it is still essentially impossible without the action of\nsome external force. What the polymer does instead is form what is called a random coil (Figure\n6.2). The different sequences of t, g+, and g\u2018 cause the chain to wander about haphazardly in\nspace, with a typical size intermediate between the dense sphere and the extended chain. Withso\n"]], ["block_2", ["6.1\nConformations, Bond Rotation, and Polymer Size\n"]], ["block_3", ["Polymer Conformations\n"]], ["block_4", ["6\n"]], ["block_5", ["217\n"]]], "page_232": [["block_0", [{"image_0": "232_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "232_1.png", "coords": [32, 278, 292, 592], "fig_type": "figure"}]], ["block_2", ["218\nPolymer Conformations\n"]], ["block_3", ["Figure 6.1\nBackbone bond conformations for polyethylene. (a) Illustration of trans and gauche arrange-\nments of the backbone bonds. (b) Schematic plot of the potential energy as a function of rotation angle about a\nsingle backbone bond.\n"]], ["block_4", ["Figure 6.2\nIllustration of a random walk in three dimensions. The walk has 4000 steps, and the walk\ntouches each face of the box. (From Lodge, A.S., An Introduction to Elastomer Molecular Network Theory,\nBannatek Press, Madison, 1999. With permission.)\n"]], ["block_5", [{"image_2": "232_2.png", "coords": [40, 122, 220, 217], "fig_type": "figure"}]], ["block_6", [{"image_3": "232_3.png", "coords": [52, 409, 302, 620], "fig_type": "figure"}]], ["block_7", [{"image_4": "232_4.png", "coords": [60, 132, 214, 220], "fig_type": "molecule"}]], ["block_8", [{"image_5": "232_5.png", "coords": [72, 53, 400, 214], "fig_type": "figure"}]], ["block_9", [{"image_6": "232_6.png", "coords": [85, 506, 266, 594], "fig_type": "molecule"}]], ["block_10", [{"image_7": "232_7.png", "coords": [91, 472, 263, 598], "fig_type": "figure"}]], ["block_11", [{"image_8": "232_8.png", "coords": [219, 28, 394, 229], "fig_type": "figure"}]], ["block_12", ["(arbitary\n"]], ["block_13", ["Potential\n"]], ["block_14", ["units)\n"]], ["block_15", ["energy\n"]], ["block_16", ["I\n\u20181\u2019\nI\nI\nI\n\u2014180 \u2014120\n\u201460\n0\n60\n120\n180\n"]], ["block_17", ["Rotation angle (deg)\n(b)\n"]], ["block_18", ["gauche\ngauche\n"]], ["block_19", ["Hans\n"]]], "page_233": [["block_0", [{"image_0": "233_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["many possibilities, cannot predict any instantaneous conformation or size, but we\nwill \nThe preceding argument, although fundamentally sound, neglects a very important aspect of\nchain conformations. ofthe t, g+, and g\u201d states are not equal; for polyethylene t is\nenergetically favorable relative to g+ or g\u2018 by about AB 3 kJ/mol (or 0.7 kcal/mol). Therefore, at\nequilibrium, the of t states will exceed that of g+ or g\u2018 by the appropriate Boltzmann\nfactor, exp(\u2014AE/RT), in thiscase is about 2 at T: 500 K (RT x42 kJ/mol or 1 kcal/mol).\n(We have injected some statistical mechanics here. For a large collection or ensemble of molecules at\nequilibrium, the of any two possible states are given by this Boltzmann factor.)\nThis bias is to put much of a dent into the vast number of possible conformations, but\nit does matter detailed conformational statistics for a given polymer chain.\nThere is a further issue of importance, namely, how high are the energy barriers among the t, g+,\nand g\" states? If these are too high, conformational rearrangements will not occur rapidly. Figure\n6.1b shows schematic plot of the potential energy as a function of rotation about the C\u2014C bond in\npolyethylene. The barrier heights are on the order of 10 kJ/mol (2.5 kcal/mol), which corresponds to\nabout 2.5 times RT, the available thermal energy per mole at 500 K. Thus rotation should be\nrelatively facile for polyethylene. (To put this energy barrier on a Chemist\u2019s scale, it is compar-\nable to the energy of a weak hydrogen bond.) However, for polymers with larger side-groups or\nwith more complicated backbone structures, these barriers can become substantial. For example,\nin poly(n-hexyl isocyanate) (\u2014N(C6H13)\u2014C(O)\u2014) the n-hexyl side chain forces the backbone to\nfavor a helical conformation, and the molecule becomes relatively extended.\nAt this point it might look like a very daunting task to calculate the probable conformation of a\ngiven polymer, and it will require some detailed information about bond rotational potentials, etc.,\nfor each structure. However, it turns out that we can go a long way without any such knowledge.\nWhat we will calculate \ufb01rst is the average distance between the ends of a chain, as a function of\nthe number of steps in a chain. We will show that this is given by a simple formula, and that all the\ndetails about chemical structures, bond potentials, etc., can be grouped into a single parameter.\nWe will also consider the distribution of possible values of this end\u2014to-end distance. The average\ncould, in principle, be taken in two different ways. One would be to follow a single chain as it\nsamples many different conformations\u2014a time average. Another would be to look at a large\ncollection of structurally identical chains at a given instant in time\u2014an ensemble average. In this\nexample, these two averages should be the same; when this occurs, we say the system is ergodic. In\n"]], ["block_2", ["a real polymer sample, a measurement will also average over a distlibution of chain lengths or\nmolecular weights; that is a different average, which we will have to reckon with when we consider\nparticular experimental techniques.\n"]], ["block_3", ["Average \n219\n"]], ["block_4", ["In this section, we calculate the root-mean-square (n'ns) end-to-end distance 022)\u201c2 for an\nimaginary chain, made up of n rigid links, each with length 6'. The model is sketched in\nFigure 6.3. At this stage there is no need to worry about whether the link is meant to represent a\nreal C\u2014C bond or not; we will make the correspondence to real polymers later. If we arbitrarily\nselect one end as the starting point, each link can be represented by a vector,\n13,-, with\ni 1,2,3, . . . n. The instantaneous end\u2014to\u2014end vector, 5, is simply the sum of the link \n"]], ["block_5", ["6.2\nAverage End\u2014to-End Distance for Model Chains\n"]], ["block_6", ["If we have a chain that wiggles around over time, or if we look at an ensemble of similar chains,\nthere13 no reason for h to point in any one direction more than any other, and the average ()1): 0;\nwe say the sample18 isotropic. What we really care about13 the average end-to\u2014end distance, which\n"]], ["block_7", [{"image_1": "233_1.png", "coords": [45, 588, 92, 628], "fig_type": "molecule"}]], ["block_8", ["\u201d2 :15}\n(6.2.1)\n"]], ["block_9", ["i=1\n"]]], "page_234": [["block_0", [{"image_0": "234_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "234_1.png", "coords": [28, 41, 164, 170], "fig_type": "figure"}]], ["block_2", ["In this simplest possible case, the orientation of link 1' is unaffected by link i\u2014l, and is equally\nlikely to point in any direction. It can even lie on top of link i\u2014l by pointing in the opposite\ndirection. (Remember we are dealing with imaginary links, not real chemical bonds, so this is\npermissible.) Mathematically we represent this approximation using the relation\n"]], ["block_3", ["The double sum can be broken into two parts, remembering also that the summations can be taken\noutside the average as follows:\n"]], ["block_4", ["where the first term, 1262, accounts for the \u201cself-terms,\u201d i.e., each of the n link vectors dotted into\nitself, and the second term accounts for the \u201ccross terms,\u201d i.e., each link vector dotted into the n\u2014\u20141\nother link vectors.\nWe now develop explicit results for (I22) ,with three different rules for how the orientation of\na given link, 19,-,IS constrained by the orientation of its predecessor, \u20ac,_ 1.\n"]], ["block_5", ["end-to\u2014end vector it extends from the start of the \ufb01rst link to the end of the last one.\n"]], ["block_6", ["we can calculate remembering that the length of a vector is obtained by taking the dot product of\nthe vector with itself:\n"]], ["block_7", ["Case 6.2.1\nThe Freely Jointed Chain\n"]], ["block_8", ["220\nPolymer Conformations\n"]], ["block_9", ["where 6 is the angle between E; and \u00a3311. For the freely jointed chain, 6 ranges freely from 0\u00b0 to\n180\u00b0. Thus on average\n"]], ["block_10", [{"image_2": "234_2.png", "coords": [35, 247, 223, 285], "fig_type": "molecule"}]], ["block_11", ["Figure 6.3\nModel chain consisting of .12 links of length 6. Each link is represented by a vector, E, and the\n"]], ["block_12", [{"image_3": "234_3.png", "coords": [36, 324, 220, 363], "fig_type": "molecule"}]], ["block_13", ["h2>=<2eza>=22\nI:\nj:\nr:\nj=\n"]], ["block_14", [{"image_4": "234_4.png", "coords": [45, 311, 230, 440], "fig_type": "figure"}]], ["block_15", ["6\", 2;, :22 cos 6\n(6.2.4)\n"]], ["block_16", ["1/2\n(112)1/2: (h- I?)_<:6o :E>\n(6.2.2)\n"]], ["block_17", ["<12} EH) 62(6639) 0\n(6.2.5)\n"]], ["block_18", [{"image_5": "234_5.png", "coords": [51, 360, 211, 401], "fig_type": "molecule"}]], ["block_19", [{"image_6": "234_6.png", "coords": [55, 317, 215, 441], "fig_type": "molecule"}]], ["block_20", ["=2??? E>+ZZ<\u20acW>\ni=1j9\u00e91\n: M + Z 2 <5- - (5)\n(6.2.3)\ni=1 #i\n"]]], "page_235": [["block_0", [{"image_0": "235_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "235_1.png", "coords": [19, 441, 269, 651], "fig_type": "figure"}]], ["block_2", [{"image_2": "235_2.png", "coords": [26, 540, 257, 640], "fig_type": "figure"}]], ["block_3", [{"image_3": "235_3.png", "coords": [28, 164, 196, 189], "fig_type": "molecule"}]], ["block_4", ["Now we make the model a bit more realistic. We will constrain the angle between adjacent links to\nbe a fixed value, 6, but still allow free rotation of the link around the cone defined by 6 (see Figure\n6.4). What happens? Now, 6,0 61-1 62\ncos6 does not average to zero because 6 is fixed. For\nsimplicity, we will define 01\u2014\n\u2014 cos 6 just to avoid writing cos 6 over and over. Returning to Equation\n62.3, what we need to calculateIS the double sum over all possible (6,- 6),i.e., the cross\nterrnzs\nas\nwell as the self-terms. We now know (6, 6-)\u2014\n- 6201 when |1'\u20141|\n\u2014 1, but what about |1' \u20141|\netc.? This1s a little sneaky: 6,- has a component parallel to 6,-_1, with length 601, but it also hzas3a\ncomponent perpendicular to 62-1 (with length 6 sin 6). However, because of the free rotation, over\ntime the perpendicular part will average to zero (see Figure 6.4). So from the point of view of bond\n6,-2, 011 average 6, looks just like a bond of length 601 pointing in the same direction as 6,-_1, and thus\n(6,-2 6-,-\u2014)62012. The same argument can be extended to any pair of bonds 1', j:\n"]], ["block_5", ["\u20141 to +1, with + or values equally probable. If we consider the relative orientations of any two\ndifferent links jointed chain, they must all be uncorrelated, so that\n"]], ["block_6", ["Average End-to-End Distance for Model Chains\n221\n"]], ["block_7", ["When the two linksthen (cos 6) 0, because cos 6 ranges from\n"]], ["block_8", ["because all the cross terms vanish. This is the classic result for the so-called random walk (or\nrandom\ufb02ight): the root-mean\u2014square excursion is given by the step length times the square root of\nthe number of steps. will invokethis result repeatedly in subsequent chapters.\nAt this point you may be thinking \u201cFine, but even if the link is not a C\u2014\u2014\u2014C bond, a real polymer\nchain cannot reverse its direction 180\u00b0 at a joint, so how can this result be relevant?\u201d Good\nquestion. Be patient for a bit.\n"]], ["block_9", ["Case 6.2.2\nThe Freely Rotating Chain\n"]], ["block_10", ["Now we define a new summation index k li\u2014j] and write\nii <1- 1-) 2\": DW\ni=1\n#1\ni=1\n#1\n"]], ["block_11", ["Figure 6.4\nDe\ufb01nition of the angles 6 and\n(15 for the freely rotating and hindered rotation chains; the\ncorrespondence to a polyethylene molecule is suggested by the locations of the carbons.\n"]], ["block_12", ["whenever\n1' ;\u00a3 j. In other words if the orientation of a given link is unaffected by its nearest\nneighbor, it must also be unaffected by more distant neighbors. From Equation 6.2.3 we now\nobtain the simple but important result\n"]], ["block_13", [{"image_4": "235_4.png", "coords": [46, 476, 232, 555], "fig_type": "molecule"}]], ["block_14", [{"image_5": "235_5.png", "coords": [47, 504, 232, 543], "fig_type": "molecule"}]], ["block_15", ["<17; 17,-) 0\n(6.2.6)\n"]], ["block_16", ["(16:11:32\nor (av/am\n(6.2.7)\n"]], ["block_17", ["<6,- 6,) tzali\u2014\ufb02\n(6.2.8)\n"]], ["block_18", ["11-1\n11\u20141\n11\u20141\n2 Z tzatm k) 21112 Z a\" 2112 Z ka\"\n"]], ["block_19", ["C\n/\n"]], ["block_20", ["\\9\n"]], ["block_21", ["60036\n"]], ["block_22", ["(6.2.9)\n"]]], "page_236": [["block_0", [{"image_0": "236_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "236_1.png", "coords": [33, 623, 217, 667], "fig_type": "molecule"}]], ["block_2", ["The factor of 2 comes about because there are two ways to get each value of k (one with i >j and\none with i < j). The term (n k) arises because there are n 1 nearest neighbors (I: 1), n 2 next\nnearest neighbors (k 2), etc. Note also that careful attention has to be paid to the limits\nof the sums.\nThe relevant summations have algebraic answers that are derived in the Appendix, namely:\n"]], ["block_3", ["222\nPolymer Conformations\n"]], ["block_4", ["4!) are possible to some extent (see Figure 6.1b). The derivation of (122) is more complicated for this\ncase, as you might expect, but it is similar in spirit to that for the freely rotating chain; it may be\nfound in Flory\u2019s second book [1]. The large :1 result is\n"]], ["block_5", ["In a real polyethylene chain, the rotation about the cone is not free; there are three preferred\nconformations (t, g+, g\u2018) as discussed in Section 6.1. Furthermore, all values of the rotation angle\n"]], ["block_6", ["forn \u2014> oo, assuming |a| < 1 (i.e., 6 79 90\u00b0), and\n"]], ["block_7", ["where again we let n\u2014> 00 to reach the last expression. Now we insert Equation 6.2.10a and\nEquation 6.2.10b into Equation 6.2.9, and we recall Equation 6.2.3 to obtain\n"]], ["block_8", ["where again we assume n is large in the penultimate step. (We also reinserted cos 6 for a.) We can\nlearn three important things from this result.\n"]], ["block_9", ["1.\n(112) is larger than the freely jointed chain result if 6 < 90\u00b0 (cos 6 > 0). This is very reasonable;\nif each link has some preference for heading in the same direction as the previous link, the\nchain will double back on itself less often. As an example, for C\u2014C single bonds, 6 is close to\n705\u00b0 (the complement to the tetrahedral angle) and for this value (112) x 21262.\n2.\n(/22) is still proportional to n62; the proportionality factor is just a number that depends on the\ndetails of the local constraints placed on link orientation. In particular, therefore, (122) is still\nproportional to (/13.\n3.\nThe previous statement applies strictly only in the large n limit, i.e., when the term propor-\ntional to 1/11 in Equation 6.2.11 is negligible and when of\u201d vanishes. This is a commonly\nencountered caveat in polymer science: we can derive relatively simple expressions, but they\nwill often be valid only in the large n limit. The answer to \u201cHow large is large enough?\u201d will\ndepend on the particular property, but when the correction is proportional to 1/12, as it is in\nEquation 6.2.11, it will drop to the order of 1% when n m 100, which is not a particularly large\nnumber of backbone bonds.\n"]], ["block_10", ["Case 6.2.3\nHindered Rotation Chain\n"]], ["block_11", [{"image_2": "236_2.png", "coords": [37, 122, 180, 158], "fig_type": "molecule"}]], ["block_12", [{"image_3": "236_3.png", "coords": [40, 181, 193, 213], "fig_type": "molecule"}]], ["block_13", [{"image_4": "236_4.png", "coords": [41, 243, 263, 319], "fig_type": "molecule"}]], ["block_14", ["Zkak \n1\u2014 (In\nN\nC!\n"]], ["block_15", ["\"\u20141\nk\n1 Gin\u2014I\na\na a\n%\n(6.2.10a)\nk=l\n1\u20140:\n1\u20140:\n"]], ["block_16", ["2\n_\n2\nl+cos6\n1+(cosqb)\n(h) \u2014n\u20ac{1\u2014cos6}{l\u2014(cosqb)}\n(6.2.12)\n"]], ["block_17", ["2 _\n2\n2\na\n_\n2\n0f\n(h)\u2014\u2014n\u20ac +2121? (1_)\n26(1_a)2\n\u201c,2 m3;\nW92 Li\u201c.\n1\u20140:\nIt\n(1\u2014092\n1\u20140:\nl+cos6\n__\n2\n_ \u201dE {1\u20140036}\n(6.2.11)\n"]], ["block_18", ["=\n(1 a)2\n(1 a)2\n(6.2.10b)\n"]]], "page_237": [["block_0", [{"image_0": "237_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a theorem:\nIf we take the limit n~>oo, and if we consider \u201cphantom\u201d chains that can double back on\nthemselves, C is a numerical constant that depends local\nconstraints and not on n.\nThe physical content of this theorem can be summarized as follows. If we have a chain of links\nwith any degree of conformational freedom, no matter how limited that freedom may be, and if we\ntrack the conformation over enough links, the orientation of the last link will have lost all memory\nof the orientation of the first. At this point we could replace that entire subset of links with one new\nlink, and it would be freely jointed with respect to the next set and the previous set. In other words,\nfor any chain of n links whose relative orientations are constrained, we can always generate an\nequivalent chain with a new (bigger) link that is freely jointed, so that the original chain and the\nnew chain have the same (h2). We will illustrate this concept in the next section.\nThe issue of how large it needs to be for this theorem to be useful was mentioned before; it of a\nhundred or so is usually more than adequate. The fact that real polymers occupy real volume means\nthat polymer chains cannot double back on themselves, or have two or more links at the same point\nin space. This so-called \u201cexcluded volume\u201d problem is actually very serious, and makes an exact\nsolution for (hz) of a real polymer much more complicated. However, it turns out that there are two\npractical situations in which we can make this problem essentially go away; one is in a molten\npolymer, and the other is in a particular kind of solvent, called a theta solvent. Under these\ncircumstances, a polymer is said to exhibit \u201cunperturbed\u201d dimensions. Thus this theorem is of\ntremendous practical importance. Further discussion of the excluded volume effect will be deferred\nto Section 6.8, and then it will be revisited in more detail in Chapter 7.7.\n"]], ["block_2", ["where (cos (I5) is the average of cos over the appropriate potential energy curve (Figure 6.1b).\nHowever, the important message is that (hz) is still proportional to n32; all that has changed is a\nnumerical prefactor that specific local constraints. This point, in fact, as\n"]], ["block_3", ["We can define a quantity Cn, called the characteristic ratio, which for any polymer structure\ndescribes the effect of local constraints on the chain dimensions:\n"]], ["block_4", ["Characteristic Ratio and Statistical Segment Length\n223\n"]], ["block_5", ["In this equation, (h2)0 is the actual mean-square end-to-end distance of the polymer chain,\nand the subscript O reminds us that we are referring to unperturbed dimensions. In Equation\n6.3.1, n denotes the number of chemical bonds along the polymer backbone, and is actual\nlength of a backbone bond, e.g., 1.5 A for C\u2014C. (For polymers containing different kinds\nof backbone bonds, such as polyisoprene or poly(ethylene oxide), it is appropriate simply to\nadd n16? + ngl\u2019g \u2014l\u2014\n\u00ab - 4 where n,- and E, are the number and length of bonds of type i, respectively.)\n"]], ["block_6", ["6.3\nCharacteristic Ratio and Statistical Segment Length\n"]], ["block_7", ["caused the chain to extend in one direction. As defined in Equation 6.3.1, C\u201d but\nit approaches a constant value at large n; this is often denoted C00. For the freely rotating C00 is\n"]], ["block_8", ["(1 + cos 6)/(1 cos 6) from Equation 6.2.11. The dependence of C,, on n is shown in 6.5 for\nseveral theoretical chains. The values of C00 for several common polymers are listed 6.1.\nFor polymers that have primarily C\u2014C or C\u2014O single bonds along the backbone, C00 from\nabout 4 to about 12. Using these values, or those provided in reference books, it is to\nestimate (h2)0 for any polymer of known structure and molecular weight.\nAlthough calculating (h2)0 for a given polymer is thus a solved problem, this C00\nis not always the most convenient. For example, it requires remembering particular lengths\n"]], ["block_9", ["C,, is a measure of chain flexibility: the larger the value of C,,, the more the local have\n"]], ["block_10", [{"image_1": "237_1.png", "coords": [37, 425, 101, 460], "fig_type": "molecule"}]], ["block_11", ["(hzlo\nCn\nr1132\n(6.3.1)\n"]]], "page_238": [["block_0", [{"image_0": "238_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "238_1.png", "coords": [23, 499, 372, 613], "fig_type": "table"}]], ["block_2", [{"image_2": "238_2.png", "coords": [29, 477, 434, 643], "fig_type": "figure"}]], ["block_3", [{"image_3": "238_3.png", "coords": [30, 31, 232, 242], "fig_type": "figure"}]], ["block_4", ["Table 6.1\nRepresentative Values of the Characteristic Ratio, Statistical Segment Length, and Persistence\nLength for Various Flexible Polymers, Calculated from the Experimental Quantities, (h2)0/M (A2 mol/g),\nvia Equation 6.3.2 and Equation 6.4.5b\n"]], ["block_5", ["The experimental chain dimensions were obtained by small-angle neutron scattering, as compiled in Fetters, L.J., Lohse,\nD.J., Witten, T.A., and Zirkel, A., Macromolecules, 27, 4639, 1994. The uncertainties in (h2)0 and M are typically a few\npercent. The temperature of the measurement is indicated, because the distribution of chain conformations depends on\ntemperature.\n"]], ["block_6", ["Figure 6.5\nCharacteristic ratio as a function of the number of bonds for three model chains. The dotted\ncurve represents the freely rotating chain with 6:68\u00b0. The long dashed curve corresponds to a parti-\ncular hindered rotation chain with the preferred values of (b 120\u00b0 apart, but in which values of qb for\nneighboring bonds are independent. The smooth curve applies to an interdependence among values\nof (b on neighboring bonds. (Reproduced from Flory, P.J., Statistical Mechanics of Chain Molecules,\nWiley-Interscience, New York, 1969. With permission.)\n"]], ["block_7", ["Poly(ethylene oxide)\n5.6\n6.0\n4.1\n0.805\n140\n1,4-Polybutadiene\n5 .3\n6.9\n4.0\n0.876\n140\n1,4\u2014Polyisoprene\n4.8\n6.5\n3.5\n0.625\n140\nPoly(dimethylsiloxane)\n6.6\n5.8\n5.3\n0.457\n140\nPolyethylene\n7.4\n5.9\n5.7\n1.25\n140\nPolyprOpylene\n5.9\n5.3\n4.6\n0.67\n140\nPolyisobutylene\n6.7\n5.6\n5.2\n0.57\n140\nPoly(methyl methacrylate)\n9.0\n6.5\n6.9\n0.425\n140\nPoly(vinyl acetate)\n8.9\n6.5\n6.8\n0.49\n25\nPolystyrene\n9.5\n6.7\n7.3\n0.434\n140\n"]], ["block_8", ["Polymer\nCo,\nb (ti)\n3,, (ti)\n(h2)0/M (.38 mol/g)\nT (\u00b0C)\n"]], ["block_9", ["224\nPolymer Conformations\n"]], ["block_10", ["can replace the number of bonds, n, with the number of monomers or repeat units, N, and subsume\nthe proportionality factor between rt and N into a new effective step length, b. The quantity b,\ndefined by Equation 6.3.2, is called the statistical segment length. The calculation of (h2)0 through\n"]], ["block_11", ["where Eeff fix/E is a new effective bond length with the following meaning: the real chain with\nlocal constraints has an end\u2014to\u2014end distance, which is the same as that of a freely jointed chain with\nthe same number of links rt, but with a different (larger) step length Eeff. Continuing in this vein, we\n"]], ["block_12", ["and the number of bonds per repeat unit. A more popular approach was suggested in the previous\nsection. We could rewrite Equation 6.3.1 in the following way:\n"]], ["block_13", ["6.87\nl\nI\nI\nI\nl\n!\n__l_\n"]], ["block_14", ["6.0\nInterdependent rotations\n"]], ["block_15", [", \n2 0\n\"\nI\n.\u2019\n\u2019\n"]], ["block_16", ["(11% Cam/32 2262.. s\n(6.3.2)\n"]], ["block_17", ["0\nI\n|\nl\n|\nI\n|\n0\n50\n100\n1 50\nNumber of bonds, it\n"]], ["block_18", ["Independent hindered rotations\n"]], ["block_19", [{"image_4": "238_4.png", "coords": [172, 503, 445, 626], "fig_type": "figure"}]]], "page_239": [["block_0", [{"image_0": "239_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["N and b treats the real chain though it were a freely jointed chain with N links of length I). This\nproves to be a useful computational scheme, but it is important to realize that b has no simple\ncorrespondence with the physical chain; it is not a measure of the real size of a real monomer.\n"]], ["block_2", ["Furthermore, it contains new information beyond that embodied in C00. Values of b are also\ngiven in Table 6.1. Note that although b varies monotonically with C00, it is not as simple a measure\nof \ufb02exibility. For example, b for polystyrene (6.7 A) and polyisoprene (6.5 A) differ by only a few\npercent, whereas polyiSOprene is considered to be a relatively \ufb02exible polymer, and polystyrene a\nrelatively stiff one. The resolution of this apparent paradox is left to Problem 6.2. Values of C00 and\n"]], ["block_3", ["It is an interesting fact that bulk polyethylene has a positive coefficient of thermal expansion (about\n2.5 x 10\u20144 per \u00b0C at room temperature), whereas the individual chain dimensions have a negative\ncoefficient (d In (h2)0/dT: $1.2 >< 10\u20143deg\u201c1). In other words, when a piece of polyethylene is\nheated, the volume increases while the individual chain dimensions shrink. How does this\ncome about?\n"]], ["block_4", ["measurements of the chain dimensions; this method will be described in detail in Chapter 8.\nThe approach taken above, and indeed for the remainder of the book, is that C0\u2018, or b are\nstructure\u2014specific parameters that we can look up as needed; the dependence of (h2)0 on molecular\nweight is universal and therefore more important to understand. However, it is of considerable\ninterest to ask whether the techniques of computational statistical mechanics can be used to\ncalculate CDO from first principles, i.e., from knowledge of bond angles, rotational potentials,\netc. A highly successful scheme for doing so, called the rotational isomeric state approach, was\ndeveloped by Flory [1]. It is beyond the scope of this book to describe it, but it is worth mentioning\nthat even today it is not a trivial matter to execute such calculations, and that controversy exists\nabout the correct values of C00 for some relatively simple chemical structures. These controversies\nare also not easily resolved by experiment; combined uncertainties in measured molecular weights\nand chain dimensions often exceed 10%.\n"]], ["block_5", ["Thermal expansion corresponds to a decrease in the density of the material, which re\ufb02ects\nprimarily an increase in the average distance between molecules; the radii of the individual\n"]], ["block_6", ["atoms and the bond lengths also tend to increase, but to a much smaller extent. In contrast, the\nreduction in (h2)0 is primarily of intramolecular origin. From Equation 6.3.1, we can see that as n is\nindependent of T and E, if anything, increases with T, then there must be a decrease in the\ncharacteristic ratio, C00. The origin of this effect can be seen from Figure 6.1. As temperature\nincreases, the Boltzmann factors that dictate the relative equilibrium populations of trans and\ngauche conformations change, and the gauche states become relatively more populated. As the\ntrans conformations favor larger (h2)0, the net result is a reduction in C00. Note that this simple\nrelation between C00 and the relative populations of trans and gauche states does not necessarily\nextend to more complicated backbone structures. For example, d In (h2)0/dT is positive 100%\ncis-l,4\u2014polybutadiene, but negative for the all-trans versions (see Problem 65). This observation is\nnot easy to anticipate based on the molecular structure.\n"]], ["block_7", ["b may be determined experimentally in various ways, but the most direct route is by scattering\n"]], ["block_8", ["For many macromolecules, the backbone does not consist of a string of single bonds with facile\nrotations, but rather some combination of bonds that tend to make the backbone in one\ndirection. Such chains are called semiflexible, and examples (see Figure 6.6) include polymers\nwith mostly aromatic rings along the backbone, such as poly(p\u2014phenylene); polymers with large\n"]], ["block_9", ["Semiflexible Chains and the Persistence Length\n225\n"]], ["block_10", ["6.4\nSemiflexible Chains and the Persistence Length\n"]], ["block_11", ["Example 6.1\n"]], ["block_12", ["Solution\n"]]], "page_240": [["block_0", [{"image_0": "240_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "240_1.png", "coords": [21, 59, 121, 107], "fig_type": "molecule"}]], ["block_2", [{"image_2": "240_2.png", "coords": [29, 497, 281, 581], "fig_type": "figure"}]], ["block_3", ["226\nPolymer Conformations\nsi\n"]], ["block_4", ["side\u2014groups that for steric reasons induce the backbone to adopt a helical conformation, such as\npoly(n-hexyl isocyanate)\nand poly('y\u2014benzyl-L\u2014glutamate); biopolymers\nsuch\nas DNA and\ncollagen that involve intertwined double or triple helices. Short versions of these molecules\nare essentially \u201crigid rods,\u201d but very long versions will wander about enough to be random coils.\nThe description of chain dimensions in terms of either COO or b turns out not to be as useful for\nthis class of macromolecules. Therefore it is desirable to have a method to calculate the\ndimensions of such molecules, and particularly to understand the crossover from rod-like to\ncoil\u2014like behavior. Such a scheme is provided by the so\u2014called worm-like chain of Kratky and\nPorod [2]; the fundamental concept is that of the persistence length, 6p, which is a measure of\nhow far along the backbone one has to go before the orientation changes appreciably. A garden\nhose provides a good everyday analogy to a semi\ufb02exible polymer, with a persistence length on\nthe order of 1 ft. A 2 in. section of hose is relatively stiff or rigid, whereas the full 50 ft hose can\nbe wrapped around and tangled with itself many times like a random coil. We will \ufb01rst define 3p\nfor \ufb02exible chains and see how it is simply related to C00. Then we will develop in terms of 6,, an\nexpression for (/12) that can be used to describe \ufb02exible, semi\ufb02exible, and rigid chains.\nThe persistence length represents the tendency of the chain to continue to point in a\nparticular direction as one moves along the backbone. It can be calculated by taking the projection\nof the end-to-end vector on the direction of the \ufb01rst bond (\u201957: /\u20ac is a unit vector in the direction\n"]], ["block_5", ["For the freely jointed chain, as discussed above, all the terms in the expansion in Equation 6.4.1 are\nzero except the \ufb01rst, and thus El, E. For chains with more and more conformational constraints\nthat encourage the backbone to straighten out, more and more terms in the expansion wili\ncontribute positively, and 6p increases. In the limit that every bond points in the same direction\nthe persistence length tends to infinity. When Ep > L, where L n3 is the contour length of the\nchain, such a molecule is called a rigid rod.\n"]], ["block_6", ["Of 61):\n"]], ["block_7", ["Figure 6.6\nExamples of polymer structures that are semi\ufb02exible or stiff chains.\n"]], ["block_8", ["Poly(n\u2014hexyl isocyanate)\nPoly(y-benzyl-L-glutamate)\n"]], ["block_9", ["Poly(p\u2014phenylene)\nPoly(p\u2014phenylene terephthalamide)\n"]], ["block_10", [{"image_3": "240_3.png", "coords": [51, 125, 110, 195], "fig_type": "molecule"}]], ["block_11", [{"image_4": "240_4.png", "coords": [54, 503, 250, 579], "fig_type": "molecule"}]], ["block_12", ["=\u00a7{<:-1>++...+}\n(64-12\n"]], ["block_13", ["Ci )\nNJ\nn\n"]], ["block_14", ["N\nn\n"]], ["block_15", ["2.\n.\n"]], ["block_16", [{"image_5": "240_5.png", "coords": [149, 52, 298, 97], "fig_type": "molecule"}]], ["block_17", [{"image_6": "240_6.png", "coords": [163, 41, 297, 135], "fig_type": "molecule"}]], ["block_18", [{"image_7": "240_7.png", "coords": [190, 121, 267, 201], "fig_type": "molecule"}]], ["block_19", [{"image_8": "240_8.png", "coords": [201, 122, 267, 163], "fig_type": "molecule"}]], ["block_20", ["O\nH\n"]]], "page_241": [["block_0", [{"image_0": "241_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "241_1.png", "coords": [33, 322, 182, 360], "fig_type": "molecule"}]], ["block_2", ["The difference between Equation 6.4.5a and Equation 6.4.5b is not particularly important, espe-\ncially for stiffer chains where CDO>>1; we will use the latter form below, because it is simpler.\nA related quantity in common use is the Kuhn length [3], 6k, which is de\ufb01ned as twice the\npersistence length:\n"]], ["block_3", ["We thus have three different, but fully equivalent expressions for the mean-square unperturbed\n"]], ["block_4", ["Where did the extra E on the right hand side of Equation 6.4.3 come from? Well, the double\ncounting was not quite complete; there were two terms with |x j | 1, two with |x \u2014j| :2, etc.,\nbut only one with x:j. We need this contribution of E in order to obtain 26p, and therefore the\nmissing \u201cself\u201d term is appended to Equation 6.4.3.\nIn order to remove the arbitrary choice of bond )6, we sum over all possible choices of x, and\nassume that we get the same answer for each it; this approximation neglects the effects of chain\nends, so it is valid only in the limit of large n.\n"]], ["block_5", ["end-to\u2014end distance of a \ufb02exible chain:\n"]], ["block_6", ["In some derivations of this relation, the joint limit it \u2014> 00 and E \u2014> 0 is taken (see following\nsection). In this case, the extra 6 on the right hand side of Equation 6.4.3 would vanish, and\nEquation 6.4.5a would become\n"]], ["block_7", ["All three are useful and frequently employed, so they are worth remembering. Estimates of the\npersistence lengths of \ufb02exible polymers are also listed in Table 6.1.\n"]], ["block_8", ["Here we divided by n in front of the double sum because we added 72 identical terms through the\nsum over x. All of these manipulations finally pay off: we recognize this double sum as exactly\n([12) from Equation 6.2.2, and hence\n"]], ["block_9", ["We now seek a relation between 6,, and COO for long, \ufb02exible chains. We can rewrite Equation\n6.4.1 as\n"]], ["block_10", ["where x is any arbitrary bond in the chain. We can make this substitution because for a \ufb02exible\nchain, only a few terms j with small\n|j x| will contribute. Now we change the limits of the sum\nover j to extend over the complete chain; in other words we look in both directions from bond 16.\nThis amounts to a double counting, so we multiply 6,, by 2:\n"]], ["block_11", ["Semiflexible Chains and the Persistence Length\n227\n"]], ["block_12", ["6.4.1\nPersistence Length of Flexible Chains\n"]], ["block_13", [{"image_2": "241_2.png", "coords": [37, 194, 157, 233], "fig_type": "molecule"}]], ["block_14", [{"image_3": "241_3.png", "coords": [38, 99, 210, 139], "fig_type": "molecule"}]], ["block_15", [{"image_4": "241_4.png", "coords": [45, 482, 211, 512], "fig_type": "molecule"}]], ["block_16", ["(2,. 2 2:2,, :C006\n(6.4.6)\n"]], ["block_17", ["2.6, :%\n12-1:\n<4; 4}) + 6\n(6.4.3)\n"]], ["block_18", ["E\n(7,0662 + E :((3,, + 1) \n(6.4.5a)\n13p\n2\n"]], ["block_19", [">>1>>\n>>1>>\nE\nE :_I22\n= \u2014\u2014C\n42 (300\u2014\n6.4.5b\n>P>\n2n}?<\n>\n2m?\n00\u201d\n2\n(\n)\n"]], ["block_20", ["1\n\u201d\n\"\n4\n4\n(2\nepzm;;<\u00a3,-\u20acj>+\u00a7\n(6.4.4)\n"]], ["block_21", ["gp:%i=%:\n(6.4.2)\n1:1\nJ=x\n"]], ["block_22", ["09),, :(3,0622 :N62 Lfk\n(6.4.7)\n"]], ["block_23", ["I\nE\n1\nE\n__\n2\n_:_____\n_2n!3\u2019<\n>+2\n2m.\u201d\n"]]], "page_242": [["block_0", [{"image_0": "242_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "242_1.png", "coords": [28, 262, 263, 333], "fig_type": "molecule"}]], ["block_2", ["So far, all that the persistence length has given us is a new way to express (h2)0 for \ufb02exible chains.\nTo obtain a useful result for semi\ufb02exible and stiff chains, we will return to the freely rotating chain\nof Section 6.2 and transform it into a continuous worm-like chain. This we do by taking a special\nlimit alluded to earlier; we will let the number of bonds, n, go to infinity, but the length of each\nbond, 6, will go to zero, while maintaining the contour length L n6 constant. We begin by\nrelating E], to a cos 6 of the freely rotating chain, starting from Equation 6.4.1:\nmm.a>++~-+}\n1\n= EM +\u20ac2a+\u20ac2a2 +~-+\u20ac2a\u201d_1}\n"]], ["block_3", ["The last transformation utilized the summation results in Equation 6.2.10a, and the large n limit.\nThus we can write\n"]], ["block_4", ["This expression is the result for the worm-like chain obtained by Kratky and Porod [2]. (Note that\nwe took (12 1).? 9: n6 L in Equation 6.4.10.) It is left as an exercise (Problem 6.6) to show\nthat in the coil limit (L >> 61,) this expression reverts to Equation 6.4.7 and that in the rod limit\n(L < \u20ac13), ([12) =L2, as it should. Examples of experimental persistence lengths for semi\ufb02exible\npolymers are given in Table 6.2.\n"]], ["block_5", ["Returning to Equation 6.4.1, we can actually extract a very appealing physical meaning for El,\nand Bk. The terms in the expansion become progressively smaller as the average orientation of\nbond\n1' becomes less correlated with that of the first bond. In fact, when bond 1' is on average\nperpendicular to the first bond (i.e., is uncorrelated), (\u00a3143) a: 0. All higher-order terms will also\nvanish. Thus the persistence length measures how far we have to travel along the chain before it\nwill, on average, bend 90\u00b0. Similarly, the Kuhn length tells us how far we have to go along the\nchain contour before it will, on average, reverse direction completely. Equation 6.4.6 also provides\na simple interpretation for COO: it is the number of backbone bonds needed for the chain to easily\nbend 180\u00b0.\n"]], ["block_6", ["Now we recall Equation 6.2.11 for the freely rotating chain, but retaining the term in a\u201d:\n"]], ["block_7", ["and thus\n"]], ["block_8", ["6.4.2\nWorm-Like Chains\n"]], ["block_9", ["where we invoke the series expansion (see the Appendix):\n"]], ["block_10", ["228\nPolymer Conformations\n"]], ["block_11", [{"image_2": "242_2.png", "coords": [40, 382, 222, 411], "fig_type": "molecule"}]], ["block_12", [{"image_3": "242_3.png", "coords": [42, 482, 188, 513], "fig_type": "molecule"}]], ["block_13", ["ex=1+x+\u00a32+m\n2!\n"]], ["block_14", ["6\na21\u2014exp(\u2014\u20ac/\u20acp)\nfor\n\u20ac\u2014>0\n(6.4.9)\np\n"]], ["block_15", ["2\n__\n21+a_\n2(1\u2014a\u201d)\n(h ) \u2014n\u00a3\u2019\n\u2014\u2014\u2014\u2014\u20141\n\u2014oz\n213\n(10 \u2014o:)2\n"]], ["block_16", ["(1?) 26,11, 2e\u00a7(1 exp[\u2014L/\u00a3p])\nas\ne \u2014+ 0\n(6.4.11)\n"]], ["block_17", [{"image_4": "242_4.png", "coords": [50, 496, 300, 569], "fig_type": "molecule"}]], ["block_18", ["1_\nn\n=\u20ac{1+a+a2+~-+a\u201d*l}=\u20ac(\na):\n6\n(6.4.8)\n1\u20140:\n1\u20140:\n"]], ["block_19", ["2\u2014_\u20ac/\u20acp)\n_\n2\n_\n1\u2014 exp[\u2014L/\u20acp]\n\u2014n\u20ac (\n3/3p\n26 (1\nE/Ep)\n(3/592\n= L\u20acp(2 E/Ep) 263(1 E/\u00e9\u2019p)(1 \u2014 exp[\u2014L/\u20acp])\n(6.4.10)\n"]], ["block_20", ["_\n2\n"]]], "page_243": [["block_0", [{"image_0": "243_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "243_1.png", "coords": [18, 82, 381, 224], "fig_type": "figure"}]], ["block_2", ["Table \nRepresentative Values of Persistence Lengths for Semiflexible Polymers.\nNote That Values May Depend Either Weakly or Strongly on the Choice of Solventa\n"]], ["block_3", ["_,__\u2014\nPolymer\nSolvent\n1?], (231)\n"]], ["block_4", ["Hydroxypropyl cellulose\nDimethylacetamide\n65\nPoly(p-phenylene)b\nToluene\n130\nPoly(p-phenylene \nMethane sulfonic acid0\n100\n96% Sulfuric acid\n180\nPoly(n-hexyl \nHexane\u2018:l\n420\nDichloromethane\n210\nDNA (double helix)\n0.2 M NaCl\n600\nXanthan (double helix)\n0.1 M NaCl\n1200\nPoly(y-benzyl-L-glutamate)\nDimethylformamide\n1500\nSchizophyllan\nWater\n2000\n"]], ["block_5", ["Semiflexible Chains and the \n229\n"]], ["block_6", ["8\u2018Data are summarized in Sato, T. and Polym. Sci, 126, 85, 1996, except for I\u201d\\l\u2019anhee,\nS., Rulkens, R., Lehmann, U., Rosenauer, M.,Wegner, G., Macromolecules,\n29, 5136, 1996; cChu, S.G., Venkatraman, 8., Y., Macromolecules, 14, 939, 1981;\nand dMuralucami, H., Norisuye, T., and Fujita, 13, 345, 1980.\n"]], ["block_7", ["Figure 6.7 shows a plot of a dimensionless form of Equation 6.4.11, obtained by dividing\nthrough by 6% and plotting against L/\u00e9\u2019p. This independent variable is the number of\nlengths in the chain, i.e., an effective degree of polymerization. The curve illustrates \ncrossover from the rod-like behavior at small L/6p, with (h2)0~M2, to the coil\u2014like \nlarge L/Q, with (h2)0 M. Thus the worm-like chain model is able to describe \nsemiflexible chains with one expression. The double logarithmic format of \nemployed in polymer science, when both the independent variable (such as M) and the \n"]], ["block_8", ["(h2)/\u20ac\u00a7\n"]], ["block_9", ["function of the number of persistence lengths \naccording to the Kratky\u2014Porod worm-like chain. The asymptotic \n"]], ["block_10", ["also shown, as is the location of a chain with length equal to \n"]], ["block_11", ["Figure 6.7\nThe mean-square end-to-end distance, normalized by the \n"]], ["block_12", [{"image_2": "243_2.png", "coords": [40, 356, 114, 615], "fig_type": "figure"}]], ["block_13", [{"image_3": "243_3.png", "coords": [49, 362, 300, 619], "fig_type": "figure"}]], ["block_14", ["1o\u20141\n"]], ["block_15", ["10-3;'\nWorm-like chain\n"]], ["block_16", ["10\u20145-\n||||||III\nn\nnunnnnl\nI\nIlllllll\nI\nIIIIIII'\nr\nllll\n-\n"]], ["block_17", ["103;\n"]], ["block_18", ["101\n"]], ["block_19", ["10-2\n10-1\n1o0\n101\n1o2\n103\n"]], ["block_20", [{"image_4": "243_4.png", "coords": [89, 94, 291, 212], "fig_type": "figure"}]], ["block_21", ["L/Ep\n"]], ["block_22", [{"image_5": "243_5.png", "coords": [192, 88, 363, 215], "fig_type": "figure"}]]], "page_244": [["block_0", [{"image_0": "244_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "244_1.png", "coords": [25, 494, 249, 653], "fig_type": "figure"}]], ["block_2", [{"image_2": "244_2.png", "coords": [25, 308, 284, 383], "fig_type": "figure"}]], ["block_3", ["230\nPolymer Conformations\n"]], ["block_4", ["variable can range over several orders of magnitude. If the functional relation is a power law, then\nin this format the plot will be a straight line and the slope gives the power law exponent.\n"]], ["block_5", ["So far we have considered chain dimensions solely in terms of the average end-to-end distance.\nHowever, there are two severe limitations to this approach. First, the end-to-end distance is\ngenerally very difficult to measure experimentally. Second, for many interesting polymer struc-\ntures (e.g., stars, rings, combs, dendrimers, etc.) it cannot even be defined unambiguously. The\nend-to-end distance assigns particular significance to the first and last monomers, but all mono-\nmers are of importance. A useful way to incorporate this fact is to calculate the average distance of\nall monomers from the center of mass. We denote the instantaneous vector from the center of mass\nto monomer i as 5}, as shown in Figure 6.8. The center of mass at any instant in time for any\npolymer structure is the point in space such that\n"]], ["block_6", ["Here, just as with the end-to-end vector, it is useful to take (5?) (5} 5}) in order to obtain\nan average distance rather than an average vector (which would zero by isotropy). In the second\ntransformation we have assumed equal masses, i.e., a homopolymer, and m,-=m cancels out.\n(Note that the summations run up to N, the number of monomers, and not 11, the number of\nbackbone bonds.) It is worth mentioning that the term radius of gyration is unfortunate, in that it\ninvites confusion with the radius of gyration in mechanics; the latter refers to the mass-weighted,\nroot-mean-square distance from an axis of rotation, not from a single point. However, the term\nradius of gyration in reference to Equation 6.5.2 appears to be firmly entrenched in polymer\n"]], ["block_7", ["6.5\nRadius of Gyration\n"]], ["block_8", ["where m,- is the mass of monomer 5. Note that the center of mass does not need to be actually on\nthe chain (in fact, it is unlikely to be). The root-mean-square, mass-weighted average distance of\nmonomers from the center of mass is called the radius ofgyration, Rg, or (52) 1/2, and is determined by\n"]], ["block_9", ["Figure 6.8\nIllustration of the vectors from the center of mass to monomers i and j, E} and 5}, respectively,\nand the vector from monomer 1' to monomer j, Ff}.\n"]], ["block_10", [{"image_3": "244_3.png", "coords": [40, 228, 106, 265], "fig_type": "molecule"}]], ["block_11", ["N\n1/2\nE\n.-\nJ;\n1/2\nRg:\n(52>l/2 \niii-25:2\n=\nfif\u2014XN:\n(6.52)\n"]], ["block_12", ["N\nEms;- 0\n(6.5.1)\n"]], ["block_13", ["i=1\n"]], ["block_14", [{"image_4": "244_4.png", "coords": [171, 320, 270, 364], "fig_type": "molecule"}]]], "page_245": [["block_0", [{"image_0": "245_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This equation turns out to be a useful alternative de\ufb01nition of Rg. It expresses Rg in terms of the\naverage distances between all pairs of monomers in the molecule; the location of the center of mass\nis not needed. Furthermore, Equation 6.5.8 is valid for any structure; it need not be a linear chain,\nand it need not have unperturbed dimensions.\nNow we can derive the specific result Equation 6.5.3 for the freely jointed chain by realizing that\n"]], ["block_2", ["This expression equals zero because\n"]], ["block_3", ["from Equation 6.5.1 (assuming all masses are equal). Returning to the second part of Equation\n6.5.5, and utilizing Equation 6.5.2, we obtain\n"]], ["block_4", ["where d) is the angle between E,- and 5}, and r?!- is the square distance between monomers i and j\n(see Figure 6.8). Now we take the average of each term in Equation 6.5.4, and then double sum\nover i and j:\n"]], ["block_5", ["which can be rearranged to\n: \ufb02\nEN;\nEN; (r5)\n(6.5.8)\n"]], ["block_6", ["science. A more correct description will emerge in Section 6.7, namely that (52) is the \u201csecond\nmoment of the monomer distribution about the center of mass,\u201d but this terminology is rather\nunwieldy for daily use.\nIt is clear that Rg can be defined for any polymer structure, and thus avoids the second objection\nto ([12) listed above. It can be measured directly by light scattering techniques, as will be described\nin Chapter 8, and indirectly through various solution dynamics properties, as explained in Chapter\n9, thereby avoiding the first objection. However, we went to some trouble to calculate (h2)0 for\nvarious chains, and to establish the utility of C00, b, and 6p. Is that all out the window? No, it is not.\nWe will now show that is, in fact, very simply related to (h2)0 for an unperturbed linear chain,\nnamely\n"]], ["block_7", ["Consider the dot product of the vectors from the center of mass to any two monomers i andj. By\nthe law of cosines\n"]], ["block_8", ["Radius of Gyration\n231\n"]], ["block_9", [{"image_1": "245_1.png", "coords": [38, 477, 229, 519], "fig_type": "molecule"}]], ["block_10", [{"image_2": "245_2.png", "coords": [39, 177, 154, 214], "fig_type": "molecule"}]], ["block_11", [{"image_3": "245_3.png", "coords": [42, 533, 142, 574], "fig_type": "molecule"}]], ["block_12", ["1 [3,? +512 \u2014r-2]\n(6.5.4)\n\ufb01'\u00e9\u2018IrZIE'z-llilws\u00e9:\n.\n2\nJ\n"]], ["block_13", ["N\nN\n1\nN\n1\nN\n1\nN\nN\n2245'?\u00bb25233536226372Z=0\n(65-5)\n"]], ["block_14", ["ii@ '51) \n(656)\n\u00a321\n"]], ["block_15", ["(r3) : |i \u2014j|b2\n(6.5.9)\n"]], ["block_16", ["g \n= \u2014 = \u2014\n(6.5.3)\n"]], ["block_17", [{"image_4": "245_4.png", "coords": [55, 405, 262, 440], "fig_type": "molecule"}]], ["block_18", ["121\nj=l\n"]], ["block_19", ["N\n"]], ["block_20", [{"image_5": "245_5.png", "coords": [88, 335, 371, 389], "fig_type": "molecule"}]], ["block_21", [" <53>25:: \n(6.5.7)\n"]], ["block_22", ["I:\nj:\n"]]], "page_246": [["block_0", [{"image_0": "246_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "246_1.png", "coords": [26, 76, 248, 130], "fig_type": "molecule"}]], ["block_2", ["which comes from the fact that there are 2(N\u20141) terms where\n|1' \u2014-j| 1; 2(N 2) terms where\nIi \u2014j| 2; 2(N 3) terms where 11' \u2014j| = 3; etc. (If this seems mysterious, draw an N x N matrix,\nwhere the rows are numbers 1': 1 . . .N and the columns are j 1 ... N. For each matrix element,\nenter If \u2014j|. There will be N 0\u2019s along the main diagonal, N 1 1\u20193 immediately adjacent to [and\non both sides of] the main diagonal, N\u20142 2\u2019s in the next place over, etc.). Therefore\n"]], ["block_3", ["232\nPolymer Conformations\n"]], ["block_4", ["or, in other words, (\u20193) represents the end-to-end mean\u2014square distance between any pair of\nmonomers 1' andj separated by k Ii j | links. We can thus write\n"]], ["block_5", ["A useful rule of thumb for polymers15 that Rg is about 100 A when M\u2014\n\u2014 105 g./mol This number\ncan be used to estimate R3 for any other M by recalling the pr0portionality of R3 to M112.Use the\ndata in Table 6.1 to assess the reliability of this rule of thumb.\n"]], ["block_6", ["We can take the values for ghz)O/M directly from the fifth column and multiply each one by 105.\nThe largest will be 125,000 A2 for polyethylene, and the smallest will be 42,500 A2 for poly(methyl\nmethacrylate). Then we need to divide by 6 and take the square root to obtain R.\n"]], ["block_7", ["The straight line fit to the data gives Rg =0.25 M051.To make things convenient, we can choose\nN\u2014\n\u2014 105, for which M\u2014\n\u2014 104 x 105\u2014\n-\u2014 104 X 107 g/mol. From the fitting equation we obtain\nRg =948 A. The number of backbone bonds :1 =2 x 105, and we use a more precise estimate\nof the bond length of 1 5.3 A. From Equation 6.3.2 and Equation 6.5.3, then\n\ufb02_\n6 x (948)\n_\n= 11.5\n462\n2 x 105(153)2\nCOO:\n"]], ["block_8", ["where once again the last formula applies in the high N limit.\n"]], ["block_9", ["All of the other polymers in Table 6.1 will give values between these two. We may conclude that the\nrule of thumb is reliable to at least one significant \ufb01gure, and is better than that for many polymers.\n"]], ["block_10", ["Use the experimental data for Rg for polystyrenes dissolved in cyclohexane in Figure 6.9 to\nestimate Coo, 6p, and b. Note that these data are for remarkably large molecular weights.\n"]], ["block_11", ["Example 6.2\n"]], ["block_12", ["Solution\n"]], ["block_13", ["Solution\n"]], ["block_14", ["Example 6.3\n"]], ["block_15", [{"image_2": "246_2.png", "coords": [42, 635, 209, 672], "fig_type": "molecule"}]], ["block_16", [{"image_3": "246_3.png", "coords": [42, 190, 249, 231], "fig_type": "molecule"}]], ["block_17", ["2\nN-l\nRg= ZNb\u2014622Ii\u2014J\n(6.5.10)\n"]], ["block_18", ["For polyethylene 11g_\u2014(125,\u2014_\u2014000/6)\u2018/2 144 A\n"]], ["block_19", ["For poly(methyl methacrylate), Rg (42,500/6)\u201d84 A\n"]], ["block_20", ["N2(N-\u20141)\nN(N\u20141)(2N\u20141)\n:=;_:{N;1_;kz}=;_:{\n2\n6\n}\n"]], ["block_21", ["\u2014 1) \u2014 (N \u20141)(2N2\n\u2014N)}\n(6.5.11)\nN2\n6\n_N62\nb2\nN62\n\u201cYT\u2014671\u2019\u201d?\n"]], ["block_22", ["_ \n{3N2(N\n"]], ["block_23", [{"image_4": "246_4.png", "coords": [72, 86, 238, 117], "fig_type": "molecule"}]], ["block_24", ["k=1\ni=lj=\n"]]], "page_247": [["block_0", [{"image_0": "247_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "247_1.png", "coords": [24, 529, 256, 578], "fig_type": "molecule"}]], ["block_2", [{"image_2": "247_2.png", "coords": [30, 407, 188, 441], "fig_type": "molecule"}]], ["block_3", ["These values are systematically larger than those given in Table 6.1. Part of this difference may be\nattributed to experimental uncertainty, but most of the difference stems from the fact that the\npolystyrene data in Table 6.1were obtained from molten polystyrenes, whereas the data in Figure\n6.9 are for dilute solutions. Although the chain dimensions in a dilute theta solution and in the bulk\nboth increase as M 2, the prefactor (e.g., C0c,) can be slightly different. In fact, the dimensions of a\ngiven polymer may differ by as much as 10% between two different theta solvents.\n"]], ["block_4", ["Figure 6.9\nRadius of gyration for very high molecular weight polystyrene in cyclohexane at the theta\ntemperature. (Data from Miyake, Y., Einaga, Y., and Fujita, M., Macromolecules, 11, 1180, 1978.)\n"]], ["block_5", ["Also from Equation 6.3.2 and Equation 6.5.3,\nz\ufb01R =\u2014\u2014\u2014\u2018/3\nx/N\ng\n\\/1><105\n"]], ["block_6", ["This result can be derived from the expression for (I12), Equation 6.4.11, by way of the relation\n6.5.8 and a transformation of sums to integrals; this is left as Problem 6.8. An example of the\napplication of the worm-like chain model is shown in Figure 6.10. The material is poly(n-hexyl\nisocyanate) (see Figure 6.6) dissolved in hexane and the coil dimensions were measured by light\nscattering (see Chapter 8). The smooth curve corresponds to Equation 6.5.12 with a persistence\nlength of 42 mn, and the contour length determined as L (nm) =M (g/mol)/715 (g/mol/nm). The\nfactor of 715 therefore re\ufb02ects the molar mass per nanometer of contour length. The correspond\u2014\nence between the data and the model is extremely good, except for the two very highest M samples.\n"]], ["block_7", ["Finally, from Equation 6.4.5b we have\n"]], ["block_8", ["Radius of Gyration\n233\n"]], ["block_9", ["H, (A)\n"]], ["block_10", [{"image_3": "247_3.png", "coords": [37, 356, 205, 394], "fig_type": "molecule"}]], ["block_11", ["4 x 103 \n-\nPolystyrene in cyclohexane at 345\u00b0C\n"]], ["block_12", ["2x103-\n-\n"]], ["block_13", ["3 x 102\n-\nHg = 0.25 lug-51\n"]], ["block_14", ["6x102\u2014\n_\n"]], ["block_15", ["b\nx948=7.3.&\n"]], ["block_16", ["The worm-like chain of Section 6.4.2 also has an expression for R2, which is\n"]], ["block_17", [{"image_4": "247_4.png", "coords": [46, 40, 287, 297], "fig_type": "figure"}]], ["block_18", ["3x103-\n-\n"]], ["block_19", ["1x103\n\u2014\n"]], ["block_20", ["g \n5\np\n\u2014\u2014\np \nE\u2014z\u2014(exp[\u2014\u2014L/\u20acp]\n\u2014 1) +7\n(6.5.12)\n"]], ["block_21", ["8X1051X107\n2x107\n4x107\n6X107\nMW\n"]], ["block_22", ["6\n1053\no\n"]]], "page_248": [["block_0", [{"image_0": "248_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We have derived results for (h2)0 for coils, rods, and everything in between. The conformation is\nassumed to be determined by the steric constraints induced by connecting the monomers chem-\nically. But we could also think about the molecule more abstractly, as a series of freely jointed\neffective subunits. Now suppose we have the ability to \u201cdial in\u201d a through-space interaction\nbetween these subunits, an interaction that might be either attractive or repulsive. The former\ncould arise naturally through dispersion forces, for example; all molecules attract one another in\nthis way, and if we put our chain in a vacuum, those forces might dominate. If the attractive\ninteraction were sufficiently strong, the chain could collapse into a dense, roughly spherical ball or\nglobule. Repulsion could arise if each subunit bore a charge of the same sign, a so-called\npolyelectrolyte. This is commonly encountered in biological macromolecules, DNA for example.\nIf the repulsive interaction were sufficiently strong, the chain could extend out to be a rod. It is\ninstructive to think of the globule, coil, and rod as the three archetypical possible conformations of\na macromolecule, and for many systems coil\n<\u2014> globule and coil\n<\u2014> rod transitions are experi-\nmentally accessible. For example, proteins in their native state are often globular, but upon\ndenaturing the attractive interactions that cause them to fold are released, and the molecule\nbecomes more coil-like. Similarly, a synthetic, neutral polymer dispersed in a bad solvent will\ncollapse into a globule when it precipitates out of solution. A relatively short DNA double helix is\n"]], ["block_2", ["Figure 6.10\nRadius of gyration versus molecular weight for poly(n-hexyl isocyanate) in hexane. The curve\ncorresponds to the worm-like chain model with a persistence length of 42 nm. (Data from Murakami, H.,\nNorisuye, T., and Fujita, H., Macromolecules, 13, 345, 1980.)\n"]], ["block_3", ["This deviation may be attributed to the onset of excluded volume effects, whereby the coil\nconformations are larger than anticipated by the freely jointed chain (Gaussian) limit.\n"]], ["block_4", ["6.6\nSpheres, Rods, and Coils\n"]], ["block_5", ["234\nPolymer Conformations\n"]], ["block_6", ["\u201cE\n104\n_\u2014\n1\n5\n_\n_\n"]], ["block_7", ["cum\n\"\n'\n0:\nI\nZ\n"]], ["block_8", [{"image_1": "248_1.png", "coords": [48, 40, 333, 309], "fig_type": "figure"}]], ["block_9", ["1\u00b0\n:\nWorm-like chain with 6,, = 42 nm\n3\n"]], ["block_10", ["1C)5\n.\u2014\n\u2014_\n"]], ["block_11", ["-\nJ\n"]], ["block_12", ["-\n"]], ["block_13", ["_\n.\n_\n"]], ["block_14", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n1\n.I\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_15", ["Poiy(n-hexy| isocyanate) in hexane\n"]], ["block_16", ["105\n1o6\n107\n"]], ["block_17", ["1\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_18", ["M (g/mol)\n"]], ["block_19", ["-\n"]]], "page_249": [["block_0", [{"image_0": "249_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "249_1.png", "coords": [13, 607, 243, 680], "fig_type": "figure"}]], ["block_2", ["and thus the radius goes as (volume)\u201d3. For the unperturbed coil V: 1/2, as we have seen.\nHowever, we will find out in Chapter 7 that in a good solvent V as 3/5 due to the excluded volume\neffect. For a rod, clearly I\u00bb 1. Equation 6.6.1 is an example of a scaling relation; it expresses the\nmost important aspect of chain dimensions, namely how the size varies with the degree of\npolymerization, but provides no numerical prefactors. The value of the exponent is universal, in\n"]], ["block_3", ["reasonably rod-like (Epw600 A), but if the double helix is denatured or \u201cmelted,\u201d the two\nseparated strands can become coils. A poly(carboxylic acid) such as poly(methacrylic in\nwater will have a different density of charges along the chain as the pH is varied. At high pH,\nvirtually every monomer charge and although the chain willout\ninto a completely rigid, chain, it will show a size that scales linearly\nwith M rather than as \\//1_/I. In short, polymers can adopt conformations varying from spheres\nthrough \ufb02exible coils to rods.\nThe scaling of the size with molecular weight is quite different in each case. We can encompass the\nvarious possibilities by writinga proportionality between the size and the degree of polymerization:\n"]], ["block_4", ["the sense that any particular value of 12 (1/3, 1/2, 3/5, or 1) will apply to all molecules in the\nsame class.\nAs illustrated at the very beginning of this chapter, the size of these various structures (globule,\ncoil, and rod) would be very different for a given polymer. Then we considered a polyethylene\nmolecule, which in the liquid state will always be a coil. Now consider a representative DNA from\nbacteriophage T2. It has a contour length of 60 um or almost 0.1 mm. As a coil, it should have\n# (Lk\n/6)1/2 N 1 mm or 10,000 A. With MN 108 and assuming a density of 1 g/mL, it would\nform a dense sphere with a radius of 340 A. It15 an amazing fact that the bacteriophage actually\npackages the DNA molecule to almost this extent. As shownin Figure 6.11, upon experiencing an\nosmotic shock the bacteriophage releases the DNA, which had been tightly wound up inside its\nhead. The mechanism by which the DNA is packed so tightly remains incompletely understood. It\nis particularly remarkable, given that DNA carries negative charges all along its contour, which\nshould create a strong repulsion between two portions of helix. This example also underscores\nanother important point about chain dimensions: they can be very sensitive to the environment of\nthe molecule, and not only to the intramolecular bonding constraints.\n"]], ["block_5", ["For the globule or dense object,V 1/3. The volume occupied by the molecule is proportional to N\n"]], ["block_6", ["I) will have an end-to\u2014end vector, h, lying between h and h + dh, as illustratedin Figure 6.12a. In\nother words, if the start of the chain defines the origin, we want the probability that end\nfalls in an infinitesimal box with coordinates between .1: and x + dx, y and y + dy, andandz + dz.\nFrom such a function, we will be able to obtain related functions for the probability P(N, h) dh that\nthe same walk has an end-to\u2014end distance, /1 [/1], lying between 11 and h + dh, and the probability\np(N, r) dr that a monomer will be found between a distance r and r + dr from the center of mass. It\nturns out that all of these distributions are approximately Gaussian functions, just like familiar\nnormal distribution for error analysis. In particular, the answer for P(N, It)IS\n"]], ["block_7", ["So far in this chapter we have only considered the average size and conformation of a polymer.\nNow we will \ufb01gure out how to describe the distribution of sizes or conformations for a particular\nchain. We seek an expression for the probability P(N, h) dh that a random walk of N steps of length\n"]], ["block_8", ["6.7\nDistributions for End-to-End Distance and Segment Density\n"]], ["block_9", ["Distributions forSegment \n235\n"]], ["block_10", ["a result that we will now derive.\n"]], ["block_11", ["R, N\u201d\n(6.6.1)\n"]], ["block_12", ["_ \n2N!)2\n(6.7.1)\n#\n3/2\nP(N,h) [ZaZ]\nex\n"]], ["block_13", [{"image_2": "249_2.png", "coords": [56, 614, 209, 654], "fig_type": "figure"}]]], "page_250": [["block_0", [{"image_0": "250_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "250_1.png", "coords": [30, 477, 367, 636], "fig_type": "figure"}]], ["block_2", ["236\nPolymer Conformations\n"]], ["block_3", ["Kleinschmidt, A.K., \n"]], ["block_4", ["Figure 6.11\nThe DNA within a single \n"]], ["block_5", ["We begin with a one-dimensional \n"]], ["block_6", ["6.7.1\nDistribution of the End-to-End \n"]], ["block_7", ["step we go a distance\n"]], ["block_8", [{"image_2": "250_2.png", "coords": [37, 476, 185, 634], "fig_type": "figure"}]], ["block_9", ["Figure 6.12\nA \ufb02exible coil with one \n"]], ["block_10", ["a spherical shell of volume 4nh2\n"]], ["block_11", [{"image_3": "250_3.png", "coords": [202, 479, 321, 625], "fig_type": "figure"}]], ["block_12", ["(b)\n"]], ["block_13", ["m\\\n"]]], "page_251": [["block_0", [{"image_0": "251_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "251_1.png", "coords": [25, 264, 187, 307], "fig_type": "molecule"}]], ["block_2", ["We insert a proportional sign here to allow for the appropriate normalization (see below). Now\nto convert to a three-dimensional N-step random walk, we take N/3 steps along x, N/3 along y, and\nNB along 2', and recognize that the probabilities along the three directions are independent.\n"]], ["block_3", ["or\n"]], ["block_4", ["The expression is simplified by means of Stirling\u2019s large N approximation, namely\n"]], ["block_5", ["and realize that (x/bN) \u2014> 0 as N gets very big. Thus ln(l +x) sex applies\n"]], ["block_6", ["which gets rid of the nasty factorials. (It is worth noting that Stirling\u2019s approximation is excellent\nwhen N is on the order of Avogadro\u2019s number, but it is not quantitatively accurate for N m 100;\nnevertheless these errors largely cancel in deriving the Gaussian distribution.) Utilizing this,\nEquation 6.7.4 can be expanded:\n"]], ["block_7", ["after some algebra. Now we recall the expansion of ln(1 +x) when x << 1, namely\n"]], ["block_8", ["This expression arises as follows. The probability of a sequence of events that are independent is\nequal to the product of the probabilities of each event; this gives the factors of (1/2)\u201d and (1/2)?\nHowever, this would underestimate what we want, namely a net displacement of x. This is because\n(1/2)p(1/2)q assumes we have all p+ steps in succession, followed by (3\u2014 steps, whereas in fact the\norder of the individual steps does not matter, only the total p and q. There are N possible choices\nfor the first step, then N l for the next and so on, which increases the total probability by a factor\nof N!. This, however, now overestimates the answer, because all of the p+ steps are indistinguish-\nable, as are all of the q\u2014 steps. There are p! possible permutations of the+ steps, all of which\nwould give the same answer, and similarly q! permutations of the steps; both of these factors are\ncounted in N!, so we have to divide by them out. Thus we arrive at Equation 6.7.2.\nNow we make the simple substitutions\nFilm! Far!)\n"]], ["block_9", ["to obtain\nr\n1\nN\n2!\n2!\nP(N,x) _ (2) N!\n(ME)! (N g)!\n(67.4)\n"]], ["block_10", ["being 1/2. At the end of the N steps, let p be the total number of + steps and q be the total number\nof steps. Clearly N p + q and the net distance traveled will be x =b( p q). The probability of\nany given outcome for this kind of process is given by the binomial theorem:\n"]], ["block_11", ["Distributions for End-to-End Distance and Segment Density\n237\n"]], ["block_12", [{"image_2": "251_2.png", "coords": [41, 95, 162, 135], "fig_type": "molecule"}]], ["block_13", ["P N\n\"x2\n6 7 9\n(\n,X) exp W\n(\n-\n-\n)\n"]], ["block_14", ["1\nl\nln(l+x):x\u2014\u00a7x2+\u00a7x3~-\n(6.7.7)\n"]], ["block_15", ["1nP(N,x) \u2014% (N + 2) 1%! + 3%) g (1 biN) 1%! {9%)\n(6.7.6)\n"]], ["block_16", [{"image_3": "251_3.png", "coords": [46, 311, 218, 355], "fig_type": "molecule"}]], ["block_17", ["1\nP\n1\nq N!\nP(N,x)= (2) (5)\npl\u2014q!\n(6.7.2)\n"]], ["block_18", ["In N! e N In N N\n(6.7.5)\n"]], ["block_19", ["lnP(N,x) : -%(N+g) (7%) \u00a3073) (Nib)\n_x2\n(6.7.8)\nI 2s\n"]], ["block_20", [{"image_4": "251_4.png", "coords": [147, 444, 324, 475], "fig_type": "molecule"}]]], "page_252": [["block_0", [{"image_0": "252_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "252_1.png", "coords": [10, 59, 291, 117], "fig_type": "molecule"}]], ["block_2", ["and thus from the triple integral of Equation 6.7.11 we find that the result is [3/21TNb2]_3/2.\nTherefore we need to multiply our exponential factor by [3/2'rrN1'32]3\u201d2 to satisfy Equation 6.7.11\nand arrive at the result given in Equation 6.7.1.\nThis Gaussian distribution function for I; is plotted against IE] in Figure 6.13a. Although it is a\ndistribution function for a vector quantity, it only depends on the length of h; this is a natural\nconsequence of assuming that x, y, and 2 steps are equally probable. It is peaked at the origin,\nwhich means that the single most probable outcome is I; 0, Le, the walk returns to the origin.\nHowever, because of the prefactor, the probability of this particular outcome shrinks as N2, even\nthough it is more likely than any other single outcome. Finally, this expression for P(N, I?) was\nobtained by assuming large N (no surprise here). How large does N have to be for the Gaussian\nfunction to be useful? It turns out that even for N a: 10, the real distribution for a random walk\nlooks reasonably Gaussian. Already for very small N it is symmetric and peaked at the origin, but it\n"]], ["block_3", ["where we have inserted |h|2 :x2 + y2 + 22. The final step is the normalization, which accounts for\nthe fact that if we look over all space for the end of our walk, we must find it exactly once. This is\nexpressed by\n"]], ["block_4", ["Therefore\n"]], ["block_5", ["Figure 6.13\nGaussian probability distributions for a chain of N steps of length I), plotted as (a) the\nprobability of an end\u2014to-end vector k versus lit] and (b) the probability of an end-to-end distance 11.\n"]], ["block_6", ["You can find in a table of integrals that\n"]], ["block_7", ["238\nPolymer Conformations\n"]], ["block_8", [{"image_2": "252_2.png", "coords": [38, 218, 155, 257], "fig_type": "molecule"}]], ["block_9", [{"image_3": "252_3.png", "coords": [38, 428, 262, 621], "fig_type": "figure"}]], ["block_10", ["a><10r6\n\u2014\na4::\n113\u2018\n\u00a30.02\u2014\n-l\ng\n.5\n_\nIf?\nD.\n6X10\ng\nL\n_.1\nEL\n*5\n_.\n4X10\n0.01 k\n\u2018\n"]], ["block_11", ["a\nN\nN\nN\nh :P \nP\n\u2014\nP\n\u2014\nm\nPW,\n)\n(3,26)\n(3.y)\n(3.2)\next)\n"]], ["block_12", ["\u201400 \u2014oo \u201400\n"]], ["block_13", ["r-OO\n"]], ["block_14", [{"image_4": "252_4.png", "coords": [46, 71, 258, 109], "fig_type": "molecule"}]], ["block_15", ["1.2)(10\u20145\nl\nl\nl\nr\nl\n_l\u2014\nI\nl\n"]], ["block_16", ["000000\n"]], ["block_17", ["00\n"]], ["block_18", ["J\nJ\nJP(N,1?)dx,dy,dz=1\n(6.7.11)\n"]], ["block_19", ["J exp(\u2014kx2) dx \ufb02\n"]], ["block_20", ["2x10\u201d6\n_\n__\n_\n"]], ["block_21", [{"image_5": "252_5.png", "coords": [53, 416, 431, 627], "fig_type": "figure"}]], ["block_22", ["0.03L\n1><1O\u20185\n-\nN=1000\n7\n"]], ["block_23", ["(a)\n(b)\n"]], ["block_24", ["O\n0\nI\n1\n1\n"]], ["block_25", ["100\n0\n20\n40\n50\n80\n100\n"]], ["block_26", ["\u20183lh12\n2s\n(6.7.10)\n"]], ["block_27", [{"image_6": "252_6.png", "coords": [238, 421, 435, 631], "fig_type": "figure"}]], ["block_28", ["#\nb=l\n_\n"]]], "page_253": [["block_0", [{"image_0": "253_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Note that the normalization integral goes from 0 to 00, as It cannot be negative.\nThis distribution function is plotted in Figure 6.13b and is rather different from P(N,}i). In\nparticular, it vanishes at the origin and has a peak at a finite value of h before decaying to zero as\nN \u2014> 00. The fact that it vanishes at the origin is due to multiplying the exponential decay by hz.\nThere is thus a big difference between finding the most probable vector position (which is the\norigin) and finding the most probable distance (the position of the peak in P(N,h), see Problem\n6.10). You have probably encountered this contrast before, for example, in the radial distribution\nfunction for the s electrons of a hydrogen\u2014like atom, or for the Maxwell distribution of molecular\nvelocities in a gas. In the former, the most probable position of the electron is at the nucleus, but\nthe most probable distance is a finite quantity, the Bohr radius. In the latter, the most probable\nvelocity in the gas is zero, but the most probable speed is finite.\nAs a simple application of this distribution function, we can ask what is the mean square\nvalue of h?\n"]], ["block_2", ["This function is already normalized correctly. It should satisfy the one-dimensional integral\n"]], ["block_3", ["and you can show that it does, armed with the knowledge that\n"]], ["block_4", ["will be \u201cbumpy\u201d because has a discrete value. As N gets larger, the fact that b is discrete\nbecomes less and less although the fact that we are representing a discrete function by a\ncontinuous one never away for finite N. Indeed, we can assert that even though the Gaussian\ndistribution may provide numerical answers that are very accurate, it can never be exactly\ncorrect for a real chain. That is because a real chain has a finite contour length that Hi] can never\nexceed, whereas Equation 6.7.1 provides a finite probability for any value of Iii] all the way out\nto infinity.\n"]], ["block_5", ["We now turn to obtaining P(N, h) for the distance h. The transformation is illustrated in Figure\n6.12b. We consider a spherical shell at a distance 11 from the origin. It has a surface area 4'rrh2 and a\nthickness dh, so its volume is 41't dh. Any walk whose end\u2014to\u2014end vector 1; lies in this shell will\nhave the same length h, so the answer we seek is just\n"]], ["block_6", ["6.7.2\nDistribution of the End-to-End Distance\n"]], ["block_7", ["Distributions for End-to-End Distance and Segment Density\n239\n"]], ["block_8", [{"image_1": "253_1.png", "coords": [39, 572, 261, 677], "fig_type": "figure"}]], ["block_9", ["00\n1\n1T\nsexp(\u2014kx2) dx If I\n0\n"]], ["block_10", ["JP(N,h) dh :1\n(6.7.13)\n"]], ["block_11", ["0\n"]], ["block_12", ["P(N, h) dh :4171121007, 7?) dh\n"]], ["block_13", ["(722) Jh2P(N,h)dh\n"]], ["block_14", [{"image_2": "253_2.png", "coords": [60, 614, 247, 669], "fig_type": "figure"}]], ["block_15", [{"image_3": "253_3.png", "coords": [78, 238, 249, 285], "fig_type": "figure"}]], ["block_16", ["00\na \n(6.7.14)\n"]], ["block_17", ["0\n"]], ["block_18", ["0\n"]], ["block_19", ["\u2018\n(6.7.12)\ndh\n:4 I22 \nTr\n[211's\n"]], ["block_20", ["2\n2\n3\n3/2\n\u20143 h\n"]], ["block_21", ["3/2\n4\u2018}?\n]\nCXP 2s\n"]], ["block_22", ["2\n"]]], "page_254": [["block_0", [{"image_0": "254_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "254_1.png", "coords": [30, 584, 211, 651], "fig_type": "figure"}]], ["block_2", ["The last thing we consider in this section is the related distribution p(N, r) dr, the probability that a\nmonomer is between r and r+dr from the center of mass. It turns out that there is no simple\nanalytical expression for this distribution, even for a Gaussian chain [4], but the resulting\ndistribution is well approximated by a Gaussian:\n"]], ["block_3", ["Thus for a solid sphere\n"]], ["block_4", ["The integral in the denominator provides the necessary normalization. As an example, consider\na solid sphere with radius R0 and uniform density p0. The distribution function for p(f\u2019) is just a\nconstant, p0, for OErSRO, and O for r>R0. (Note that this simple function is p07), not p(r)\nbecause the latter must increase as r2 for r SR0; there are more monomers near the surface of the\nsphere than at the center.) Substituting p07) into Equation 6.7.18 we obtain\n"]], ["block_5", ["where we use the second moment (52) :Nb2/6 explicitly in the expression (compare Equation\n6.7.1, wheres could have been replaced by 012)). One important point is the factor ofN in front.\nThis is simply a new normalization so that\n"]], ["block_6", ["Applying this in Equation 6.7.14 we obtain\n"]], ["block_7", ["240\nPolymer Conformations\n"]], ["block_8", ["6.7.3\nDistribution about the Center of Mass\n"]], ["block_9", ["as expected. The quantity (/12) is also called the second moment of the distribution P(N, h). (You\nmay recall the discussion of moments in the context of molecular weights in Chapter 1.7.)\n"]], ["block_10", ["which re\ufb02ects the fact that when we look for monomers over all space, we must find all N of them.\nThe segment density distribution can also be used for a solid object, where there is no need to\nidentify N separate subunits. In such a case p(r) can be used to find Rg (32)\u201d2, from\n"]], ["block_11", ["where now we need to know\n"]], ["block_12", [{"image_2": "254_2.png", "coords": [35, 449, 192, 529], "fig_type": "molecule"}]], ["block_13", [{"image_3": "254_3.png", "coords": [42, 586, 205, 650], "fig_type": "molecule"}]], ["block_14", ["N,\n: N4\n\u2014\n\u2014\u2014-\n6.7.16\n6\nr)\n\u201cr i2wi\n\u201dp i2i\n(\n)\n"]], ["block_15", ["00\n3\nit\nJx4 exp(\u2014kx2) dx w\nE\n"]], ["block_16", ["<52) 000\n2 30\n(6.7.18)\nI p(r) dr\nI 411731007) dr\n0\n0\n"]], ["block_17", ["R0I 904117\u201d4 dr\nR5 5\n3\n(32) 20\n= R2/3 = 5R3\n(6.7.19)\nI p04'rrr2 dr\n0/\n"]], ["block_18", ["3\n3/23\n3\n\u20182\n2s \n2\n2\n=\n_.\n=\n.7.\n(h)\n4w[2w2]\n8(2Nb2) \ufb01<\n3 \nNb\n(6\n15)\n"]], ["block_19", ["Jp(N, r) dr N\n(6.7.17)\n"]], ["block_20", ["0\n"]], ["block_21", ["T r2p(r) dr\n3? 41174p(F') dr\n"]], ["block_22", ["0\n"]], ["block_23", [{"image_4": "254_4.png", "coords": [96, 140, 276, 174], "fig_type": "molecule"}]]], "page_255": [["block_0", [{"image_0": "255_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "255_1.png", "coords": [0, 60, 373, 304], "fig_type": "figure"}]], ["block_2", [{"image_2": "255_2.png", "coords": [28, 71, 210, 309], "fig_type": "figure"}]], ["block_3", ["This result serves to emphasize an important point that Rg re\ufb02ects an average of the monomer\ndistribution and not the total Spatial extent of the object. There are always monomers further than\nRg from the center of mass as well as monomers closer than R3 to the center of mass. Expressions\nfor Rg for other shapes are listed in Table 6.3.\n"]], ["block_4", ["We mentioned at the end of Section 6.2 that there is a further important issue in chain conforma-\ntions, that of excluded volume. The simple fact is that a real polymer occupies real Space and no\ntwo monomers in the chain can share the same location concurrently. This means that the statistics\nof the conformation are no longer those of a random walk, but rather a self-avoiding walk. This\ndifference might appear subtle, which it is, but the consequences are profound.\nMost importantly, we can no longer write down an analytical expression for any of the desired\ndistribution functions such as P(N,h). Nor can we calculate in any simple yet rigorous way the\ndependence of Rg on N. The mathematical reason for this difficulty is the way the problem\nbecomes more complicated as N increases. In the case of the random walk, the orientation of\nany link or step is dictated by random chance and its position in space is determined only by where\nthe previous link is. For the self\u2014avoiding walk, in contrast, we would need to ask where every\nprevious link is in order to establish whether a particular orientation would be allowed for the link\nin question. It would not be allowed if it intersected any previous link. Consequently, the\ncalculation becomes more and more complicated as N increases. Some very sophisticated mathe-\nmatics has been employed on this problem, but we will not discuss this at all. We can draw an\nimportant qualitative conclusion, however: the excluded volume effect will tend to make the\naverage coil size larger, as the chain seeks conformations without self\u2014intersections. From the\nmost sophisticated analysis, nNO'SBQ instead of NW; in other words the exponent\n12 from\nEquation 6.6.1 is 0.589 (although most people use 0.6 as a reasonable approximation).\n"]], ["block_5", ["1\n1\nCylinder\n5L2 +\n"]], ["block_6", ["Thin rod\n1\u201471?\nRod length L\n"]], ["block_7", ["1\nThin disk\nEr?\nDisk radius r\n"]], ["block_8", ["Nbg\nStatistical segment length b\nGaussian ring\n\u201412\u2014\nDegree of polymerization N\n"]], ["block_9", ["6.8\nSelf-Avoiding Chains: A First Look\n"]], ["block_10", ["Solid ellipsoid\nEllipsoid principal radii R1, R2, R3\n"]], ["block_11", ["Table 6.3\nFormulae for the Radii of Gyration for Various Shapes\n"]], ["block_12", ["3\nStatistical segment length b\nSolid sphere\n3R2\nSphere radius R\n"]], ["block_13", ["Structure\nR:\nParameters\n"]], ["block_14", ["_\n.\nN62\n,\n,\nGauss1an \n.?\nDegree of polymerization N\n"]], ["block_15", ["3f \nb2\nStatistical segment length b\nGaussian \nT\u2014 \nArm degree of polymerization Narm\n"]], ["block_16", ["Self\u2014Avoiding Chains: A First Look\n"]], ["block_17", [{"image_3": "255_3.png", "coords": [39, 318, 103, 356], "fig_type": "molecule"}]], ["block_18", ["3\nR, : \\n0\n(6.7.20)\n"]], ["block_19", ["\u20142\u2014\n"]], ["block_20", ["2\nCylinder radius r, length L\n"]], ["block_21", ["Number of armsf\n"]], ["block_22", ["241\n"]]], "page_256": [["block_0", [{"image_0": "256_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1.\nExperimental chain dimensions for poly(ethylene terephthalate) (PET) at 275\u00b0C are given by\n(h2)0/M @090 2&2 mol/g. Calculate Coo, the statistical segment length, and the persistence\nlength for this polymer. Based on these numbers, is PET a \ufb02exible polymer, or not? What\nwould you expect based on the molecular structure?\n2.\nResolve the paradox noted in Section 6.3: polystyrene is considered to be relatively stiff,\nand\npolyisOprene\nrelatively\n\ufb02exible,\nyet\ntheir statistical\nsegment\nlengths\nare\nalmost\nidentical.\n3.\nFor freely jointed copolymers with nA steps of length EA and nB steps of length EB \ufb01nd (I12)\n(large n limit) for strictly alternating, random, and diblock architectures. Are the answers the\nsame or different? Why?\n"]], ["block_2", ["As also mentioned in Section 6.2, there are two experimental situations in which this compli-\ncation goes away, and v = 0.5 again. One case is a molten polymer. Here a chain still cannot have\ntwo monomers occupying the same place, but there is no benefit to expanding the coil. The reason\nis that space is full of monomers, and a monomer on one chain cannot tell if its immediate neighbor\nin space belongs to a different chain, or is attached by many bonds to the same chain. Conse-\nquently, it does not gain anything by expanding beyond the Gaussian distribution. The second case\nis for a chain dissolved in a particular kind of solvent called a theta solvent. A theta solvent is\nactually a not-very-good solvent, in the sense that for energetic reasons monomers would much\nprefer to be next to other monomers than next to solvent molecules. This has a tendency to shrink\nthe chain, and a theta solvent refers to a particular solvent at a particular temperature where the\nexpansion due to the self-avoiding nature of the chain is exactly canceled by shrinking due to\nunfavorable polymer\u2014solvent interactions. We will explore this in more detail in the next chapter,\nwhere we consider the thermodynamics of polymer solutions.\n"]], ["block_3", ["1.\nA single chain can adopt an almost infinite number of possible conformations; we must settle\nfor describing the average size.\n2.\nFor any chain with some degree of conformational freedom the average size will grow as the\nsquare root of the degree of polymerization: this is the classic result for a random walk.\nFurthermore, the distribution of chain sizes is approximately Gaussian.\n3.\nThe prefactor that relates size to molecular weight is a measure of local \ufb02exibility; three\ninterchangeable schemes for quantifying the prefactors are the characteristic ratio, the statis-\ntical segment length, and the persistence length.\n4.\nChemical structures for which the chain orientation can reverse direction in about 20 backbone\nbonds or less are called \u201c\ufb02exible\u201d; much stiffer polymers are termed \u201csemi\ufb02exible.\u201d This\nlatter class is best considered through the worm-like chain model using the persistence length\nas the key measure of local \ufb02exibility.\n5.\nThe radius of gyration is the most commonly employed measure of size; it can be de\ufb01ned for\nany chemical structure and it is directly measurable.\n6.\nBecause of excluded volume, real chains dissolved in a good solvent are not random walks, but\nself-avoiding walks. The corresponding size grows with a slightly larger power of molecular\nweight. In theta solvents or in the melt, the excluded volume effect is canceled out and the\nrandom walk result applies.\n"]], ["block_4", ["In this chapter, we have examined the spatial extent of polymer chains as a function of molecular\nweight and chemical structure. The principal results are the following:\n"]], ["block_5", ["242\nPolymer Conformations\n"]], ["block_6", ["6.9\nChapter Summary\n"]], ["block_7", ["Problems\n"]]], "page_257": [["block_0", [{"image_0": "257_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["4.\nAssume a freely rotating, strictly alternating copolymer chain, with alternating bond angles 6A\nand 63 and step lengths EA and EB; find (222).\n5.\nThe unperturbed dimensions of polymer chains depend weakly on temperature. Interestingly,\nthe sign of the temperature coefficient can be either positive or negative. For example,\nfor\n100%\ncis-polybutadiene,\nd(ln (h2))/dTr\u2014e 0.0004 deg\u2018l,\nwhereas\nfor\n100%\ntrans-\npolybutadiene, the same quantity is \u20140.0006 deg\u20141. Provide an explanation for the observa-\ntion that this quantity can, in general, be either positive or negative, and speculate how your\nexplanation might apply in this particular case. Note that it would require a very careful\ncalculation to actually show why these two coefficients have different signs.\n6.\nShow that the expression for the mean-square end-to-end distance of the worm-like chain\ngiven in Equation 6.4.11 reduces to the expected answers for a random coil and a rigid rod,\nwhen the limits L >> 6p and 6p > L are taken, respectively.\n7.\nFrom the following data of KirsteT estimate the persistence length of calf thymus DNA and the\nmass per persistence length Mp:\n"]], ["block_2", ["9.\nMiyaki, Einaga, and Fujita3F reported measurements of R3 for very high M polystyrenes in\nbenzene at 25\u00b0C (a good solvent) and cyclohexane at 345\u00b0C (theta conditions). Find the\nrelation between Rg and MW in the case of benzene (the cyclohexane data were analyzed in\nExample 6.3). How does the exponent compare with expectations? If you extrapolate to lower\n"]], ["block_3", ["tRo. Kirste, Disc. Farad. Soc. 49, 51 (1970).\n"]], ["block_4", ["Problems\n243\n"]], ["block_5", ["1Y. Miyaki, Y. Einaga, and H. Fujita, Macromolecules 11, 1180 (1978).\n"]], ["block_6", ["Equation 6.5.12 is a good place to start. One approach is graphical, i.e., to compare the data\n(plotted logarithmically) against a theoretical plot of the dimensional quantities Rg/Ep versus\nM/Mp. Alternatively, the data can be fit to Equation 6.5.12 using a nonlinear regression\nroutine. However, some care must be taken in weighting the data in the fit. Is there a reason\nwhy the highest molecular weight data should be accorded less significance?\n8.\nDerive the expression for the mean square radius of gyration of the worm\u2014like chain, Equation\n6.5.12, from the corresponding expression for the mean-square end-to-end distance, Equation\n6.4.11. First show that this relation is equivalent to Equation 6.5.8.\n"]], ["block_7", [{"image_1": "257_1.png", "coords": [37, 531, 157, 576], "fig_type": "molecule"}]], ["block_8", [{"image_2": "257_2.png", "coords": [47, 232, 236, 337], "fig_type": "figure"}]], ["block_9", [{"image_3": "257_3.png", "coords": [49, 446, 141, 492], "fig_type": "molecule"}]], ["block_10", ["1\nn\nj\u2014 1\n2 _\n2\n4r6224>\nj=2\ni=1\n"]], ["block_11", ["M,\nRg (.41)\nM...\nR4251)\n"]], ["block_12", ["3.5 x 105\n450\n3.45 x 106\n1700\n4.6 x 105\n480\n4.6 x 106\n2000\n6.9 x 105\n650\n6.3 x 106\n2300\n1.15 x 106\n900\n6.9 x 106\n2500\n1.6 x 106\n1100\n9.2 x 106\n3000\n2.3 x 106\n1390\n1.35 x 107\n3600\n2.8 x 106\n1550\n"]], ["block_13", ["Then make the correspondences L = 1212, x = if, and y \ufb02? when using Equation 6.4.1 1 for (\u00a325.).\nThe final step is to equate the double sum above with the following integrals and carry out the\nintegration:\n"]], ["block_14", ["\u2014d\ne I\n0\n"]], ["block_15", ["1\ny\n"]], ["block_16", ["ml:\u2014\n"]], ["block_17", ["dx\n"]]], "page_258": [["block_0", [{"image_0": "258_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["P9P?\u201c\n"]], ["block_2", ["244\n"]], ["block_3", ["References\n"]], ["block_4", ["10.\n"]], ["block_5", ["14.\n"]], ["block_6", ["11.\n"]], ["block_7", ["12.\n"]], ["block_8", ["13.\n"]], ["block_9", ["15.\n"]], ["block_10", ["16.\n"]], ["block_11", [{"image_1": "258_1.png", "coords": [42, 85, 285, 185], "fig_type": "figure"}]], ["block_12", ["Flory, P.J., Statistical Mechanics of Chain Molecules, Wiley-Interscience, New York, 1969.\nKratky, O. and Porod, G., Rec. Trav. Chim., 68, 1106 (1949).\nKuhn, W., Kollold 2., 68, 2 (1934).\nYamakawa, H., Modern Theory of Polymer Solutions, Harper & Row, New York, 1971.\n"]], ["block_13", ["MW x 10\u20196\nR3 (nrn), Benzene\nRg (nm), Cyclohexane\n"]], ["block_14", ["56.2i 1\n506 _-_|; 10\n228 i 5\n39.5i 1\n32.0i 0.6\n23.5 i 0.5\n15.1 _-|; 0.5\n8.77i 0.3\n"]], ["block_15", ["For the Gaussian distribution function for the end-to-end distance, calculate the most\nprobable distance, the mean distance, and the root\u2014mean-square distance. Generate a good\nplot of this distribution function for some value of N and indicate where these three\ncharacteristic distances fall on the plot.\nFind the mean square radius of gyration for an infinitely thin rod of length L, with mass\ndensity (per unit length) p in the center L/2 section and mass density (per unit length) 2p for\nthe U4 sections at each end.\nFind the radius of gyration of a sphere with an outer shell of different density, where the inner\nspherical core has a radius of R1 and a density of p, and the outer shell (the \u201ccorona\u201d) extends\nto a radius of R2=2R1 and has a density equal to 2p. (This geometry is a reasonable\nrepresentation of a spherical micelle, which can be formed in a solution of block copolymers\nwhen one block is relatively insoluble; see Chapter 4.4.)\nConsider the coil dimensions of an A\u2014B diblock copolymer, with total degree of polyme-\nrization N and the fraction of A monomers given byf. Assuming that the mass of each A and\nB monomers is the same, show that\n"]], ["block_16", ["where (RE), is the mean square radius of gyration for block i and (22)AB is the mean square\nseparation of block centers-of\u2014mass.\nConsider a statistical A\u2014B 00polymer, with total degree of polymerization N and the fraction\nof A monomers given by f. Derive a simple expression for (/12) in terms of N,f, bA, and b3,\nassuming Gaussian statistics. This equation will probably not be exactly correct, in practice,\neven for large N. One reason is that thermodynamic interactions between A and B monomers\n(usually effectively repulsive) will tend to expand the chain. However, there is another\nreason, connected to the nature of the statistical length; what is it?\nA distinguished polymer scientist is reputed to have remarked \u201can infinite steel girder would\nbe a random coil.\u201d Explain the important point that this striking comment is intended to\nillustrate.\nHow does the average polymer concentration inside an individual coil vary with M in good\nand theta solvents? Estimate this concentration in g/cm3 for polystyrene with M = 106 under\nboth conditions, using Miyake et al.\u2019s data given in Problem 9. Explain why this concentra-\ntion is often referred to as the coil overlap concentration, c*.\n"]], ["block_17", [" =fA+ (1 f)AB\n"]], ["block_18", ["M, by what degree of polymerization would excluded volume increase Rg (i) by only 10%?\n(ii) by a factor of 2?\n"]], ["block_19", ["392i8\n353i7\n297_-|;9\n227i7\n164i4\n"]], ["block_20", ["87.9;l\u2014_2\n"]], ["block_21", ["183:4\n167i4\n145i?)\n116i2\n"]], ["block_22", ["Polymer Conformations\n"]]], "page_259": [["block_0", [{"image_0": "259_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Graessley, W.W., Polymeric Liquids and Networks: Structure and Properties, Garland Science, York,\n2003.\nRubinstein, M. and Colby, R.H., Polymer Physics, Oxford University Press, New York, 2003.\nVolkenstein, M.V., Configurational Statistics of Polymeric Chains, Wiley\u2014Interscience, London, 1963.\nYamakawa, H., Modern Theory of Polymer Solutions, Harper & Row, New York, 1971.\n"]], ["block_2", ["Doi, M. and Edwards, S.F., The Theory of Polymer Dynamics, Clarendon Press, Oxford, \nFlory, P.J., Statistical Mechanics of Chain Molecules, Wiley\u2014Interscience, New York, 1969.\n"]], ["block_3", ["Further Readings\n245\n"]], ["block_4", ["Further Readings\n"]]], "page_260": [["block_0", [{"image_0": "260_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this chapter we shall consider some thermodynamic properties of solutions in which a polymer\nis the solute and some low-molecular\u2014weight species is the solvent. Our special interest is in the\napplication of solution thermodynamics to problems of phase equilibrium.\nAn important fact to remember about the field of thermodynamics is that it is blind to details\nconcerning the structure of matter. Thermodynamics is concerned with observable quantities and\nthe relationships among them, although there is a danger of losing sight of this fact in the\nsomewhat abstract mathematical formalism of the subject. For example, we will take the position\nthat entropy is often more intelligible from a statistical, atomistic point of view than from purely\nphenomenological perspective. It is the latter that is pure thermodynamics; the former is the\napproach of statistical thermodynamics. In this chapter we shall make extensive use of the\nstatistical point of view to understand the molecular origin of certain phenomena. The treatment\nof heat capacity in physical chemistry provides an excellent example of the relationship between\npure and statistical thermodynamics. Heat capacity is defined experimentally as the heat required\nto change the temperature of a sample in, say, a constant-pressure experiment. Various equations\nrelate the heat capacity to other thermodynamic quantities such as enthalpy, H and entropy, S. The\nalternative approach to heat capacity would be to account for the storage of energy in molecules in\nterms of the various translational, rotational, and vibrational degrees of freedom. Doing thermo-\ndynamics does not even require knowledge that molecules exist, much less how they store energy,\nwhereas understanding thermodynamics benefits considerably from the molecular point of view.\nOne drawback of the statistical approach is that it depends on models, and models are bound to\noversimplify. Nevertheless, we can learn a great deal from the attempt to evaluate thermodynamic\nproperties from molecular models, even if the effort falls short of quantitative success.\nThere is probably no area of science that is as rich in mathematical relationships as thermo-\ndynamics. This makes thermodynamics very powerful, but such an abundance of riches can also be\nintimidating. In this chapter we assume that the reader is familiar with basic chemical and\nstatistical thermodynamics at the level that these topics are treated in undergraduate physical\nchemistry textbooks. This premise notwithstanding, a brief review of some pertinent relationships\nwill be a useful way to get started.\nNotation frequently poses problems in science and this chapter is an example of such a situation.\nOur problem at present is that we have too many things to count: they cannot all be designated n. In\nthermodynamics n is widely used to designate the number of moles and so we will adhere to this\nconvention. Since we deal with (at least) two-component systems in this chapter, any count of the\nnumber of moles will always carry a subscript to indicate the component under consideration. We\nshall use the subscript 1 to designate the solvent and 2 to designate the solute. We have consistently\nused N to designate the degree of polymerization and shall continue with this notation, although the\ndefinition of \u201cmonomer\u201d will be modified slightly.\nTo describe the state of a two-component system at equilibrium, we must specify the number of\nmoles n1 and n2 of each component, as well as the pressure p and the absolute temperature T. It is\n"]], ["block_2", ["7.1\nReview of Thermodynamic and Statistical Thermodynamic Concepts\n"]], ["block_3", ["Thermodynamics of Polymer Solutions\n"]], ["block_4", ["7\n"]], ["block_5", ["247\n"]]], "page_261": [["block_0", [{"image_0": "261_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["the Gibbs free energy, G, that provides the most familiar access to a discussion of equilibrium. The\nincrement of G associated with increments in the independent variables mentioned above is given\nby the equation\n"]], ["block_2", ["where V is the volume, 5 is the entropy, and p.) is the chemical potential of component 1'. An\nimportant aspect of thermodynamics is the fact that the state variables (in the present context,\nespecially H, G, and the internal energy U) can be expanded as partial derivatives of the\nfundamental variables. Hence we can also write\n"]], ["block_3", ["Comparing Equation 7.1.1 and Equation 7.1.2 gives\n"]], ["block_4", ["The chemical potential is an example of a partial molar quantity: p.) is the partial molar Gibbs free\nenergy with respect to component 1'. Other partial molar quantities exist and share the following\nfeatures:\n"]], ["block_5", ["4.\nA useful feature of the partial molar properties is that the property of a mixture (subscript m)\ncan be written as the sum of the mole-weighted contributions of the partial molar properties of\nthe components:\n"]], ["block_6", ["248\nThermodynamics of Polymer Solutions\n"]], ["block_7", ["where Y V, H, or 5, respectively. Except for the partial molar Gibbs free energy, we shall use\nthe notation 7; to signify a partial molar quantity, where Y stands for the symbol of the\nappropriate variable.\n2.\nPartial molar quantities have per mole units, and for Y, this is understood to mean per mole\nof component i. The value of this coefficient depends on the overall composition of the\nmixture. Thus VH20 is not the same for a water\u2014alcohol mixture that is 10% water as for\none that is 90% water.\n3.\nFor a pure component the partial molar quantity is identical to the molar (superscript A) value\nof the pure substance. Thus for pure component i\n"]], ["block_8", ["and\n"]], ["block_9", ["1.\nWe may define, say, partial molar volume, enthalpy, or entropy by analogy with Equation 7.1.5:\n"]], ["block_10", [{"image_1": "261_1.png", "coords": [41, 242, 125, 316], "fig_type": "molecule"}]], ["block_11", [{"image_2": "261_2.png", "coords": [41, 327, 119, 367], "fig_type": "molecule"}]], ["block_12", ["dG=Vdp-\u2014SdT+Zp.,-dm\n(7.1.1)\n"]], ["block_13", ["3G\n3G\n(9G\n36\ndG <\u2014\u2014)\ndp + <\u2014)\ndT \u2014|\u2014 (\u2014)\ndnl + (\u2014)\ndag\n(7.1.2)\nap\nT,n1,n2\n6T\np,m,r12\nanl\np,T,nz\nan;\np,T,n1\n"]], ["block_14", ["1/ z\n(\u201810.)\n(7.1.3)\nap\nT,m,n2\n"]], ["block_15", [{"image_3": "261_3.png", "coords": [47, 93, 180, 131], "fig_type": "molecule"}]], ["block_16", ["3G\nS = _ _\n7.1.4\n(3T) p,nl ,Hz\n(\n)\n"]], ["block_17", ["Y.- (21:)\n(7.1.6)\nam\nplay-5E,-\n"]], ["block_18", ["Ym 111.71 -|- H2372\n(1.18)\n"]], ["block_19", ["ant.\npiranja\ufb01\n"]], ["block_20", ["A\n"]], ["block_21", ["3G\n"]], ["block_22", ["i=1,2\n"]], ["block_23", [{"image_4": "261_4.png", "coords": [157, 184, 374, 222], "fig_type": "molecule"}]], ["block_24", [{"image_5": "261_5.png", "coords": [250, 186, 365, 220], "fig_type": "molecule"}]]], "page_262": [["block_0", [{"image_0": "262_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The quantity 11.? is called the standard state (superscript 0) value of #4; it is the value of 11.,- when\na; 1. Neither 11.,- nor G can be measured absolutely; we deal with differences in the quantities and\nthe standard state value disappears when differences are taken. Although the standard state is\nde\ufb01ned differently in various situations, we shall generally take the pure component as the\nstandard state and 11.? Gf\u2019.\n"]], ["block_2", ["6.\nRelationships that exist among ordinary thermodynamic variables also apply to the corre-\nsponding partial molar quantities. Two such relationships are\n"]], ["block_3", ["5.\nTo express the value of property Ym on a per mole basis, it is necessary to divide Equation\n7.1.8 by the total number of moles, n1 + 11;. The mole fraction x,- of component i is written\n"]], ["block_4", ["It is convenient to define a thermodynamic concentration called the activity a, in terms of\nwhich the chemical potential is given by the relationship\n"]], ["block_5", ["Regular solution theory illustrates how a relatively simple statistical model can provide a useful\nexpression for the free energy of mixing for a binary solution of two components. Furthermore, the\nFlory\u2014Huggins theory, which we will develop in the next section and which is the starting point for\ndiscussions of both polymer solutions and polymer blends, is really nothing more than regular\nsolution theory extended to polymers. We will derive these results using a lattice approach because\nthe physical picture is particularly clear, but it should be realized that the expression for AG,1n that\nresults could be obtained without such a seemingly artificial assumption.\nWe will designate the number of molecules of the two species ml and m2 and the number of\nmoles 121 and 122. We assume that the two molecules have equal volumes and also that their partial\nmolar volumes are equal (and concentration-independent): V1 V2. The cell size of the lattice is\nchosen to equal the molecular volume, and the lattice has a coordination number 2. The total number\nof molecules m m1 + m2 is also the total number of lattice sites. A section of a two-dimensional\nsquare lattice is illustrated in Figure 7.1.\n"]], ["block_6", ["Regular Solution Theory\n249\n"]], ["block_7", ["The entropy of mixing is obtained from the Boltzmann definition of entropy:\n"]], ["block_8", ["7.2\nRegular Solution Theory\n"]], ["block_9", ["7.2.1 Regular Solution Theory: Entropy of Mixing\n"]], ["block_10", ["521(a\n(7.2.1)\n"]], ["block_11", ["n,-\nxi \n(7.1.9)\n25:12 \"i\n"]], ["block_12", ["In this expression n, and n; are the numbers of moles of components 1 and 2 in the mixture\nunder consideration.\n"]], ["block_13", ["therefore\n"]], ["block_14", ["Ym\n_\n_\n=x1Y1+x2Y2\n(7.1.10)\n111 + n;\n"]], ["block_15", ["p,=\ufb01,\u2014\u2014T\u00a7,\n(7.1.11)\n"]], ["block_16", ["and\n_\n6\n.\nV1: (\u20141\u201c)\n(7.1.12)\n3p\nTan-#1\n"]], ["block_17", ["rt, =u?+RTlna,\n(7.1.13)\n"]], ["block_18", ["As noted above, all of the partial molar quantities are, in general, concentration-dependent.\n"]]], "page_263": [["block_0", [{"image_0": "263_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "263_1.png", "coords": [12, 35, 303, 327], "fig_type": "figure"}]], ["block_2", ["where the terms in the denominator take account of the fact that the m1 molecules of type\n1 are\nindistinguishable from one another, as are the m2 molecules of type 2 (compare this with Equation\n6.7.2). Stirling\u2019s approximation (In N I azN ln N\u2014N for large N) is used to get rid of the factorials,\njust as in Section 6.7:\n"]], ["block_3", ["Now the con\ufb01gurational entropy for each pure component (and we are assuming that configur-\national entropy is the only source of entropy) is zero:\n"]], ["block_4", ["where Q is the total number of possible configurations that the system can adopt. For placing m1\nobjects of type 1 and m; objects of type 2 on a lattice of m sites the total number of configurations is\n"]], ["block_5", ["250\nThermodynamics of Polymer Solutions\n"]], ["block_6", ["Figure 7.1\nSection of a two\u2014dimensional square lattice with each site occupied by either of two species.\n"]], ["block_7", ["I\nQ \nm-\n(7.2.2)\nmllmgl\n"]], ["block_8", ["I\n51 21min] :klnT\u2014E: 0\nm1.\n(7.2.4)\nI\n52 kind2 =k1nm\u20142;= 0\nm2.\n"]], ["block_9", ["Sm :k{lnm!~\u2014\u2014lnm1!\u2014lnm2!}\n= k{mlnm m mllnm1+ m1\u2014 m2 lnmg + m2}\n2 k{(m1+ m2)ln(m1+ m2) m1 ln m1 m2 ln m2}\n= ~\u2014k{m1(ln m1 ln m) + 1722(1a In 172)}\n= \u2014k{m11nx1 + 1722111162}\n(723)\n"]], ["block_10", ["00000000000000\n00000000000000\n00000000000000\n0000000000000.\n00000000000000\n00000000000000\n00000000000000\n00000000000000\n00000000000000\n00000000000000\n00000000000000\n0.000000000000\n00000000000000\n00000000000000\n"]]], "page_264": [["block_0", [{"image_0": "264_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We now compute the enthalpy of mixing for this model, Mn\u201c, and will assume that the resulting\nenergy term is sufficiently small that it does not matter in A5,\u201c. First, we assume that AH\u201c, = AUm,\nin other words no p V terms contribute to H; this is consistent with the lattice approach and\nemphasizes that we assume there is no volume change on mixing. We introduce the interaction\nenergies in the pure state, w\u201c and W22, which act between two molecules of type 1 and between\ntwo molecules of type 2, respectively; this is shown schematically in Figure 7.2a. All molecules\nattract one another by dispersion forces, so w\u201d and W22 are negative. Under these assumptions the\nenthalpy in the pure state is given by\n"]], ["block_2", ["which means each molecule has 2 neighbors and therefore 2 interactions, but we divide by 2\nbecause there is only w\u201d worth of interaction energy per pair of molecules. We are also assuming\nthat molecules only interact with their nearest neighbors.\nIn the mixture we assume that the placement of\n1 and 2 is completely random, as in the\ncalculation of A5,\u201c. Thus the probability that a neighboring site is occupied by a molecule of type\n"]], ["block_3", ["7.2.2\nRegular Solution Theory: Enthalpy of Mixing\n"]], ["block_4", ["It is important to recognize that Equation 7.2.5a through Equation 7.2.5d are all the same in\nphysical content, but different in units, and one must be careful to use the apprOpriate form of AS\u201c,\nin calculations.\nThere are three important features to this expression for A3,\u201c.\n"]], ["block_5", ["1.\nAs x1 and x2 are always between 0 and 1 in a mixture, the natural logarithm terms are always\nnegative and the overall AS\u201c, > 0. Therefore, configurational entropy always favors spontan-\neous mixing.\n2.\nThe expression is symmetric with respect to exchange of 1 and 2, which is a consequence of\nthe assumption of equal molecular sizes. In real mixtures this condition will be hard to satisfy.\n3.\nThis calculation of the entropy assumed that all configurations on the lattice were equally\nprobable, i.e., there was no energetic benefit or price for having 1 next to 1 and 2 next to 2,\nversus having 1 next to 2. If there were such an energy term, each configuration ought to be\nfurther weighted by the appropriate Boltzmann factor exp[\u2014E/RT], where E is the total energy\nof that con\ufb01guration.\n"]], ["block_6", ["or per mole of lattice sites (and therefore, in this case, per mole of molecules)\n"]], ["block_7", ["where we have factored out Nav. It is often useful to have an intensive expression, i.e., the entropy\nof mixing per lattice site (and therefore per molecule)\n"]], ["block_8", ["This expression applies to the entire mixture, and could also be written\n"]], ["block_9", ["so Equation is actually the change in entropy with mixing\n"]], ["block_10", ["Regular Solution Theory\n251\n"]], ["block_11", ["ASm -\u2014R{x11nx1 +x2 111.152}\n(725d)\n"]], ["block_12", ["ASm 46051111151 +x2 111362)\n(7.2.5c)\n"]], ["block_13", ["ASm \u2014R{n1 lnxl + nglnxg}\n(7.2.5b)\n"]], ["block_14", ["ASm Sm Sl 52 \u2014k(m11nx1 + m2 lnxg)\n(7.2.5a)\n"]], ["block_15", ["H2 =\n\u201dimzZs\n"]], ["block_16", ["H1 \nEmlzwn\n(7.2.6)\n"]]], "page_265": [["block_0", [{"image_0": "265_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "265_1.png", "coords": [26, 45, 226, 144], "fig_type": "figure"}]], ["block_2", ["which is the exchange energy per molecule, normalized by the thermal energy Id\". In other words,\nX is the fraction of H\" you must pay in order to lift one molecule of type 1 out of its beaker, one\n"]], ["block_3", ["We define an exchange energy Aw :(wlrwl 1/2 w22/2), which represents the difference between\nthe attractive cross-interaction of l and 2 and the average self-interaction of 1 with 1 and 2 with 2.\nFor dispersion forces, and for the regular solution theory, Aw 2 0, as we will discuss in Section 7.6;\nthis is a manifestation of \u201clike prefers like,\u201d indicating that the self-interactions are more attractive\nthan cross-interactions. For now, however, we can just take it as an energy parameter that is\nprobably positive.\nWe make a further definition, that of the interaction parameter, x:\n"]], ["block_4", ["252\nThermodynamics of Polymer Solutions\n"]], ["block_5", ["and thus\n"]], ["block_6", ["Figure 7.2\nIllustration of pairwise nearest\u2014neighbor interaction energies, WU, in (a) the pure components and\n(b) a mixture.\n"]], ["block_7", ["g\n4\u2014)-\nO\n"]], ["block_8", ["1 and 2 Wu (wlz is also negative), then (see Figure 7.2b)\n"]], ["block_9", ["1 or 2 is given by x1 or x2, respectively. If we call the interaction energy between molecules\n"]], ["block_10", ["<(-)> \nH\nH\n(a)\n14/20\nw=0\n"]], ["block_11", [".\nO .\n\u00a2W11\nW22\u00a2 \nO . O\n0\n(b)\nW12\n"]], ["block_12", ["__ \n2\u2014\n.2.9\nX\nkT\n(7\n)\n"]], ["block_13", ["1\nHnn \nEm1(zx1w\u201c + zx2w12)+ \u00a7m2(zx1w12 + zx2w22)\n(7.2.7)\n"]], ["block_14", ["AHm =Hm \u2014H1\u2014H2\n"]], ["block_15", [{"image_2": "265_2.png", "coords": [65, 453, 309, 522], "fig_type": "molecule"}]], ["block_16", [{"image_3": "265_3.png", "coords": [65, 187, 192, 272], "fig_type": "molecule"}]], ["block_17", [": \n\u00a7z{w12(m1x2 \n"]], ["block_18", ["mlmz\nml\ufb02\u2019Q\n=\u2014Z\nW12\n\u2018W11(\n)\u2018W22(\n)\n2\nm\nm\nm\n:\n2(w12 _K1_l_323) :zAw\n(7.2.8)\nm.\n2\n2\nm\n"]], ["block_19", ["1\n"]], ["block_20", ["{\n(21711\u201d?2)\n"]], ["block_21", ["0..\n000\nO\n"]], ["block_22", [{"image_4": "265_4.png", "coords": [108, 52, 226, 144], "fig_type": "molecule"}]]], "page_266": [["block_0", [{"image_0": "266_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["For the mixture of 2 mol the total free energy of mixing is therefore (\u2014 1.3 x 2 X 8.3 x 298) m -6.5\nkJ. Although we should not take the exact numbers too seriously, the overall negative sign con\ufb01rms\nour expectation that toluene and cyclohexane should be quite happy to mix, even though X is\npositive.\n"]], ["block_2", ["Assuming that w\u201c,\nW12,\nand\nW22\nare approximately\n\u20141.17 x 10\u201920,\n\u2014 1.08 x 10\u201420,\nand\n\u2014 1.01 x 10\u201420 J for toluene and cyclohexane, respectively, estimate the free energy of mixing\nfor 1 mol of toluene with 1 mol of cyclohexane at room temperature. (We will see in Section 7.6\nhow these interaction energies can be estimated from experimental measurements.)\n"]], ["block_3", ["We will first calculate a value for X, and then substitute into Equation 7.2.11a. We need to assume\na value of the coordination number, 2; it is typically about 10 for small molecule liquids (recall that\nin a close\u2014packed lattice of spheres there are 12 nearest neighbors).\n"]], ["block_4", ["per site. The first two terms represent the entropy, and, as noted above, they favor mixing, whereas\nthe last enthalpic term is assumed to be positive and therefore opposes mixing. The second form of\nA6,\u201c, Equation 7.2.11b, is probably the easier one to remember because of its obvious symmetry.\nHowever, when we subsequently compute chemical potentials, we will need to take partial\nderivatives with respect to 121 and 112, and the form in Equation 7.2.11a will be more appropriate.\nThe implications of regular solution theory, particularly in terms of the predictions for the phase\nbehavior, will be discussed in Section 7.5.\n"]], ["block_5", ["We have r11: r12: 1, and x1 =x2=O.5, so\n"]], ["block_6", ["for the whole system, or\n"]], ["block_7", ["per lattice site or molecule, and\n"]], ["block_8", ["per mole. When X is positive AH\", is positive, and therefore Opposes Spontaneous mixing. We can\nnow express the free energy of mixing, AGm AHm\u2014TASm as\n"]], ["block_9", ["for the system as a whole, or\n"]], ["block_10", ["Solution\n"]], ["block_11", ["molecule ofbeaker, and exchange them. Note that although X is dimensionless,\nits value does depend on the chosen size of the lattice site, through Aw. Using this definition we\ncan write\n"]], ["block_12", ["Example 7.1\n"]], ["block_13", ["Regular Solution Theory\n253\n"]], ["block_14", ["ZAW\n10\n1.17 +1.01\n_,0\n:_g\n\u20141.08 \u201c\n10\n20.24\nX\n H \n(1.4 x 10-23)(298){\n+\nl \n"]], ["block_15", ["AGm\nF \n"]], ["block_16", ["77\u2014,\u2014 \n(7-2-11b)\n"]], ["block_17", ["AGm\nF\n= (r11 lnxl + n; lnX2 + HIJQX)\n(72-113)\n"]], ["block_18", ["AGm\n_\n"]], ["block_19", ["AHm xlxsT\n(7.2.10c)\n"]], ["block_20", ["AHm m1X2XkT nixsT\n(7.2.10a)\n"]], ["block_21", ["AH,\u201c xlk\n(7.2.10b)\n"]]], "page_267": [["block_0", [{"image_0": "267_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "267_1.png", "coords": [24, 408, 259, 640], "fig_type": "figure"}]], ["block_2", ["Before proceeding to polymer solutions, one further comment about this derivation of AG\", is in\norder. In computing AHm we assumed completely random mixing, as in Equation 7.2.7. But, if X is\nnonzero, there will presumably be some finite preference for \u201cclustering,\u201d e.g., 1 with 1 and 2 with\n2 if x> 0. Thus the probability that a lattice site immediately adjacent to a type\n1 molecule is\noccupied by another\n1 may actually be larger than x1 and the same for component 2. This\npossibility is simply not accommodated in the model and represents a fundamental limitation.\nA theory that assumes that the local interactions are determined solely by the bulk average\ncomposition, i.e., x1 and x2 in this case, is called a \u201cmean\u2014field\u201d theory. We will encounter\nother examples of mean\u2014field theories in this book. They are popular, as one might expect, because\nthey are relatively tractable; it also tums out that in many polymer problems, especially in\nundiluted polymers, they are remarkably reliable.\n"]], ["block_3", ["Flory and Huggins independently considered AG\", for polymer solutions, and the essence of their\nmodel is developed here [1,2]. As noted in the previous section, the Flory\u2014Huggins theory is a\nnatural extension of regular solution theory to the case where at least one of the components is\npolymeric. To proceed, we will adopt the same lattice model as before, with one important\ndifference. We choose the lattice site to have the volume of one solvent molecule (subscript 1)\nand each polymer (subscript 2) occupies N lattice sites. This is illustrated in Figure 7.3. Thus N is\nproportional to the degree of polymerization, but the monomer unit is now defined to have the\nsame volume as the solvent. (Equivalently, N is the ratio of the molar volume of the polymer to that\nof the solvent.) We will refer to this subunit of the polymer as a \u201csegment.\u201d We will also switch\nfrom mole fractions to volume fractions, qbl and (152, in describing composition. We do this because\nin order to use moles accurately with polymers, one needs to know the molecular weight precisely,\nand one also has to know the full molecular weight distribution, whereas the volume fraction is\neasily obtained from the measured mass and known densities. (However, we will need to be a little\ncareful in some thermodynamic manipulations in subsequent sections, where for example the\n"]], ["block_4", ["254\nThermodynamics of Polymer Solutions\n"]], ["block_5", ["Figure 7.3\nSection of a two-dimensional square lattice with each site occupied either by a solvent molecule\nor a polymer segment.\n"]], ["block_6", ["7.3\nFlory\u2014Huggins Theory\n"]], ["block_7", ["0\nO\nO\nO\nO\nO\nO\nO\n0000\n"]], ["block_8", ["OO O\n0\n"]], ["block_9", ["00000000000\n"]], ["block_10", ["00000000\n"]]], "page_268": [["block_0", [{"image_0": "268_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This expression has the same physical content as Equation 7.3.4a, but it brings out the importance\nof N. Because N is usually a large number, the contribution to the entropy of mixing from the\npolymer is very small, often almost negligible. Thus immediately we can predict that for systems\nwith X > 0, where spontaneous mixing is driven only by A5,\u201c, it will be harder to mix a solvent with\na polymer than with another small molecule.\n"]], ["block_2", ["Herewith a more detailed derivation is presented. Let us suppose we have added 5 polymers onto\nthe lattice and we inquire about the number of possibilities for placing the (i + 1)th chain. At this\nmoment, the probability that a lattice site does not contain a monomer isf, =1 (iN/m), where m is\n"]], ["block_3", ["To obtain AS\", for the entire system, we multiply the respective terms by the number of molecules\nof each type, and sum:\n"]], ["block_4", ["If we compare this expression with Equation 7.2.5 for regular solution theory, the only change is the\nswitch to volume fractions. Furthermore, if we let N =1, then (:51 =xl and qbz=x2, and regular\nsolution theory can be recovered as a special case of the Flory\u2014Huggins model. To obtain the\nexpression for AS\", per site, we divide by m 2m1 + Nmz:\n"]], ["block_5", ["The polymer molecule has many internal degrees of freedom, as described in Chapter 6. However,\nwe can assume that the number of conformational degrees of freedom is the same in the solution as\nin the pure polymer; in other words, the polymer is the same random coil with the same segment\ndistribution function in the bulk as in solution. Thus the only configurational entropy of mixing\ncomes from the increased possibilities for placing the center of mass. The polymers occupy n\nsites in the bulk polymer, and the number of possible locations for the center of mass is thus\nproportional to Nmz. On the other hand, in the mixture the number of locations is proportional to\n"]], ["block_6", ["where the unspecified proportionality factor drops out in the ratio. A similar argument holds for\neach solvent molecule, and therefore\n"]], ["block_7", ["m1 + Nmz. Thus for one polymer molecule (recall Equation 7.2.1)\n"]], ["block_8", ["7.3.2 Flory\u2014Huggins Theory: Entropy of Mixing by a Longer Route\n"]], ["block_9", ["where the total number of lattice sites m is now given by m, +Nm2.\nThere are at least two ways to obtain the Flory\u2014Huggins entropy of mixing; one is rather\ninformal, but simple, the other is more careful in the application of the model, but\nalgebraically tedious. We shall go through both, but you may wish to skip the longer version on the\nfirst pass.\n"]], ["block_10", ["7.3.1\nFlory\u2014Huggins Theory: EntrOpy of Mixing by a Quick Route\n"]], ["block_11", ["chemical respect to the number of\nmoles.) Accordingly, the volume fractions are defined as\n"]], ["block_12", ["Flory\u2014Huggins \n255\n"]], ["block_13", ["ASH-l _k(m1 111$] + m2 ln\u00e9z) \u2014\u2014R(n1 1nd\u201d + n21n\u00a22)\n(7.3.43)\n"]], ["block_14", ["as... =\u2014k(\u00a3;1\u2014lln<;bl N;- lnaz) = \u2014k (a, ma, +% 111(1),)\n(7.3.4b)\n"]], ["block_15", ["as\u201c,l \u2014\u2014klnd>1\n(7.3.3)\n"]], ["block_16", ["ASm; kln(m1 + Nmz) klnNmz \u2014\u2014klnd>2\n(7.3.2)\n"]], ["block_17", ["a \nm\u2018\n"]], ["block_18", ["2\u2014m1+Nm2\n"]], ["block_19", ["1\u2014m1+Nm2\n731\n"]], ["block_20", [{"image_1": "268_1.png", "coords": [101, 508, 306, 538], "fig_type": "molecule"}]]], "page_269": [["block_0", [{"image_0": "269_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where the last line took a little algebra. Anyway, this now gives us .3m k In 0,0,. To obtain S for\nthe pure polymer and pure solvent, we can set either ml or m; 0 in Equation 7.3.9, respectively:\n"]], ["block_2", ["where the last transformation invokes a dilute solution approximation: iN < m always, and W > N\nfor all but a few values of i. This now allows us to replace the product of powers in the last version\nof Equation 7.3.6 with a ratio of factorials:\n"]], ["block_3", ["where the factor of l/mgl takes care of the fact that the polymers are indistinguishable. Now we\nneed to play a trick to deal with the term inside the product. Consider the following ratio of\nfactorials:\n"]], ["block_4", ["where we neglect the difference between 2 and z\u2014l in one term. The total number of configur-\nations for adding m2 polymers in succession is therefore\n"]], ["block_5", ["where the last step recognizes that m1 2m mgN. Now we are ready for help from Mr. Stirling:\n"]], ["block_6", ["again the total number of sites. Now we add the next polymer segment by segment (and assume f,- is\nconstant throughout the process). The \ufb01rst segment has 771 iN choices, the second has zf,, and the\nthird and all subsequent segments have (2 l)f,\u2014 choices. The number of configurations available to\nthe (i + 1)th chain, 0,, is thus\n"]], ["block_7", ["256\nThermodynamics of Polymer Solutions\n"]], ["block_8", [{"image_1": "269_1.png", "coords": [39, 396, 280, 431], "fig_type": "molecule"}]], ["block_9", ["m_2!( m ) \n1;! (m\u2014iN\u2014N)!:?nj( m ) \n"]], ["block_10", [{"image_2": "269_2.png", "coords": [47, 97, 225, 164], "fig_type": "molecule"}]], ["block_11", ["ln\ufb02l :mllnml \u2014m11nm1=0\n(7310)\n11102 = mzlnmzN m2 1a mgav \u20141)(1\u2014ln(z 1))\n' \u2019\n"]], ["block_12", ["ln\ufb02mt=1nm!\u2014lnm1!\u2014lnm2!+m2(N l)ln(z \nm \n= mlnm m mllnm1+m1\u2014 mzlnmz +m2 +m2(N \u20141)ln(z 1) m2(N ~1)lnm\n"]], ["block_13", ["\u201c(+1 \n"]], ["block_14", ["1\n(Hg\u20141\n1\n3712\u20141\nm iN\nN-1\n0 \u2014\u2014,\n0m = \u2014, H (m \u2014\u2014 7N)(z \u2014\u2014 1)N\"\u2018(\u2014\u2014\u2014\u2014)\n\u201d12.\ni=0\n7712.\n\u00a320\n772\n"]], ["block_15", ["(maiN)!\n__(m\u2014iN)(m\u2014iN-\u20141)---(m\u2014\u2014iN\u2014N)!\n(m\u2014-(i+1)N)! \n(m\u2014 iN\u2014N)!\n=(m\u2014iN)(m\u2014iN\u20141)---(m\u2014\u2014iN\u2014N+ 1)::s(m\u2014\u2014iN)N\n(7.3.7)\n"]], ["block_16", ["1\n2-1 m2(\"\"\u201d\u2019\"2\u201c1\n(m\u2014iN)!\n1\n2\u20141 WW\u201d)\n"]], ["block_17", [".\nN\n= 7<\nm \nH (m \u2014 7N)\n(7.3.6)\n'\ni=0\n"]], ["block_18", ["m iN\nN\u2014l\n% (m WXZ \nDIV\u20141(7)\n(73.5)\n"]], ["block_19", [":(m1+m2)ln(m1+Nm2)H m1 lnml m2 11'l m2(N -\u20141)(1\u2014ln(z 1))\n"]], ["block_20", ["1\nZ _1\nMg(N\u2014I) 012\u20141\n"]], ["block_21", [{"image_3": "269_3.png", "coords": [170, 189, 323, 228], "fig_type": "molecule"}]], ["block_22", ["#777;( m )\n(m\u2014mzN)!\ufb02\u2014m1!m2!(m )\n(7.3.8)\n"]], ["block_23", ["{m!(m\u2014N)!-~(m\u2014(m2\n"]], ["block_24", ["1\nz\u2014l\n\"MN\u2014l)\nm!\nm'\nz\u2014l\nm2(N\u2014l)\n"]], ["block_25", ["\u20141)N)!\n}\n(m \u2014N)!(m 2N)! (m mzN)!\n"]], ["block_26", [{"image_4": "269_4.png", "coords": [264, 456, 436, 491], "fig_type": "molecule"}]], ["block_27", ["(7.3.9)\n"]]], "page_270": [["block_0", [{"image_0": "270_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "270_1.png", "coords": [23, 454, 204, 485], "fig_type": "molecule"}]], ["block_2", ["per site.\nThe main features of Equation 7.3.13a and Equation 7.3.13b are that the entropy terms always\nfavor mixing, the enthalpy opposes mixing when x>0, and the big difference from regular\nsolution theory is the factor of N reducing the polymer contribution to the entropy of mixing.\nThese expressions are very powerful, as they can be used to calculate many thermodynamic\nquantities of interest. For example, in the next two sections we will develop the explicit predictions\nof the model for two experimentally important quantities, the osmotic pressure and the phase\ndiagram, respectively.\nOne further point to bring out now, however, is how problematic the mean-field assumption can\nbe for dilute polymer solutions. This is illustrated schematically in Figure 7.4, where a dilute\nsolution is pictured, along with a trajectory through the solution that happens to pass through\ntwo coils. Also shown is the \u201clocal\u201d monomer concentration along that trajectory. It has two\nregions where the trajectory passes through the coils, and the local concentration of monomers is\nsignificantly higher than the solution average, (:52. However, there are also substantial regions\nbetween coils where the actual monomer concentration is zero. In other words, chain connectivity\n"]], ["block_3", ["The expression for the enthalpy of mixing in Flory\u2014Huggins theory is exactly that for regular\nsolution theory, once the substitutions of (p1 and(152 for x1 and x2 have been made. This can be seen\nbecause the enthalpy was computed on a lattice site basis with only local interactions and the\ncalculation would not be changed by linking the monomers together. (This is not strictly true,\nbecause now for each monomer there are only 2\u20142 neighboring sites that could be occupied by\neither monomer or solvent; two sites are required to be other monomers by covalent attachment.\nHowever, as 2 and Aw do not appear independently in the final expression for A0,\u201c, but are\nsubsumed into the parameter x, and we do not know Aw exactly anyway, we ignore this\ncomplication.) Thus we can write\n"]], ["block_4", ["for the system, and\n"]], ["block_5", ["per site. Combining the expressions for AS\"1 and AH\"1 we arrive at the final result:\n"]], ["block_6", ["for the system, and\n"]], ["block_7", ["Finally, then\n"]], ["block_8", ["7.3.3\nFlory\u2014Huggins Theory: Enthalpy of Mixing\n"]], ["block_9", ["which is exactly the result obtained in the previous section. Next time, you will probably just settle\nfor the simple argument!\n"]], ["block_10", ["Flory\u2014Huggins \n257\n"]], ["block_11", ["AG,1n\nF: \n(7.3.133)\n"]], ["block_12", ["AH,\u201c cplq\ufb01zk\n(7.3.12b)\n"]], ["block_13", ["AH,\u201c \u201d11%k ma\ufb01a/RT\n(7.3.12a)\n"]], ["block_14", ["AG,1n\n(b\nH \n(7.3.13b)\n"]], ["block_15", ["ASm k{(m1 + m2) ln(m1 + mzN) m1 ln m1 m2 ln(m2N)}\n"]], ["block_16", ["N\nN\n: k{m1 Wm)\n+ m, 1,,(ml+_mz) }\nm1\nm2N\n= \u2014\u2014k{m1 lnqbl + m2 lnqbz}\n= \u2014R{n1 ms, + n; 111%}\n(7.3.11)\n"]]], "page_271": [["block_0", [{"image_0": "271_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "271_1.png", "coords": [33, 42, 219, 275], "fig_type": "figure"}]], ["block_2", ["Figure 7.4\nIllustration of the failure of the mean-field assumption in dilute solutions: The local concentra-\ntion along some arbitrary trajectory \ufb02uctuates between higher than average, inside a polymer coil, and lower\nthan average, outside a coil.\n"]], ["block_3", ["guarantees that monomers are clustered in space, whereas the model assumes that the local\nconcentration is uniformly c152 throughout the sample. This turns out to be a major limitation to\nthe quantitative application of Flory\u2014Huggins theory to dilute solutions; however, it also suggests\nthat the theory should get progressively better when the concentration is increased, and the coils\nbegin to interpenetrate. This turns out to be the case.\n"]], ["block_4", ["At this point it is worthwhile to summarize the main assumptions employed in order to arrive at the\nexpression for the free energy of mixing, Equation 7.3.13.\n"]], ["block_5", ["258\nThermodynamics of Polymer Solutions\n"]], ["block_6", ["In this section we consider the osmotic pressure, H, of a dilute, uncharged polymer solution. First,\nwe will develop the virial expansion for I], which is based on general thermodynamic principles\nand therefore completely model-independent. We will learn how measurements of H can be used\nto determine the number\u2014average molecular weight of a polymer, and how the so-called second\nvirial coef\ufb01cient, B, is a diagnostic of the quality of a solvent for a given polymer. Then we will\nreturn to the Flory\u2014Huggins theory, and see what it predicts for H and B. This will lead us to our\nfirst working definition of a theta solvent, a very important concept in polymer solutions. Finally.\n"]], ["block_7", ["7.3.4\nFlory\u2014Huggins Theory: Summary of Assumptions\n"]], ["block_8", ["99:359.\u201c\n"]], ["block_9", ["7.4\nOsmotic Pressure\n"]], ["block_10", ["1.\nThere is no volume change on mixing, and VI 171, V; \n172 are independent of concen-\ntration.\nAS\", is entirely the ideal combinatorial entropy of mixing.\nAH\", is entirely the internal energy of mixing.\nBoth AS\"1 and AH\", are computed assuming entirely random mixing.\nThe interactions are short-ranged (nearest neighbors only), isotropic, and pairwise additive.\nThe local concentration is always given by the bulk average composition (the mean\u2014field\nassumption).\n"]], ["block_11", [{"image_2": "271_2.png", "coords": [71, 45, 203, 183], "fig_type": "figure"}]]], "page_272": [["block_0", [{"image_0": "272_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "272_1.png", "coords": [19, 496, 247, 617], "fig_type": "figure"}]], ["block_2", [{"image_2": "272_2.png", "coords": [24, 425, 256, 662], "fig_type": "figure"}]], ["block_3", [{"image_3": "272_3.png", "coords": [32, 454, 214, 526], "fig_type": "figure"}]], ["block_4", ["Figure 7.5\nSchematic diagram of the osmotic pressure experiment, and the operational definition of the\nosmotic pressure, H.\n"]], ["block_5", ["Rearranging, we obtain\n"]], ["block_6", ["has a dilute polymer solution of known concentration, c (in g/mL), and therefore at the instant the\nsolution is introduced into its compartment, the solvent component has a different chemical\npotential, ul. To reach equilibrium, so that the solvent chemical potential is equal on both sides\nof the membrane, there must be a net \ufb02ow of solvent from the left compartment to the right. This\ncan be easily seen, in that the simplest way to equalize the two chemical potentials would be to\nhave equal polymer concentrations on each side. As the polymer cannot move from right to left,\nsome solvent must move from left to right. However, because the solution in the right is contained,\nthe influx of solvent increases the column of solvent in the tube, i.e., the pressure goes up. This\nincrease in pressure defines the osmotic pressure, H, and ultimately U will oppose further solvent\ntransfer. At equilibrium, then,\n"]], ["block_7", ["From Equation 7.1.12 we recall that the integrand in Equation 7.4.1 is just the partial molar volume\nof the solvent, V1. We can assume this is constant over the relatively small pressure change, H,\n"]], ["block_8", ["and temperature compartments by\na semipermeable membrane; the\nmembrane passes solvent easily, but is impermeable to polymers (e.g., because of their size).\nThere is emerging from the top of each compartment and the height of the fluid in each\ntube reflects the pressure in that compartment. The compartment on the left is full of pure solvent,\nwhich therefore its standard state chemical potential it? (T, p0). The compartment on the right\n"]], ["block_9", ["and thus\n"]], ["block_10", ["the concept of osmotic pressure will turn out later to be central to understanding scattering\nexperiments (see Chapter 8). In short, this is a very important section.\n"]], ["block_11", ["The experiment schematically Figure 7.5. A thermostated chamber at pressure p0\n"]], ["block_12", ["Osmotic \n259\n"]], ["block_13", ["7.4.1\nOsmotic Pressure: General Case\n"]], ["block_14", ["MilTaPO) m (Tape) + UV:\n(7.4.2)\n"]], ["block_15", ["P0+H\n8M1\nmm) m(m + H) MAT-.100) +l\n(7,M\n (7.4.1) \nPo\np\nT\n"]], ["block_16", ["1i1 [1,? 1.1.1 \u2014\u2014RT 1n a1 \u2014RTln71x1 e \u2014RTlnx1\n(7.4.3)\n"]], ["block_17", ["Pure solvent\n#i (T. P0)\nSolution\n"]], ["block_18", ["Semipermeable membrane\n"]], ["block_19", [{"image_4": "272_4.png", "coords": [76, 460, 191, 509], "fig_type": "figure"}]], ["block_20", ["#1 (T, P0 \u2018l' H)\n"]]], "page_273": [["block_0", [{"image_0": "273_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "273_1.png", "coords": [22, 468, 214, 680], "fig_type": "figure"}]], ["block_2", ["The last simplification in Equation 7.4.6 is equivalent to ignoring n2 relative to n1 in the\ndenominator. For a solution with c = 0.01 g/mL, M 100,000 g/mol, and V in mL, n2/V will be\n104, whereas nl/V will be about 10\u20182 (if the solvent density n 1 g/mL and molecular weight\na 100 g/mol), so this is a very reasonable approximation. Substituting Equation 7.4.6 into Equation\n7.4.5 leads to\n"]], ["block_3", ["To replace x2 with \u201cpractical\u201d polymer units, such as c, we recall that c nzM/V where M is the\nmolecular weight and V is the solution volume. Thus\n"]], ["block_4", ["where B is called the second virial coef\ufb01cient: it has units of cm3 mol/gz. The quantity B3, the third\nvirial coefficient, re\ufb02ects ternary or three-body interactions, and is important when (3 is large\nenough and when B is small enough; we will not consider it further here. Equation 7.4.7a is the\ncentral result of this section. The quantity on the left\u2014hand side is measurable; of the quantities on\nthe right, c is determined by solution preparation, and M and B are determined by examining the c\ndependence of H. In Figure 7.6a H/RT is plotted against c for three different values of B: B > 0,\n"]], ["block_5", ["Figure 7.6\nGeneric plots of osmotic pressure versus concentration for dilute polymer solutions: (a) [NET\nversus 6 and (b) H/CRT versus 0, for the indicated values of B. Note the curvature at high c clue to three-body\ninteractions (the 63 term in Equation 7.4.7a).\n"]], ["block_6", ["where we have recalled the definition of the solvent activity, al, from Equation 7.1.13, and\nrecognized that 71, the activity coefficient, will approach\n1\nin sufficiently dilute solution.\nConverting to the polymer concentration, x2 = 1 \u2014- x1, gives\n"]], ["block_7", ["260\nThermodynamics of Polymer Solutions\n"]], ["block_8", ["recalling again that ln(1 \u2014\u2014x) x \u2014x for small x. In order to allow for the effects of finite solute\nconcentration, it is customary to expand the right-hand side of Equation 7.4.4 in powers of x2, a\nvirial expansion:\n"]], ["block_9", ["(a)\n0.9/ml-\n(b)\nc,g/mL\n"]], ["block_10", [{"image_2": "273_2.png", "coords": [35, 463, 354, 625], "fig_type": "figure"}]], ["block_11", [{"image_3": "273_3.png", "coords": [36, 329, 277, 380], "fig_type": "molecule"}]], ["block_12", ["I\u2018l/FIT\nI\u2018l/cFlT\n"]], ["block_13", [{"image_4": "273_4.png", "coords": [42, 462, 178, 610], "fig_type": "figure"}]], ["block_14", ["n2\ncV/M\nN \n2\n:\n__\n7.4.\nn1+n2\n(V/v,)+(cV/M)\nM\n(\n6)\n352\n"]], ["block_15", ["H\n1\nx2\n\u2014:\u2014:'ln1\u2014x\n95:-\n7..44\nRT\nV1\n(\n2)\nV1\n(\n)\n"]], ["block_16", ["%=KZ\u2014+BV\u00a7(\ufb01)2+o--=\ufb01+3c2+33c3+m\n(7.4.7a)\n"]], ["block_17", ["\ufb01:\u2014V\u2018I+BX2+B3X2+\n\"'\n(7.4.5)\n"]], ["block_18", ["H\nx2\n2\n3\n"]], ["block_19", ["I\u2014l\u2014t\u2014II\u2014l\u2014I\u2014II\u2014I\u2014rIUIIuu'nlllluv\u2014I\u2014r\n\"\"\"\"'l\"\"\"\"'I\"\"\"\"'\nI'\n"]], ["block_20", ["I.\n"]], ["block_21", ["l-\n.r\n..\n"]], ["block_22", [{"image_5": "273_5.png", "coords": [169, 461, 393, 639], "fig_type": "figure"}]], ["block_23", [{"image_6": "273_6.png", "coords": [198, 458, 332, 610], "fig_type": "figure"}]]], "page_274": [["block_0", [{"image_0": "274_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["B = O, and B < O. In Figure 7.6b we choose a second format, obtained by dividing Equation 7.4.7a\nthrough \n"]], ["block_2", ["This format often employed in practicereveals clearly that the molecular weight determined\n"]], ["block_3", ["l.\nThe \ufb01rst term in Equation 7.4.7a, c/M, is the number of molecules per unit volume (in\nmol/cm3). Thus, in the very dilute limit, the osmotic pressure is determined only by the\nnumber of solute molecules, whatever they may be. This is characteristic of all colligative\nproperties, including H, freezing point depression, and boiling point elevation. However, as c\nincreases, there will be binary solute\u2014solute interactions. The second term, proportional to c2,\naccounts for these. Now there are three possibilities:\n2.\nIf the polymers are in a good solvent, monomers on different chains are happy enough to be\nsurrounded by solvent; when two coils approach one another, there is steric repulsion (i.e.,\nexcluded volume) and the chains separate. Therefore, the coil\u2014coil interaction is effectively\nrepulsive. This corresponds to B > 0; it means there is an even greater drive for solvent to \ufb02ow\ninto the right compartment and dilute the solution.\n3.\nConversely, if the solvent is poor, such that the polymer is barely able to stay dissolved in\nsolution, monomers find it energetically favorable to be close to other monomers. Thus when\ntwo coils approach one another, there is a tendency to cluster. In this case, B <0 and the\neffectively attractive solute\u2014solute interactions resist the uptake of further solvent.\n4.\nThere is a special intermediate case where B =0. This corresponds to a \u201cnot-very\u2014good\u2018\u201d\nsolvent where the excluded volume and relatively unfavorable solvent\u2014solute interactions\ncancel one another in the net H. This case is given a special name; following Flory, it is\ncalled a theta solvent [3]. We will explore the significance of a theta solvent in more detail\nsubsequently.\n5.\nIt is worth pointing out that there is an analogy with the van der Waals equation of state for 1\nmol of an imperfect gas:\n"]], ["block_4", ["Osmotic \n261\n"]], ["block_5", ["from plotof to the sign of the initial slope.\nSo, the experimental approach is clear enough: Measure H for a series of solutions of known c,\nwith c sufficiently low that higher order terms in the virial expansion are not too important. But,\nwhat does Equation 7.4.7a mean physically?\n"]], ["block_6", ["We noted previously that moles were a troublesome unit for polymers, in part because of the\ninevitable molecular weight distribution. Now we can address the important issue of what average\n"]], ["block_7", ["7.4.1.1\nNumber-Average Molecular Weight\n"]], ["block_8", [{"image_1": "274_1.png", "coords": [40, 510, 194, 541], "fig_type": "molecule"}]], ["block_9", ["H\n1\n__ = _\nB\n. . .\n7.4.7b\ncRT\nM \nC +\n(\n)\n"]], ["block_10", ["1\n\u2014 a RT\nP\n_ + E_L_\n5,2\n_ 2\n+ __ + . . .\n7.4.9\nRT\nV\nV2\nV3\n(\n)\n"]], ["block_11", ["in which B accounts for the excluded volume and a the intermolecular interactions. When this\nequation is expanded in a virial series in the density (UV)\n"]], ["block_12", ["the second virial coefficient is B\u2014(oz/RT). This virial coefficient vanishes at a special\ntemperature, known as the Boyle point, when the excluded volume (B) and interaction (a/\nRT) terms exactly cancel one another. Similarly, the theta temperature for a particular poly-\nmer\u2014solvent system is the temperature at which B vanishes due to a cancelation of effects from\nexcluded volume and net polymer\u2014polymer interactions. It is important to realize, however,\nthat it is not the excluded volume itself that vanishes, but just its effect on the osmotic pressure.\n"]], ["block_13", ["aV2) (V \u2014 ,3) 2 RT\n(7.4.8)\n(P \n"]]], "page_275": [["block_0", [{"image_0": "275_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "275_1.png", "coords": [30, 161, 242, 229], "fig_type": "molecule"}]], ["block_2", ["The three values of Mn are very comparable, as they should be. The values of B are consistent with\nthe notion that cyclohexane is a theta solvent for polystyrene at 345\u00b0C. The experimental value of\nB is essentially zero at this temperature, considering the experimental uncertainty. Note also how\nthe plotting format of Figure 7.7b accentuates the scatter in the data.\n"]], ["block_3", ["For each data set, divide H by CRT, where R 82.1 cm3 atm/K mol and T is in kelvin. The results are\nplotted in Figure 7.7b. Then fit each data set to a straight line by linear regression (a hand calculator\nis sufficient). The results are\n"]], ["block_4", ["0.005\n0.0061\n0.0063\n0.0067\n0.010\n0.0114\n0.0124\n0.0136\n0.015\n0.0173\n0.0187\n0.0203\n0.020\n0.0215\n0.0255\n0.0276\n0.025\n0.0276\n0.0316\n0.0354\n0.030\n0.0332\n0.0381\n0.0421\n0.040\n0.0405\n0.0502\n0.0580\n"]], ["block_5", ["To illustrate the typical magnitude of the osmotic pressure and the extraction of values of Mn and\nB, consider the following data for IT (in atm) for a polystyrene sample in cyclohexane at three\ntemperatures. The data are also plotted in Figure 7.7a.\n"]], ["block_6", ["M will the osmotic pressure experiment measure for a polydisperse sample? The answer is\nappealingly simple and rigorous; as a colligative property, H at infinite dilution depends only\non the number of solute molecules per unit volume, and therefore should determine the number-\naverage molecular weight, Mn. To see that this is indeed so, we can write for very dilute\nsolutions:\n"]], ["block_7", ["c (g/mL)\n20.0\u00b0C\n34.5\u00b0C\n50.0\u00b0C\n"]], ["block_8", ["where we recall the definition of Mn from Chapter 1. Thus a properly conducted osmotic pressure\nmeasurement can determine the absolute value of Mn.\n"]], ["block_9", ["for a distribution of molecular weights, M,-. Next we form the ratio\n"]], ["block_10", ["Solution\n"]], ["block_11", ["262\nThermodynamics of Polymer Solutions\n"]], ["block_12", ["Example 7.2\n"]], ["block_13", [{"image_2": "275_2.png", "coords": [41, 116, 152, 150], "fig_type": "molecule"}]], ["block_14", ["RTC\nc,-\n: __ :\n__\n7.4.\nH\nM\nRT:\nM1\n(\n10)\n"]], ["block_15", ["200\u00b0C:\nSlope B \u20142.0 x 10\u201c4 cm3 mol/g2\n"]], ["block_16", ["345\u00b0C:\nSlope B 1.7 X 10\u20145 cm3 mol/g2\n"]], ["block_17", ["50.0\u00b0C:\n316pe B :1.2 X 10\u20144 cm3 mol/g2\n"]], ["block_18", ["C-\u2014\u20190\nCRT\n-\nE: C;\nE\nZ n;M,-/V\n"]], ["block_19", ["mm\nH\n- _ \n"]], ["block_20", ["1/intercept :Mn 1.97 x 104 g/mol\n"]], ["block_21", ["1/intercept Mn 2.02 x 104 g/mol\n"]], ["block_22", ["1 /intercept Mn 2.00 x 104 g/mol\n"]], ["block_23", ["_ \n__\nl\n\u2014 ZniM, T Mn\n(7.4.11)\n"]], ["block_24", ["____.-\u2014\n"]]], "page_276": [["block_0", [{"image_0": "276_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "276_1.png", "coords": [25, 251, 273, 480], "fig_type": "figure"}]], ["block_2", ["To conclude this section, we return to the Flory\u2014Huggins expression for the free energy of mixing\nfor a polymer solution, Equation 7.3.13, and see what it predicts for H. From Equation 7.1.5 we can\nwrite\n"]], ["block_3", ["Figure 7.7\nOsmotic pressure for polystyrene in cyclohexane plotted as (a) H versus 0 and (b) H/CRT versus c,\nat the indicated temperatures. The straight lines are linear regression fits. The data are provided in Example 7.2.\n"]], ["block_4", ["7.4.2\nOsmotic Pressure: Flory\u2014Huggins Theory\n"]], ["block_5", ["so we need to take the derivative with respect to m of Equation 7.3.13, Le,\n"]], ["block_6", ["Osmotic \n263\n"]], ["block_7", ["(a)\nc,g/mL\n"]], ["block_8", ["\u2014B\u2014 50.0\u00b0C\n4x1o\u20145...........\no\n0.01\n0.02\n0.03\n(b)\nc, g/mL\n"]], ["block_9", ["I\u2018UCRT,\n"]], ["block_10", ["mol/g\n"]], ["block_11", [{"image_2": "276_2.png", "coords": [41, 49, 268, 247], "fig_type": "figure"}]], ["block_12", ["4.5 X 1075 \n+2000\n"]], ["block_13", ["5%:{RT(n11n\u00a21\u2014l\u2014\n712111s + 721(352X)}\n"]], ["block_14", ["5.5><1o-5 \n/\n"]], ["block_15", ["\u2014\n6\n\u201cHi/1 [.Ll M? (\u201c32; AGm)\n(7.412)\n"]], ["block_16", [{"image_3": "276_3.png", "coords": [49, 579, 207, 617], "fig_type": "molecule"}]], ["block_17", ["a\n"]], ["block_18", ["g\n0.04\".\na\nI\n.J\n(U\n:\nEl\n:\n1:\u201c\ng\na\nn\n-\n0.03\n-\nD\n.2\n_\na\nI\n.\nI\n'3'\nI\n(102 '_\na\nI\nI 20.0\u00b0C\n_'\n"]], ["block_19", ["I\nj\nE\n-\n5x10-5>\u00a3\nD\nU\na\nn\na\n\u2014-_\n"]], ["block_20", ["6X10_5\u201cHun-WHHII-Hl-m\n"]], ["block_21", ["E\nP 345\u00b0C\n-\n0.01\n\u2014\nO\n_'\n"]], ["block_22", ["0.05\n:\u2014\nu\n_'\n"]], ["block_23", ["0.07\ufb01r1www.mmlwwlwrr.\n"]], ["block_24", ["0.06 \nPolystyrene in cyclohexane\n_\n..\nE\n-\n"]], ["block_25", ["0\n_ i\n.\ni\nl\nJ\n.\n.\n.\n.\nI\n.\n.\nI\n.\n.\nI\n.\n.\n.\n. '\n"]], ["block_26", ["O\n0.01\n0.02\n0.03\n0.04\n0.05\n"]], ["block_27", [{"image_4": "276_4.png", "coords": [83, 393, 149, 439], "fig_type": "figure"}]], ["block_28", ["+34,5\u00b0C\n"]], ["block_29", ["Polystyrene in cyclohexane\n"]], ["block_30", [{"image_5": "276_5.png", "coords": [181, 171, 234, 221], "fig_type": "figure"}]], ["block_31", ["PI\u201d:\n"]], ["block_32", ["0.04\n0.05\n"]]], "page_277": [["block_0", [{"image_0": "277_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "277_1.png", "coords": [21, 412, 185, 456], "fig_type": "molecule"}]], ["block_2", [{"image_2": "277_2.png", "coords": [25, 63, 253, 140], "fig_type": "figure"}]], ["block_3", [{"image_3": "277_3.png", "coords": [28, 461, 219, 517], "fig_type": "molecule"}]], ["block_4", [{"image_4": "277_4.png", "coords": [32, 308, 202, 366], "fig_type": "molecule"}]], ["block_5", [{"image_5": "277_5.png", "coords": [32, 204, 310, 263], "fig_type": "molecule"}]], ["block_6", [{"image_6": "277_6.png", "coords": [34, 365, 136, 409], "fig_type": "molecule"}]], ["block_7", ["To find the answer, we recall \ngrazml-l\u2014Nn\ufb02\u2014mIL\n"]], ["block_8", ["Proceeding with the differentiation \n"]], ["block_9", ["254\nThermodynamics of Polymer Solutions\n"]], ["block_10", ["This equation has two important features. First, when X: 1/2, then B =0 and we have a theta\nsolvent. Thus x: 1/2 represents a second operational definition of the theta point. For X > 1/2,\nB < 0, and the solvent is poor, whereas for X < 1/2, the solvent is good. Figure 7.8 shows data for\n3(7) for several polystyrenes in cyclohexane and the theta temperature is determined to be 345\u00b0C.\nSecond, 8 is predicted to be independent of molecular weight (note that N mM in Equation 7.4.17\nand thus the M dependence cancels out). It turns out experimentally that this is not quite true; for\nexample B varies approximately as M in good solvents. This incorrect prediction is a direct\nconsequence of the mean\u2014field assumption that we discussed in the previous section.\n"]], ["block_11", ["where we choose to write everything terms polymer concentration. The expansion of\nln(1 (152) \u2014q\u00a72 chi/2 is used to get rid of logarithmkeep the dig/2 term here\nbecause of the virial expansion to second order), over to the side:\n"]], ["block_12", ["Finally, we convert from(152 to c:\n"]], ["block_13", ["similarly,\n"]], ["block_14", ["and Obtain\n"]], ["block_15", ["In this section we examine the phase behavior of a polymer solution, or, more precisely, we\nconsider the temperature-composition plane at fixed pressure and locate the regions where a\n"]], ["block_16", ["7.5\nPhase Behavior of Polymer Solutions\n"]], ["block_17", ["and thus for the Flory\u2014Huggins model,\n"]], ["block_18", [{"image_7": "277_7.png", "coords": [36, 145, 247, 196], "fig_type": "molecule"}]], ["block_19", [{"image_8": "277_8.png", "coords": [40, 63, 220, 143], "fig_type": "molecule"}]], ["block_20", ["H\n1\n(162\n1\n1\n2\n _ z :\u2014 __ _ __\n_ _\n+ . . .\nRT\nv1 N\nv1 (X\n2) $2\n(7'4\u201c)\n"]], ["block_21", ["a\na 1 \na\n\u2014\n2\n\u2014-\n\ufb02=\u2014\u2014\u2014#(\n(\u201951): \u2014i=\n952:\nQM\"\n(7.4.13b)\n8111\n6711\n8111\nN712\nI11\n'\n"]], ["block_22", ["H\nC +\n1\nV \n2 +\n_ 2 _\n_ _\n_..__C\n. . .\nRT\nM\n2\nX\n1W\n(7.4.17)\n"]], ["block_23", ["11171\n{\nm am\n122\n(\u201415%)\n}\n\u2014\u2014\u2014= \n1nd) +\u2014\n+-\u2014\u2014\n\u2014\u2014\n+\n\u2014\nRT\n1\nCb]\n\u201d1\n$2\na\n($2\n\u00a21\u00a22)X\n"]], ["block_24", ["3\u201911\n(111 + Nn2)2\n(\u201dI + \n2 = \n"]], ["block_25", ["_ _ \n2 \nM\n\u2014\nM\n(7.4.16)\n"]], ["block_26", ["1\n_ N2\n1\n\"172\n1\n(2\nM2\n2\nv1 M2\n(\n)\n"]], ["block_27", ["1\n:\u2014\nIn 1\u20141;!)\n+\u00a2 (144)+\n(1)2}\n{\n(\n2)\n2\nN\nX 2\n(7.4.14)\n"]], ["block_28", ["N112\nn1\n(7.4.133)\n"]]], "page_278": [["block_0", [{"image_0": "278_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 7.8\nSecond virial coef\ufb01cient for three polystyrenes in cyclohexane as a function of temperature.\nThe theta temperature based on these data alone lies between 34\u00b0C and 35\u00b0C. (From Yamakawa, H., Abe, F.,\nand Einaga, Y., Macromolecules, 27, 5704, 1994.)\n"]], ["block_2", ["one\u2014phase solution is stable, and where the mixture will undergo liquid-liquid phase separation\ninto two phases. We will do this first for regular solution theory, as a means to illustrate the\nvarious concepts and steps in the procedure. Then we will return to Flory\u2014Huggins theory and\nsee the consequences of having one component substantially larger in molecular weight than\nthe other.\n"]], ["block_3", ["The phase diagram for a regular solution is shown schematically in Figure 7.9. It has the following\nimportant features, which we will see how to calculate:\n"]], ["block_4", ["1.\nA critical point (Tc, x.) such that for T > Tc 3 one-phase solution is formed for all compositions.\n2.\nA coexistence curve, or binodal, which describes the compositions of the two phases xi\u2019 and xi\"\nthat coexist at equilibrium, after liquid\u2014liquid separation at some fixed T< TC. Any solution\nprepared such that (T,x1) lies under the binodal will be out of equilibrium until it has\nundergone phase separation.\n3.\nA stability limit, or spinodal, which divides the two-phase region into a metastable window,\nbetween the binodal and the spinodal, and an unstable region, below the spinodal. The\nsigni\ufb01cance of the terms metastable and unstable will be explained subsequently. Note that\nthe binodal and spinodal curves meet at the critical point.\n"]], ["block_5", ["Qualitatively, of course, we should expect one-phase behavior at high T because AS,1n > 0, and\ntherefore -\u2014TASm contributes an increasingly negative term to A6,\u201c. However, although AGm < 0\nis the criterion for spontaneous mixing, it by no means guarantees a single mixed phase, as we shall\n"]], ["block_6", ["7.5.1\nOverview of the Phase Diagram\n"]], ["block_7", ["Phase Behavior of Polymer Solutions\n265\n"]], ["block_8", ["\u201829\no x 10\u00b0\n'5\nE\nO)\nE\nO\n"]], ["block_9", ["\u201c3 \u20145 x 10'5\n-\n'\n\u2014EB\u2014 M: 9,980\n\u2018\n"]], ["block_10", ["__1.5 \nI\nI\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_11", ["-1 x 10\"4\n\u2014E\u2014 M: 40,000\n\u2018\n"]], ["block_12", [{"image_1": "278_1.png", "coords": [49, 43, 302, 295], "fig_type": "figure"}]], ["block_13", ["Polystyrenes\n'\nin\n5 10\n_'\ncyclohexane\n"]], ["block_14", ["1\nX 10\u20144\nl\nl\nI\n'\nI\nl\n|\n|\n"]], ["block_15", [{"image_2": "278_2.png", "coords": [186, 189, 278, 248], "fig_type": "figure"}]], ["block_16", ["\u2014E\u2014 M: 20,200\n"]]], "page_279": [["block_0", [{"image_0": "279_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "279_1.png", "coords": [23, 38, 242, 255], "fig_type": "figure"}]], ["block_2", ["Figure 7.9\nPhase diagram (temperature versus mole fraction of component 1) for regular solution theory.\nThe binodal (coexistence curve) separates the one-phase region at high temperature from the two-phase region\n"]], ["block_3", ["266\nThermodynamics of Polymer Solutions\n"]], ["block_4", ["at low temperature. The spinodal curve (stability limit) separates the unstable and metastable windows within\nthe two-phase region. The binodal and spinodal curves meet at a critical point.\n"]], ["block_5", ["and recall from its definition (Equation 7.2.9) that )(N l/T. These two functions are plotted in\nFigure 7.10a and Figure 7.10b, respectively. Note that both are symmetric about x1 1/2, and that\nin this format the entropy term is independent of T, whereas the enthalpy term is not (due to x).\nFurthermore, we take X > 0, as expected by the theory. In Figure 7.10c we combine the two terms,\nat two generic temperatures, one \u201chigh\u201d and one \u201clow.\u201d At the higher T, X is so small that AG\u201c.1\nlooks much like the AS\", term; it is always concave up. However, at the lower T, the larger X in the\nenthalpy term produces a \u201cbump,\u201d or local maximum in the free energy. This will turn out to have\nprofound consequences. Note that even at the lower T, AG,n < 0 for all compositions considered in\nthis example.\nPhase separation will occur whenever the system can lower its total free energy by dividing into\ntwo phases. If we prepare a solution with overall composition (x1), and then ask will it prefer to\nseparate into phases with compositions x; and xi\" , we can find the answer simply by drawing a line\nconnecting the corresponding points on the AG,n curve (i.e., AGm(xf) to AGm(xf\u2019)), as shown in\nFigure 7.11a. Because AG\u201c, is an extensive property, this line represents the hypothetical free\nenergy of a combination of two phases, xf and xi\", for any overall composition On) that lies in\nbetween. (Note that the relative proportions of the two phases with compositions x{ and x{\u2019 are\ndetermined once (x1) is selected, by the so-called lever rule.)\nWhat we now realize is that, so long as AG,\u201c is concave up, this straight line will lie above AG,1n\nat (x1) for any choice of xf and x1\u201d , and therefore phase separation would increase the free energy.\n"]], ["block_6", ["see. To begin the analysis, we resolve the two contributions to AGm/RT from regular solution\ntheory (Equation 7.2.1 1):\n"]], ["block_7", ["0\nx1\n1\n"]], ["block_8", ["AS\n_Tm xllnx1+ x2 11l\n(7.5.1)\n"]], ["block_9", ["RT\n"]], ["block_10", ["(To: Xe)\nBinodal\n,\nMetastable\n"]], ["block_11", ["Two-phase\n"]], ["block_12", ["One-phase\n"]], ["block_13", ["Spinodal\n"]], ["block_14", ["Unstable\n"]]], "page_280": [["block_0", [{"image_0": "280_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "280_1.png", "coords": [28, 240, 233, 436], "fig_type": "figure"}]], ["block_2", [{"image_2": "280_2.png", "coords": [30, 42, 227, 436], "fig_type": "figure"}]], ["block_3", ["where the second derivative can be taken with respect to the mole fraction of any component. The\nmeaning of stability is this: In any mixture at a finite temperature, there will be spontaneous, small\nlocal \ufb02uctuations in concentration 8x, such that there are small regions that have x1 bigger than the\naverage, and some regions where it is smaller. Now, by the argument given above, any such\n\ufb02uctuation will actually increase the free energy; the straight line connecting (x1) 8x1 and\n(x1) + 8.151 will fall above AGm(x1). Consequently all these \ufb02uctuations will relax back to (11).\nThe importance of these spontaneous \ufb02uctuations will be taken up again in Chapter 8, where we\nwill show how they are the origin of light scattering.\n"]], ["block_4", ["Thus \u201cconcave up\u201d gives us the criterion for stability of the one\u2014phase solution; the mathematical\nexpression of concave up is\n"]], ["block_5", ["Figure 7.10\nPredictions of regular solution theory for (a) entropy of mixing, plotted as\n\u2014ASm/R; (b)\nenthalpy of mixing, plotted as AHm/RT, for two temperatures; (c) free energy of mixing obtained by\ncombining panels (a) and (b), plotted as AGm/RT.\n"]], ["block_6", ["Phase Behavior of Solutions\n267\n"]], ["block_7", ["J\na:\nF\n2\nE\n'|\nME\n_\nTI,\n25 \u201d\"0 5 _\n j \nE:\n:\n:\n"]], ["block_8", ["AGm/RT\n9;: N\n"]], ["block_9", [{"image_3": "280_3.png", "coords": [38, 29, 232, 250], "fig_type": "figure"}]], ["block_10", ["I\n0.4 :-\n\u2014\n"]], ["block_11", [{"image_4": "280_4.png", "coords": [40, 518, 122, 563], "fig_type": "molecule"}]], ["block_12", [{"image_5": "280_5.png", "coords": [42, 39, 405, 239], "fig_type": "figure"}]], ["block_13", ["\u20140.25 l\n\u2014-\n0 8\n3\n3\n"]], ["block_14", ["_0_75 _\n_\n.\u2014\nHigh T075: 1-5)\n':\n"]], ["block_15", ["(C)\nX1\n"]], ["block_16", ["_-...II..I...II..I...\n0'40\n0.2\n0.4\n0.6\n0.8\n1\n"]], ["block_17", ["(a\nGm)\n>0\n(7.5.2)\n"]], ["block_18", ["(a)\nX1\nX1\n"]], ["block_19", ["_.1\nI\n'\nl\ni\n'\n'\n\u2014l\nI\nl\nI\nl\nl\n1\nl\nl\nI\n'\n'\n'\n-\n"]], ["block_20", ["TIP\n3x?\n"]], ["block_21", ["0\nI\n.\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI r I\nr\u2014r\nr\nJ\n1-2\n_\u2014r\nI\nI\nI\nI\nI\nI\nI\nI\nI\n.\nI\nI\nI\n.\nI\nI\nI\nI\n"]], ["block_22", ["T\n0\n0.2\n0.4\n0.6\n0.8\n1\n(b)\no\n0.2\n0.4\n0.6\n0.8\n1\n"]], ["block_23", ["-\n1\nF\n'\nT\nLow T0: = 2.5)\n\"-\n"]], ["block_24", ["High T(z= 1.5)\n"]], ["block_25", ["i\n0\n-_\n_\n"]], ["block_26", [{"image_6": "280_6.png", "coords": [229, 38, 417, 242], "fig_type": "figure"}]], ["block_27", ["I\nI\nI\nI\nI\nI\nI\n[\nI\nI\nI\nI\nI\nI\nI\nl\nl\nI\nI\n"]]], "page_281": [["block_0", [{"image_0": "281_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "281_1.png", "coords": [26, 44, 226, 321], "fig_type": "figure"}]], ["block_2", ["where we have divided by the total number of moles to get to mole fractions, and where\nAir,- it, pf. Now imagine we draw a straight line that is tangent to AG\", at some composition,\nx\u2019], as shown in Figure 7.11b. This line can be written generically as\n"]], ["block_3", ["Now consider the lower T curve in Figure 7.100, where AGm shows the bump. Here we can see that\nif we prepared a solution with (x1) somewhere near the local maximum of A0,\u201c, we could find an\nx{ and am\u201d such that the straight line between them would fall below the AG\", curve for our (x1),\nand phase separation should occur. In fact, there are many such pairs x1\u2019 and x1\" that would lower\nAGm, so which pair is chosen? We recall the criteria for phase equilibria: T and p must be identical\nin the two phases, and\n"]], ["block_4", ["The chemical potential of component 1 is the same in both phases and the chemical potential of\ncomponent 2 is equal in both phases. (Be careful with this; both relations must be satisfied\nsimultaneously, but it is not an equality between #1 and #2.) It turns out that there will be only\none solution (xl\u2019,x1\u201d) for both of these relations at a particular T, which we can identify by the\ncommon tangent construction. We can write the free energy as the mole\u2014weighted sum of the\nchemical potentials (which are the partial molar free energies, Equation 7.1.5):\n"]], ["block_5", ["Figure 7.11\nGeneric free energy of mixing versus composition curves. (a) If a solution with overall\ncomposition between xf and x1\u201d were to separate into two phases with compositions x{ and xi\u201d, the resulting\nfree energy (the dashed line) would lie above the one-phase case (smooth curve). (b) Tangent construction\nshowing how the chemical potentials of the two components may be obtained for a given composition x{.\n(c) Tangent construction finds the compositions of the two phases xf and xi\" that would coexist at equilibrium,\nfor a system with overall composition between xl\u2019 and x1\u201d . Points a and [9 denote the in\ufb02ection points of AGm,\nwhich separate the metastable (x{ < x1 < (1,!) < x1 < xi\") and unstable regions (a < x1 < 3)).\n"]], ["block_6", ["or\n(7.5.4)\n"]], ["block_7", ["7.5.2\nFinding the Binodal\n"]], ["block_8", ["0\nx1I\nX1\"\n1\n(C)\n"]], ["block_9", ["(a)\ni,\nx,\u00bb\n(b)\nX\u2018\n"]], ["block_10", ["268\nThermodynamics of Polymer Solutions\n"]], ["block_11", ["A(120(1\")\n_\n"]], ["block_12", ["3511209!) \u201d//\n"]], ["block_13", ["AGm\nAGm\n"]], ["block_14", [{"image_2": "281_2.png", "coords": [45, 46, 211, 172], "fig_type": "figure"}]], ["block_15", ["mm) mm\u201d),\n#2060 \u201c20\u201c\u201d)\n(7.5.3)\n"]], ["block_16", ["AG.\u201c xlAP\u2018q + (1 \u2014x1)A#2 A\ufb02z \"I\u2018xlmiui Ape)\n"]], ["block_17", ["AGm mApLI + ngAptz\n"]], ["block_18", ["AG\u201d,\n' \n371109\u201d)\n"]], ["block_19", [{"image_3": "281_3.png", "coords": [52, 169, 218, 296], "fig_type": "figure"}]], ["block_20", [{"image_4": "281_4.png", "coords": [79, 36, 418, 173], "fig_type": "figure"}]], ["block_21", ["A7120(1')\n"]], ["block_22", [{"image_5": "281_5.png", "coords": [227, 44, 397, 164], "fig_type": "figure"}]]], "page_282": [["block_0", [{"image_0": "282_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "282_1.png", "coords": [31, 426, 191, 465], "fig_type": "molecule"}]], ["block_2", ["Returning to Figure 7.11c, we see that there are two in\ufb02ection points, marked a and b, on each side\nof the bump. Between these two compositions, the free energy is concave down, and we say the\nsolution for that (x1,7)-is unstable. What does this mean? For any small local \ufb02uctuation in\nconcentration 8x1, the straight line connecting x1 8x1 and x1 + 8x1 will fall below AGm(x1). These\n\ufb02uctuations will therefore grow in amplitude and spatial extent; the mixture will spontaneously\nphase separate into two phases with compositions x\u2019l and x\u2019l\u2019 . Thus in a region where the AGm curve\nis concave down, the solution is unstable with respect to any \ufb02uctuation in concentration. The\nmechanism by which this phase separation occurs is called Spinoa\u2019al decomposition, and it is quite\ninteresting in its own right. However, in this chapter we are concerned with thermodynamics, not\nkinetics, so we will not pursue this here.\nYou may have noticed that there are two regions on the curve in Figure 7.11c, between x\"1 and a,\nand between b and x\u2019l\u2019 , where the curve is locally concave up, indicating stability, yet we already\nknow that the equilibrium state in these intervals should be liquid\u2014liquid coexistence with\nconcentrations x\u2019l and x\u2019l\u2019. What does this mean? These regions fall between the binodal and\nspinodal, and are termed metastable. They are stable against small, spontaneous \ufb02uctuations, but\nnot globally stable against phase separation. Consequently, a system in the metastable region may\nremain there indefinitely; it requires nucleation of a region of the new phase before separation\n"]], ["block_3", ["The next issue to address is the location of the spinodal, or stability limit. We have already\nindicated the condition for stability, namely Equation 7.5.2. The stability limit, then, is found\nwhere the second derivative of AGm changes sign, which de\ufb01nes an in\ufb02ection point:\n"]], ["block_4", ["In other words, if we tangent to the x1: 0 intercept, we obtain b Au2(x\u20191), and if we\nfollow it to the x1 1 intercept, k + b Anibal).\nThe argument so far applies for any AGm curve. Now if we have a AGm curve with a bump as\nin Figure 7.11c, we can draw one straight line that is tangent to AGm at two particular points,\ncall them x1\u2019 and x1\u201d. From the argument above, the x1:0 intercept gives us both Att2(xf)\nand Auz(x{\u2019), so these two chemical potentials must be equal. By the same reasoning the other\nintercept gives Au\ufb01xf) Auloq\"), and therefore we have shown that x1\u2019 and x1\u201d defined by the\ncommon tangent are indeed the compositions of the two coexisting phases. (Warning: for regular\nsolution theory, where the AGm curve is symmetric, xf and x1\u201d coincide with the local minima in the\nAG,\u201d curve, but this is not generally true.) So, in summary, one can locate the coexistence\nconcentrations by geometrical construction on a plot of AGm versus composition, or one could\ndo it from the analytical expressions for the two chemical potentials. However, the latter is\nalgebraically a little tricky, particularly because of the natural logarithm terms (see, for example,\nEquation 7.4.14).\n"]], ["block_5", ["But this relation holds whatever x\"1 we choose, so we can match the intercepts and slopes to obtain\n"]], ["block_6", ["Where it is the slope and b is the x1 0 intercept. But we chose )1 AGm for x1 xi, so inserting\nEquation 7.5.4 into Equation 7.5.5 we find\n"]], ["block_7", ["7.5.3\nFinding the Spinodal\n"]], ["block_8", ["Phase Behavior of Polymer Solutions\n269\n"]], ["block_9", ["y M \n(7.55)\n"]], ["block_10", ["kx\u2019, + b Ange) + x\u2019l[A,u.1(x'1) Att2(x\u20191)]\n(7.5.6)\n"]], ["block_11", ["5\u2019 A\ufb02zlxl)\nk Au1(x\u20191) Ayala)\n"]], ["block_12", ["(62\u201413(33)\n= 0\non the spinodal\n(7.5.8)\n8x,-\nTa)\n"]], ["block_13", ["(7.5.7)\n"]]], "page_283": [["block_0", [{"image_0": "283_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "283_1.png", "coords": [26, 16, 188, 211], "fig_type": "figure"}]], ["block_2", ["We can understand this by realizing that as T approaches TC from below, one in\ufb02ection point\nmoves to\nthe\nright,\nand\none\nto\nthe\nleft.\nThe\nrate of change of the\nin\ufb02ection\npoint,\n3/3xg(82AGm/3x,2), vanishes when the two meet.\nAlgebraic expressions for the spinodal and the critical point of regular solution theory can be\ndirectly obtained as follows. The chemical potential for component 1 (and of course, by symmetry,\nwe could equally well use component 2) comes from differentiating AGm:\n"]], ["block_3", ["proceeds. Nucleation, and the ensuing process of domain growth, is another interesting kinetic\nprocess that we will not discuss here. However, metastability can be a wonderful thing; diamond is\nmetastable with respect to graphite, the equilibrium phase of carbon at room T and p, but no one\nworries about diamonds transforming to graphite in their lifetime. A mechanical analogy is helpful\nin distinguishing among stable, metastable, and unstable systems, as shown in Figure 7.12. The ball\nin panel (a) may rattle around near the bottom of the bowl, but it will never come out; the system is\nstable. The ball in panel (b) is precariously perched on top of the inverted bowl, and the slightest\nbreeze or vibration will knock it off; the system is unstable. The ball in panel (c) can rattle around\nin the small depression, and may appear to be stable for long periods of time, but with a sufficiently\nlarge impulse it will roll over the banier and downhill to a lower energy state; the system is\nmetastable. Only state (a) is an equilibrium state, but state (c) might not change in our lifetime.\n"]], ["block_4", ["The \ufb01nal feature to locate in the phase diagram is the critical point. We know it lies on the\nspinodal, so it must satisfy Equation 7.5.8. But, we need another condition to make it a single,\nspecial point. The easiest way to visualize this is to return to Figure 7.10c and the plots of AGm at\ndifferent temperatures. Phase separation occurs only when we have the bump in AGm, so the\ncritical point marks the temperature where the bump first appears. This also corresponds to the\ntemperature where the two in\ufb02ection points merge into one and this is determined by\n"]], ["block_5", [{"image_2": "283_2.png", "coords": [34, 559, 215, 594], "fig_type": "molecule"}]], ["block_6", ["7.5.4\nFinding the Critical Point\n"]], ["block_7", ["Stable\nO\n"]], ["block_8", ["270\nThermodynamics of Polymer Solutions\n"]], ["block_9", ["Unstable\n"]], ["block_10", ["Metastable\n"]], ["block_11", ["Figure 7.12\nSchematic illustration of the difference between stable, unstable, and metastable states.\n"]], ["block_12", ["63%\u201c,\n"]], ["block_13", ["3x? \n= O\nat the critical point\n(7-5.9)\n"]], ["block_14", ["\u201911P\n"]], ["block_15", ["O\n"]], ["block_16", ["O\n"]]], "page_284": [["block_0", [{"image_0": "284_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "284_1.png", "coords": [24, 123, 182, 169], "fig_type": "molecule"}]], ["block_2", [{"image_2": "284_2.png", "coords": [29, 248, 199, 277], "fig_type": "molecule"}]], ["block_3", ["Now we can repeat this entire procedure for the Flory\u2014Huggins theory. The main difference will be\nthat the value of N breaks the symmetry of the AS\u201c, expression and will produce an asymmetric\n"]], ["block_4", ["The stability limit can now be obtained by taking the derivative with respect to x1:\n"]], ["block_5", ["Equation 7.5.14a and Equation 7.5.15 constitute two simultaneous equations that can be solved to\nobtain the critical point (see Problem 9): The result is x1 1,, :1/2 (which we could have guessed from\nthe outset, due to symmetry) and Xe :2. This means that unless it costs at least 2kT to exchange one\nmolecule of type 1 with one molecule of type 2, there will be no phase separation. To obtain the\ncritical temperature for a particular system, Tc, we need to know the value of X (i.e., zAw):\n"]], ["block_6", ["Generically, however, we can see that the larger Aw, the larger TC will be, and therefore the larger\nthe two-phase window. If, perhaps due to some specific interactions, Aw happens to be negative,\nthere will be no critical point according to regular solution theory; the system will be completely\nmiscible at all temperatures and in all proportions.\n"]], ["block_7", ["The critical point requires that we differentiate Equation 7.5.13 once more:\n"]], ["block_8", ["Note that this relation can be rewritten in the appealingly symmetric form\n"]], ["block_9", ["Now\n"]], ["block_10", ["where the subscript s denotes the value of X on the spinodal. (You should convince yourself that if\nwe followed the prescription for the stability limit given by Equation 7.5.8, and took the second\nderivative of AGm/kT from Equation 7.2.1 lb with respect to x1 instead of \ufb01rst obtaining [L1, we\nwould get the same relation.) This equation is a quadratic in x1:\n"]], ["block_11", ["7.5.5\nPhase Diagram from Flory\u2014Huggins Theory\n"]], ["block_12", ["so\ni :lnx1+(1\u2014x1)\u2014(1*X1)+X(1*xl)2\nRT\n=lnx1+ )((1-351)2\n(7'5'12)\n"]], ["block_13", ["Phase Behavior Polymer Solutions\n271\n"]], ["block_14", [{"image_3": "284_3.png", "coords": [42, 331, 128, 366], "fig_type": "molecule"}]], ["block_15", ["1\nx? \u2014x1+\u2014=0\n(7.5.l4a)\n2X5\n"]], ["block_16", ["X1\nIQ\n"]], ["block_17", ["6):\nn + n\nH n\nx x\n2\n4:\n1\n2\n21:\n12:12\n(7.5.11)\n3111\n(HI + n2)\n)7]\n722\n"]], ["block_18", ["3\nM1\n1\n3x1 (RT)\nx1\nXS(\n161)\n(7 5 13)\n"]], ["block_19", ["1\na (\u2014\u20142X(1 \u2014x1)) = \u2014\u00a7,+2Xc:0\n(75.15)\n6.17]\n1C]\n1\n"]], ["block_20", [{"image_4": "284_4.png", "coords": [47, 437, 227, 464], "fig_type": "molecule"}]], ["block_21", ["3\n1% :a\u2014nl.{mlnx1+n21n(1\u2014x1)+\u201d1(1_xl)X}\n"]], ["block_22", ["1\n1\n"]], ["block_23", ["__ _ \n_\n_ ___.\n7.5.16\n1%\n2k\n(\n)\nC\n"]], ["block_24", ["121 6x1\n\u201d2\n(9)61\n6x1\nl\u201c(1+)\u201d 3m\n1\u2014x1 3n1+(\nxl)X\nn1 (.3a\n(7'5'10)\n"]]], "page_285": [["block_0", [{"image_0": "285_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "285_1.png", "coords": [17, 199, 217, 269], "fig_type": "figure"}]], ["block_2", [{"image_2": "285_2.png", "coords": [31, 209, 209, 245], "fig_type": "molecule"}]], ["block_3", [{"image_3": "285_3.png", "coords": [34, 500, 300, 550], "fig_type": "molecule"}]], ["block_4", ["phase diagram. We already found the expression for AMI/RT, in Equation 7.4.14, so we can start\nfrom there. We should take derivatives with respect to x1, but in fact we can get away with the\nmuch easier task of differentiating with respect to ($2. This is because 3/811); (8x2/6\u00a22)6/3x2 and\nas we will be setting the expressions equal to zero, 8x2/8gb2 will divide out. Also, 8/6qb2=\n"]], ["block_5", ["Thus the critical polymer concentration depends inversely on x/N. For larger and larger N, qbc will\nbecome lower and lower, approaching 0 as N\u2014roo. Now to complete the analysis, we return to\nEquation 7.5.19 and find Xci\n"]], ["block_6", ["So, as N increases, Xe approaches 1/2. Recall from regular solution theory (which we can obtain\nfrom Flory\u2014Huggins theory by setting N 1) that Xe 2 and gbc 1/2. Thus, as we increase N, the\ncritical concentration moves more and more toward solutions dilute in polymer, and Xe becomes\nsmaller. For a given zAw, therefore, TC increases as N increases, meaning that polymers become\nless likely to form a homogeneous solution at a given T as N increases. This is illustrated in Figure\n7.13 for various values of N.\nRecall an important result from the previous section: At the theta temperature B 0 and X = 1/2.\nNow we have a third definition of the theta temperature; it is the critical temperature for a given\npolymer\u2014solvent system in the limit of infinite molecular weight. A polymer will be completely\nmiscible with a solvent above the theta temperature, but anywhere below the theta temperature, there\n"]], ["block_7", ["The critical point comes from\n"]], ["block_8", ["272\nThermodynamics of Polymer Solutions\n"]], ["block_9", ["or\n"]], ["block_10", ["which is again a quadratic:\n"]], ["block_11", ["We can now substitute this relation into Equation 7.5.17\n"]], ["block_12", [{"image_4": "285_4.png", "coords": [36, 361, 186, 403], "fig_type": "molecule"}]], ["block_13", ["\u2014\u2014 8/8q51, so we can work with 1.1.1. Thus the spinodal curve can be found from\n"]], ["block_14", ["1\n1\n_1(1+x/N)2_1(r+_2_+1)\nN\nx/N\nWm (m2 \n"]], ["block_15", ["63\nA1121\n_ \n_\n_i\n2\nEb; (F)\n_\n6962\n{1171(1\n$2) + 9152 (1\nN) +X\u00a22}\n"]], ["block_16", ["1\nl\nXe \n(7.5.19)\n2 (1 962,92\n"]], ["block_17", ["345%\nRT\n(1 952,192\nX\n(\n)\n"]], ["block_18", ["a \nm1\n1\u00b0\n2(1\u2014,i,,)\nN\u2014l\n(x/N\u20141)(\\/N+1)\nl\n1\n2\ng _\n7.5.22\n1 + m m\n\u2018\n)\n"]], ["block_19", ["_1\n1\n(i520\n+ 1--) \n(7.5.20)\n1 \u20ac523\n(\nN\n(1 <52,c)2\n"]], ["block_20", ["82\n(Am)\n~l\n"]], ["block_21", [{"image_5": "285_5.png", "coords": [52, 312, 217, 347], "fig_type": "molecule"}]], ["block_22", ["1\n2d)\n1\n2\n2c\n"]], ["block_23", [{"image_6": "285_6.png", "coords": [62, 117, 329, 192], "fig_type": "molecule"}]], ["block_24", [{"image_7": "285_7.png", "coords": [98, 152, 311, 185], "fig_type": "molecule"}]], ["block_25", ["\u2014\u2014l\n1\nl\n=.___+1_s_+2\n=\u2014'+\u2014\"\u20182Xs=0\n(7.5.17)\n1\u2014 an\nN\nm\ncm M\n"]], ["block_26", [{"image_8": "285_8.png", "coords": [105, 389, 332, 443], "fig_type": "molecule"}]], ["block_27", ["(7.5.23)\n"]]], "page_286": [["block_0", [{"image_0": "286_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "286_1.png", "coords": [34, 20, 315, 344], "fig_type": "figure"}]], ["block_2", ["is a danger of phase separation. Furthermore, as a polydisperse solution is cooled below the theta\npoint, the higher molecular weight chains will tend to phase separate first, a feature which can be\nused to advantage in fractionation.\nExamples of the phase behavior of polymer solutions are presented in Figure 7.14. The classic\nresults of Shultz and Flory for polystyrene in cyclohexane and polyisobutylene in diisobutyl ketone\nare reproduced in Figure 7.14a and Figure 7.14b, respectively. The data are experimental estimates\nof the coexistence concentrations (binodal), with smooth curves drawn to guide the eye. The\ndashed lines correspond to the predictions of the Flory\u2014Huggins theory. It is clear from these\nfigures that the theory indeed captures the main features of the data, namely a critical concentration\nthat is small and decreases with increasing M, and a critical temperature that increases with\nincreasing M. Neither the shape of the binodal nor the exact concentration dependence of the\ncritical composition is correct, however. The critical temperatures for these two systems are plotted\nas a function of M in Figure 7140, in a format suggested by Equation 7.5.23. The plots are linear,\nand permit reliable determination of the theta temperatures, with values that are in good agreement\nwith those determined by locating B(T) 0 (see Figure 7.8).\nA third system, polystyrene in acetone, is illustrated in Figure 7.15. Here we see phenomena that are\nnot described by the theory at all, namely phase separation upon heating for certain values of M. For\nexample, forM = 10,300 there appear to be two critical temperatures, one just below 0\u00b0C and the other\njust above 140\u00b0C. The solution would be two-phase for temperatures below the former and above the\nlatter and one-phase at intermediate temperatures. For M 19,800, there is no temperature at which a\nsolution with 0.1 < qbz < 0.15 would be one-phase. What are we to make of this? First, there is nothing\nin thermodynamics that forbids this kind of behavior. The Flory\u2014Huggins theory, however, cannot\n"]], ["block_3", ["Figure 7.13\nPhase diagrams for a polymer-solvent system with a theta temperature of 400 K, and N\nvalues of 102, 3 X 102 and 103. Smooth curves are binodals, dashed curves are spinodals. (Reproduced from\nMunk, P. and Aminabhavi, TM. in Introduction to Macromolecular Science, 2nd ed., Wiley, New York, 2002.\nWith permission.)\n"]], ["block_4", ["Phase \n273\n"]], ["block_5", ["T\n(K)\n"]], ["block_6", ["400 \nT: 6\n-\n"]], ["block_7", ["380 \n_\n"]], ["block_8", ["360\n"]], ["block_9", ["340\n"]], ["block_10", ["320\n"]], ["block_11", ["300\n|\nl\nI\n"]], ["block_12", ["r----\u2014---\n"]], ["block_13", ["\u201cq-\n"]], ["block_14", ["-\n"]]], "page_287": [["block_0", [{"image_0": "287_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "287_1.png", "coords": [2, 40, 416, 302], "fig_type": "figure"}]], ["block_2", ["Figure 7.14\nExperimental coexistence curves compared to the Flory\u2014Huggins theory (dashed curves) for\n(a) polystyrene in cyclohexane and (b) polyisobutylene in diisobutylketone. (c) Resulting critical temperatures\nversus degree of polymerization, plotted as suggested by Equation 7.5.23. (From Shultz, AR. and Flory, P.J.,\nJ. Am. Chem. Soc, 74, 4760, 1952. With permission.)\n"]], ["block_3", ["describe it (at least with )(> 0, see the next section) and thus this system illustrates some of the\nqualitative limitations of the theory. The critical point on a phase boundary that separates a two~phase\nregion at low temperature from a one-phase region at high temperature is called an upper critical\nsolution temperature (UCST), whereas a critical point on a phase boundary that separates a two\u2014phase\nregion at high temperature from a one-phase region at low temperature is called a lower critical\nsolution temperature (LCST). Thus for polystyrene in acetone with M:10,300, both a UCST and an\n"]], ["block_4", ["274\nThermodynamics of Polymer Solutions\n"]], ["block_5", ["250,000\na\n\\.\n:9\u201c\nEL 201\n_\ni:\nl-\no\n.\n-.\n_\n.\n\\,\n' -\n89,000\n"]], ["block_6", [{"image_2": "287_2.png", "coords": [40, 49, 255, 281], "fig_type": "figure"}]], ["block_7", ["30\n_\n1,270,000\n\u2014\n"]], ["block_8", ["25 :\n\\\n- .\n\u201c\n"]], ["block_9", ["(a)\n\u20ac92\n"]], ["block_10", ["35\n1\u2014\n1\u2014\nI_\n1\u2014\n"]], ["block_11", ["10.\n\u2018\n"]], ["block_12", ["15\nl:\n\\.\n\u2014\n"]], ["block_13", ["5\n_l\nl\ni\u2014\nI_\n0\n0.1\n0 2\n0.3\n0 4\n0 5\n"]], ["block_14", [".-\n\\\n'\n43,600\n"]], ["block_15", [{"image_3": "287_3.png", "coords": [125, 304, 348, 520], "fig_type": "figure"}]], ["block_16", ["g 3.3 \n_\n"]], ["block_17", ["+50\n3\n8\n3.2 \n_\n"]], ["block_18", ["3-5\nI\nI\n\u2014I\u2019\n\u2014r\n./\n3.4 \n_\n"]], ["block_19", ["O\n3.1 /\n_\n"]], ["block_20", ["30\nl\nl\nl\nl\n0\n0.02\n0.04\n0.06\n0.08\n0.10\n"]], ["block_21", ["(0)\n1__\n1\nW2N\n"]], ["block_22", ["C\n"]], ["block_23", [{"image_4": "287_4.png", "coords": [228, 29, 411, 288], "fig_type": "figure"}]], ["block_24", [{"image_5": "287_5.png", "coords": [235, 156, 439, 286], "fig_type": "figure"}]], ["block_25", [{"image_6": "287_6.png", "coords": [241, 50, 374, 216], "fig_type": "figure"}]]], "page_288": [["block_0", [{"image_0": "288_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "288_1.png", "coords": [25, 102, 281, 274], "fig_type": "figure"}]], ["block_2", ["What's in x?\n275\n"]], ["block_3", [{"image_2": "288_2.png", "coords": [33, 24, 279, 367], "fig_type": "figure"}]], ["block_4", ["LCST are observed; the Flory\u2014Huggins theory is only capable of predicting UCST behavior. (Note a\npossible source of confusion: the lower critical temperature lies above the upper critical temperature.)\n"]], ["block_5", ["As we have seen, the phase behavior of a polymer solution is determined largely by the interaction\nparameter X and by N. In this section we inspect some of the various ways to look at X- This will\ngive us some insight into how intermolecular interactions determine phase behavior. Furthermore,\nit turns out that X is used in the research literature in a range of different specific ways and it is\nimportant to see how that comes about.\n"]], ["block_6", ["We begin by returning to regular solution theory, and the initial definition of X in terms of zAw in\nEquation 7.2.9. If we know something about the intermolecular interactions, we ought to be able to\nsay something more specific about wij rather than just leaving it as a parameter. For London\n(dispersion) interactions, which are caused by spontaneous dipolar \ufb02uctuations on one molecule\ninducing a dipole on another, the interaction energy is\n"]], ["block_7", ["Figure 7.15\nExperimental phase diagrams for polystyrene in acetone. (From Siow, K.S., Delmas, G., and\nPatterson, D., Macromolecules, 5, 29, 1972. With permission.)\n"]], ["block_8", ["7.6.1\nX from Regular Solution Theory\n"]], ["block_9", ["7.6\nWhat's in X?\n"]], ["block_10", ["a\n\u2018:\n.\na,\n70\n:\n:\nw\n1-.\nt\n:'\n19300\n.\n"]], ["block_11", ["F50\nh\n/\u2014-\\.\\\n"]], ["block_12", ["6.0\n(7.6.1)\nWI.\n.\nN\n\u2014-\nJ\nI\u201c\n"]], ["block_13", ["110 \n-\n"]], ["block_14", ["190[\n4,800\n1L\n\ufb01r\n'\no\n140 t\n10300\nL\n"]], ["block_15", ["30\nl.\n..\n"]], ["block_16", ["50 \n"]], ["block_17", ["901 \\x\n\u2014\n"]], ["block_18", ["A?\n10300\n5%\n0 .\n"]], ["block_19", ["Jk\n4\n800\n,%\n"]], ["block_20", ["/\n"]], ["block_21", ["Olga);\n"]], ["block_22", ["0.05\n0.10\n0.15\n0.20\n0.25\n0.30\n"]], ["block_23", ["(02\n"]], ["block_24", ["'\n"]], ["block_25", ["-\n"]], ["block_26", ["-\n"]]], "page_289": [["block_0", [{"image_0": "289_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "289_1.png", "coords": [28, 196, 287, 278], "fig_type": "molecule"}]], ["block_2", [{"image_2": "289_2.png", "coords": [29, 143, 225, 169], "fig_type": "molecule"}]], ["block_3", [{"image_3": "289_3.png", "coords": [30, 196, 290, 282], "fig_type": "figure"}]], ["block_4", [{"image_4": "289_4.png", "coords": [31, 419, 194, 454], "fig_type": "molecule"}]], ["block_5", ["It turns out that the solubility parameter approach to X does not work well as a predictive method\nfor polymer solutions. As a first example, consider polystyrene in cyclohexane, which is a theta\nsolvent at 345\u00b0C (see Figure 7.7 and Figure 7.8). We can calculate X by Equation 7.6.6, using\n"]], ["block_6", ["This is a very simple expression for X that is directly related to fundamental physical quantities (\ufb01vap,\nCED, or 5). Note that1n the Flory\u2014Huggins theory we have assumed no volume change on mixing, so\nV1217] and the lattice size is defined by the solvent size. By examining a table of solubility\nparameters (see examples in Table 7.1), an estimate of X can be quickly obtained. To obtain a\ngood solvent for a polymer, one could begin by seeking solvents with similar solubility parameters. ..\nFor small molecules the heat of vaporization can be measured precisely, but for polymers this is not\nthe case; thus solubility parameter values are obtained indirectly, and can be quite uncertain.\n"]], ["block_7", ["Here we have introduced the solubility parameter, 5, defined as \\/ CED. The CED is usually given in\nunits ofcal/cm3, and thus 5 has the unusual units of (cal/cm?) \u201d2. Now we can rewrite Equation 7.6.3 as\n"]], ["block_8", ["7.6.2\nX from Experiment\n"]], ["block_9", ["This particular \u201cmixing rule\u201d for w,\u2014,- is called the Berthelot rule, and is certainly a plausible starting\npoint. Under this assumption, we have\n"]], ["block_10", ["Now we can see that because Aw can be written as a perfect square, it must be greater than or equal\nto zero, which is why X2 0 in regular solution\ntheory and the Flory\u2014Huggins theory.\nA\nWe can go further with this approach. The molar heat of vaporization for a pure substance, Uvap,\nshould be directly related to w, i.e.,\n"]], ["block_11", ["meaning that zw,,/2 is the interaction energy lost by removing one i molecule from the pure\nsubstance The cohesive energy density (CED)1s defined by dividing U,- m, by the molar volume,\n"]], ["block_12", ["where a,- is the polarizability of molecule i, and n,\u2014 is the distance between molecules i and j.\nBecause wg falls off with distance so rapidly, we are justified in considering only nearest-neighbor\ninteractions (as in regular solution or Flory\u2014Huggins theory). All atoms and molecules experience\nthis attractive (and fundamentally quantum mechanical) interaction, and for many molecules\nwithout strong dipole moments or hydrogen bonds, it is the only interaction that matters. For\nexample, simple alkanes and inert gases form condensed phases primarily because of London\nforces. Now if we have a lattice where we have imposed r,-,- = r},- = n, by design, then\n"]], ["block_13", ["V,\u2014, which1n regular solution theory13 just Avogadro\u2019 3 number times the lattice site volume:\n"]], ["block_14", ["276\nThermodynamics of Polymer Solutions\n"]], ["block_15", [{"image_5": "289_5.png", "coords": [39, 485, 121, 514], "fig_type": "molecule"}]], ["block_16", ["x= Lo] 52?\n(7.6.6)\nRT\n"]], ["block_17", ["CED\u2014UMP = s\u2014ii _=_ 5?\n(7.6.5)\nI\ni\n1'\n"]], ["block_18", ["W12 \u201c2% = \u201c\u20143371- %Z = \u20141/W11W22\n(7.6.2)\n\u201912\nr 11 r22\n"]], ["block_19", ["aiyap \navz\ufb02\n7..4\n2\n(6)\n"]], ["block_20", ["2A\n\u2014\u2014z\nw\n1w\n1\nw\u2014\n\u2014\u2014\u2014~\n\u2014\u2014\n12\n2\n11\nzwzz\n"]], ["block_21", ["= .3 (4m + [(m)2+(m)z])\n(7.6.3)\n"]]], "page_290": [["block_0", [{"image_0": "290_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "290_1.png", "coords": [26, 623, 167, 661], "fig_type": "molecule"}]], ["block_2", [{"image_2": "290_2.png", "coords": [27, 56, 420, 230], "fig_type": "figure"}]], ["block_3", ["What's in x?\n277\n"]], ["block_4", ["Source: EA. Grulke and Immergut, E.H. (Eds), 3rd ed., Wiley, New York, 1989.\n"]], ["block_5", ["V1 m M/p (84 g/mol)/(0.78 g/cm3) 108 cm3/mol and R 1.987 cal/K mol; the result is\nX m 0.14. But, we know that X =0.50 at the theta temperature, so we are off by almost a factor of\n4! To see that this is not an isolated case, consider the following example.\n"]], ["block_6", ["Table 7.1\nSolubility Values for Common Polymers and Solvents (Values for Polymers Are\nRepresentative Only)\n"]], ["block_7", ["What should we conclude from these calculations? First, the solubility parameter approach is\nnot quantitatively reliable for polymer solutions; the predicted value of X can be greater than, less\nthan, or very close to an experimental value. For at least one of the examples above, poly(vinyl\nacetate) in ethanol, the possibility of hydrogen bonding invalidates the basic assumptions of the\ntheory, so we should not be too surprised by this result. For poly(methyl methacrylate) in carbon\ntetrachloride, the predicted X is much too small. Based on comparisons of experiments with many\npolymers and solvents, the following empirical equation is found to be a much more reliable route\nto estimate X when the predicted value is less than about 0.3:\n"]], ["block_8", ["According to the theory, all three X values should be 0.5. In fact, one value is spot on, one is much\ntoo small, and one is much too big.\n"]], ["block_9", ["We will also need molecular weights and densities for the three solvents; they are approximately\n88 g/mol and 1.03 g/cm3 for dioxane; 154 g/mol and 1.59 g/cm3 for carbon tetrachloride; 46 g/mol\nand 0.79 g/cm3 for ethanol:\n"]], ["block_10", ["1 ,4-Polybutadiene\n8.3\nTetrahydrofuran\n9.1\nPolystyrene\n9.1\nChloroform\n9.3\nAtactic polypropylene\n9.2\nCarbon disulfide\n10.0\nPoly(methyl methacrylate)\n9.2\nDioxane\n10.0\nPoly(vinyl acetate)\n9.4\nEthanol\n12.7\nPoly(vinyl chloride)\n9.7\nMethanol\n14.5\nPoly(ethylene oxide)\n9.9\nWater\n23.4\n"]], ["block_11", ["Three other reported theta systems are polyisoprene in dioxane at 34\u00b0C, poly(methyl methacrylate)\nin carbon tetrachloride at 27\u00b0C, and poly(vinyl acetate) in ethanol at 19\u00b0C. Estimate X for each of\nthese cases, using Equation 7.6.6 and Table 7.1.\n"]], ["block_12", ["Polymer\n6 (cal/cm3)\u201d2\nSolvent\n6 (cal/cm3)\u201d2\n"]], ["block_13", ["Poly(tetrafluoroethylene)\n6.2\nn-Hexane\n7.3\nPoly(dimethylsiloxane)\n7.4\nCyclohexane\n8.2\nPolyisobutylene\n7.9\nCarbon tetrachloride\n8.6\nPolyethylene\n7.9\nToluene\n8.9\nPolyisoprene\n8.1\nEthyl acetate\n9.1\n"]], ["block_14", ["Solution\n"]], ["block_15", ["Example 7.3\n"]], ["block_16", ["Av\nx R\u2014T\u2018 (81 82)2\n(7.6.7)\n"]], ["block_17", ["(88/ 1.03)\n_\n(1.987x307)(1\n1)\n05\nPolyisoprene/dioxane: X m\n"]], ["block_18", ["Poly(methyl methacrylate)\u2014carbon tetrachloride: X m\n(8.6 9.2)22 0.06\n"]], ["block_19", ["(46/079)\nPolyvinyl acetate-693\u201c\u201c X mg?)\n"]], ["block_20", [{"image_3": "290_3.png", "coords": [238, 81, 416, 220], "fig_type": "figure"}]], ["block_21", ["(12.7 9.4)2= 1.1\n"]], ["block_22", ["(154/159)\n(1.987 x 300)\n"]]], "page_291": [["block_0", [{"image_0": "291_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Two questions should immediately come to mind. First, have we wasted our time examining the\nFlory\u2014Huggins theory in such detail, given that it fails to describe the thermodynamic properties of\npolymer solutions quantitatively, and, in many cases, even qualitatively? Second, can we do\nanything to rectify the situation? The answer to the \ufb01rst question is simple: no, we have not\nwasted our time. All of the developments that we went through (finding expressions for IT, #4: and\nmapping out the phase diagram, etc.) were model\u2014independent thermodynamics. The model really\nonly entered through the explicit expression for AGm that we used (i.e., Equation 7.3.13), and all\ndifferences between experiment and model predictions are directly attributable to inadequacies of\nEquation 7.3.13. The answer to the second question is not so clear-cut. There are three general\nstrategies employed in the polymer community. One is to try and improve the model, for example\nby identifying further interaction terms, additional sources of entropy, etc. This approach has been\npursued for many years. It has the virtue that it is possible to keep track of both the intended\nmeaning and quantitative effect of any added term in the expression for AGm. It has the drawback\nthat it is not yet generally successful if we try to restrict ourselves to only a small number of new\nterms. A second strategy is known as the \u201cequation of state\u201d approach. In this case, the AGm\nexpression is recast in a general form, with a finite number of parameters. In many cases the\nbehavior of mixtures can be well predicted based on knowledge of the parameters of the pure\ncomponents, so it can be a very practical strategy. One disadvantage is that it is difficult to gain\nmuch physical insight into the underlying molecular processes from the parameter values. The\nthird approach is the one most favored by experimentalists in polymer science, and so we will\nexamine this one a little more carefully. In essence it amounts to using X as a fittingfunction, not a\nnumber, so that X takes on whatever attributes are necessary to describe the data.\nIn general the free energy of mixing can be divided into two parts, an ideal part (superscript id)\nand an excess part (superscript ex):\nA0\u201c, AGE + A0: Mfg rasig + AH: ms:\n(7.6.8)\n"]], ["block_2", ["Furthermore, an ideal solution is defined as one for which\n"]], ["block_3", ["In the poly(methyl methacrylate)\u2014-carbon tetrachloride and polystyrene\u2014cyclohexane cases Equa-\ntion 7.6.7 does a much better job, albeit still not perfect. From Equation 7.6.7 there appears to be a\nnearly constant (and substantial) temperature-independent contribution to X in polymer\u2014solvent\nsystems, which is not anticipated by the regular solution theory approach. The fact that the 0.34\nterm does not have an explicit temperature dependence suggests that it re\ufb02ects an additional\nentropy of mixing contribution, rather than the purely enthalpic X anticipated by the model.\nThis is but one example of the limitations of this theory; another was provided by the\npolystyrene\u2014acetone phase diagrams in Figure 7.15, where the observed LCST behavior cannot\nbe explained by an interaction parameter that follows Equation 7.6.6. Still other problems are\napparent from the solubility parameter values in Table 7.1. For example, poly(ethylene oxide) is\nactually water-soluble, even though the difference between the two solubility parameters is huge.\nThis particular case involves hydrogen bonding and the rather unusual properties of water. In fact,\nthe assumptions of the theory are not consistent with any kind of strong or directional interaction,\nsuch as those involving permanent dipoles or hydrogen bonds. The packing of dipoles can\ninfluence the entropy of mixing through the various orientational degrees of freedom, which are\nnot included in the purely combinatorial entropy of the lattice model. Similarly, the energy of\ninteraction between two dipoles is very sensitive to the relative orientation, and the orientation of\none dipole will be sensitive to the positions and orientations of all neighbors. In short, there are\nmany ways in which the basic theory fails to incorporate important features of real systems,\nespecially when the interactions are strong and directional.\n"]], ["block_4", ["7.6.3\nFurther Approaches to X\n"]], ["block_5", ["278\nThermodynamics of Polymer Solutions\n"]], ["block_6", ["ASE \u2014\u2014k(x1 1a + x2 lnxg),\nM3 = 0\n(7.6.9)\n"]]], "page_292": [["block_0", [{"image_0": "292_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["What's in x?\n279\n"]], ["block_2", ["In other words, the entropy of mixing of an ideal solution is the purely combinatorial entropy of\nmixing, and there is no enthalpy of mixing (i.e., it is an athermal solution). Now we can view\nregular solution theory, as making specific \n"]], ["block_3", ["This Xeff will therefore include all the ingredients that pertain to the system under study, but it is an\nexperimental result not a model prediction. The scientific literature is often very confusing on this\npoint. For example, relatively few authors will actually mention Equation 7.6.11, even though that\nis what they are doing when they fit the data. Furthermore, quite a few will refer to this Xeff as the\n\u201cFlory\u2014Huggins interaction parameter,\u201d which it is not; the Flory\u2014Huggins X is correctly given by\nEquation 7.2.9.\nEmpirically, the Xeff function obtained by fitting data usually follows the form\n"]], ["block_4", ["where Xeff is sometimes resolved into two components, the \u201centhalpic part\u201d Xh and the \u201centropic\npart\u201d Xs- Interestingly, the parameters a and B may each be positive or negative. A negative at\nimplies some kind of specific attractive interaction between the components, such as might occur if\none component was a hydrogen bond donor and the other an acceptor. The sign of B is harder to\ninterpret, but presumably re\ufb02ects details of molecular packing. (Recall that in the lattice model we\nassume that each molecule has the same shape, and fits neatly into the lattice site, so there is no\nentropy associated with how the molecule is oriented within a site. For real molecules anisotropy\nof shape is almost inevitable, but it is not necessarily obvious whether there will be more or less\npacking possibilities per unit volume in the mixture compared to in the pure components, and\nhence the sign of the excess entropy can be hard to predict.) These various possibilities for the\nsigns of a and B give some insight into the various kinds of phase diagram that arise; for example,\nan LCST system would be the natural consequence of having oi < 0 and B > 0.\nOne final point about this experimental Xeff parameter: it very often exhibits a concentration\ndependence, which is equivalent to saying that the excess free energy of mixing is not entirely\nquadratic and symmetric with respect to components 1 and 2 (i.e., not AGE? 0-: (blag). This should\nnot be too surprising, given the great disparity in size and shape between a polymer and a solvent\nmolecule. This feature does have an important practical consequence, however. We can write the\nfree energy of mixing as\n"]], ["block_5", ["and if we actually measure AGm in an experiment, we can extract Xeff directly. However, it is not\neasy to measure AGm; more often we make a measurement that is really sensing a chemical\npotential, such as the osmotic pressure. In this case, if we take Equation 7.4.14 and substitute Xeff\nfor x, we would have\n"]], ["block_6", ["The approach that is adOpted is to de\ufb01ne an effective interaction parameter in of\nthe experimentally accessible free energy of mixing (i.e., the experimentally AGm\n"]], ["block_7", ["less the ideal part from Equation 7.6.9):\n"]], ["block_8", ["quantities, namely\nl\ufb02\u00a7:O,AH\u00a7:xmmU\"\n061m\n"]], ["block_9", [{"image_1": "292_1.png", "coords": [38, 615, 262, 666], "fig_type": "molecule"}]], ["block_10", ["AGE;i\nE \n(7.6.11)\nXeff\n\u00a21\u00a22kT\n"]], ["block_11", ["a\nn\ufb01=f+6=n+xs\nnan)\n"]], ["block_12", [{"image_2": "292_2.png", "coords": [49, 157, 116, 192], "fig_type": "molecule"}]], ["block_13", ["H_V1__\n1\nRT \n{1n(1\u20141\u2014\u2014)+cfqbg}\n(7.6.13b)\nN\n"]], ["block_14", ["A\ufb01i\n\u00a2\nk\u2014T\n= 9\u201851 111q + F2 111952 + Xeff\u00a7b1e1: is proportional to \u2014TS, where S is the entropy lost on stretching. If we\nassume that we begin with a Gaussian chain, we can write\n"]], ["block_7", ["Thus the scaling argument says\n"]], ["block_8", ["and combining Equation 7.7.3 and Equation 7.7.5 we have\n"]], ["block_9", ["where c is the number of segments per unit volume, and v(T) is the strength of the excluded\nvolume interaction between any two segments. (In particular, v B, and the c2 dependence is just a\nre\ufb02ection of the probability of two segments coming into contact. Basically Equation 7.7.2 is the\nsecond term of the virial expansion, Equation 7.4.7a; the first term does not matter because we can\nlet N become very large.) Now 0 N/R3 and when we integrate F05 over the volume of the coil (we\nassume FOS is the same everywhere, i.e., a mean-field approximation within the coil), we gain\nanother factor of R3:\n"]], ["block_10", ["random-walk statistics, and the larger the chain dimensions, the fewer the possible conformations.\nThus entropy resists unlimited coil expansion. The balance between the osmotic drive to swell and\nthe entropic drive to stay coiled up sets the ultimate dimensions. We estimate the \u201cfree energy\u201d per\nunit volume of this chain by a scaling argument, as follows [4]. (A scaling argument means that we\nleave out unimportant numerical factors, and emphasize the dependence of R on the main variable,\nin this case N. Also, we assume R will have the same dependence on N as does Rg.) The osmotic\npart is driven by the segment\u2014segment interactions, which we can write as V(T)(32 per unit volume:\n"]], ["block_11", ["Figure 7.16\nSchematic illustration of the swelling of a coil in a good solvent as a \u201csingle chain osmotic\npressure\u201d experiment.\n"]], ["block_12", ["Excluded Volume and Chains in a Good Solvent\n281\n"]], ["block_13", [{"image_6": "294_6.png", "coords": [36, 526, 221, 591], "fig_type": "molecule"}]], ["block_14", ["F\nN2\nR2\nkt; v(T) E + TV'\n(7.7.6)\n"]], ["block_15", ["Fe]\nR2\n\u2014\u2014 m \u2014_\n7.7.5\nkT\nN\n(\n)\n"]], ["block_16", ["F05\n\u2014 w v T 6\n7.7.2\nkT\n( )\n(\n)\n"]], ["block_17", ["Jeoil\nV(T) Fd(V01ume)\n: V(T) EER3 :\n1411)?\n(713)\n5.515 N\nH\"\n"]], ["block_18", ["17,, \u2014TS \u2014lnP(N,R)\n"]], ["block_19", ["N2\nN2\nN2\n"]], ["block_20", ["Swell\n_\u2014_b.\nC\n"]], ["block_21", ["-\u20143R2\n3\nit2\n"]], ["block_22", [{"image_7": "294_7.png", "coords": [144, 53, 302, 174], "fig_type": "figure"}]]], "page_295": [["block_0", [{"image_0": "295_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "295_1.png", "coords": [34, 348, 234, 399], "fig_type": "molecule"}]], ["block_2", ["The quantities I72 and V, are the molar volumes of the polymer and solvent, reSpectively (which\nare equal to the partial molar quantities in the Flory\u2014Huggins theory). The differences between\nEquation 7.7.9 and Equation 7.7.10 are the term in (13, which arises from an additional entropy\nchange for the chain due to the increased volume accessible to it, and the explicit expression for the\nprefactor. The factor of (1/2 x) in Equation 7.7.10 is proportional to B (recall Equation 7.4.18),\njust as was v(T) in Equation 7.7.9.\nNow we can brie\ufb02y examine the implications of Equation 7.7.10.\n"]], ["block_3", ["R34) NNW, and then raising each side to the fifth power:\n"]], ["block_4", ["To find the desired relation between R and N, we minimize Fto, with respect to R:\n"]], ["block_5", ["The preceding argument is not quite right, but it actually succeeds by a partial cancelation of\nerrors. The osmotic term re\ufb02ects a mean\u2014field argument-\u2014\u2014the probability of segment\u2014segment\ncontact is taken to be uniformly 02 across the coil\u2014which overestimates its importance. On the\nother hand, the elastic part presupposes a Gaussian chain, which is also not quite correct.\nNevertheless, the argument is reasonably simple and it captures the essence of the problem:\n"]], ["block_6", ["1.\nIn a theta solvent, X 1/2, and so a 1; there is no swelling. Thus in a theta solvent the chain\nbehaves as a random walk for all N large enough to be random walks.\n2.\nIn a good solvent, x<1/2, the chain swells (0: >1). In the limit of a very good solvent,\n"]], ["block_7", ["282\nThermodynamics of Polymer Solutions\n"]], ["block_8", ["and so\n"]], ["block_9", ["We can also recast this equation in terms of the expansion factor, a, by dividing each side by\n"]], ["block_10", ["a balance between excluded volume and coil distortion.\nFlory and Krigbaum worked out a much more detailed version of this calculation [5], with\nthe result\n"]], ["block_11", ["where the prefactor CM is given by\n"]], ["block_12", [{"image_2": "295_2.png", "coords": [38, 73, 194, 105], "fig_type": "molecule"}]], ["block_13", [{"image_3": "295_3.png", "coords": [43, 295, 172, 335], "fig_type": "molecule"}]], ["block_14", ["R v(T)1/5N3/5 B(T)1/5N3/5\n(7.7.8)\n"]], ["block_15", ["d\nF\nN2\n2R\nHE our)??? F \n(7.7.7)\n"]], ["block_16", ["a5 v(T)N1/2\n(7.7.9)\n"]], ["block_17", ["C\n=\n2 ,.\n0\n7.7.11\nM (7m) (M2N,.v.)( M i\n(\n)\n"]], ["block_18", ["a5 a3 2CM G @7117:\n(7.7.10)\n"]], ["block_19", [{"image_4": "295_4.png", "coords": [49, 542, 196, 582], "fig_type": "molecule"}]], ["block_20", ["This result, that v=3/5,\nis\nthe classic result for the excluded volume\n(self\u2014avoiding\nwalk) exponent that we cited before. The high N behavior in good and theta solvents is\nillustrated in Figure 7.17, for polystyrene in benzene (a good solvent), cyclohexane (theta),\nand trans-decalin (theta).\n"]], ["block_21", ["0:5 >> a3, and thus a5 N(Equation 7.7.9). Therefore,\n"]], ["block_22", ["N1/2\n0C \n"]], ["block_23", ["Nu oc \n(7.7.12)\n"]], ["block_24", ["R\n5\nN\u201d\n5\n5 =\ng\n1/2\n\u20191\n(Ran)\n0C (NI/2) \u201cN\n"]], ["block_25", ["NV\n"]]], "page_296": [["block_0", [{"image_0": "296_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "296_1.png", "coords": [24, 49, 343, 246], "fig_type": "figure"}]], ["block_2", ["Chapter Summary\n233\n"]], ["block_3", ["and tmns-decalin, both theta solvents. Data obtained by light scattering by several authors. (From Miyake, A.,\nEinaga, Y., and Fujita, H., Macromolecules, 11, 1180, 1978. With permission.)\n"]], ["block_4", ["In this chapter we have covered a great deal of material concerning the thermodynamic prOperties\nof polymer solutions. We have interwoven results that are strictly thermodynamic with those that\ndepend on a particular model, the Flory\u2014Huggins theory. The main points are as follows:\n"]], ["block_5", ["Figure 7.17\nExperimental radii of gyration for polystyrenes in benzene, a good solvent, and cyclohexane\n"]], ["block_6", ["7.8\nChapter Summary\n"]], ["block_7", ["1.\nUsing thermodynamics alone we were able to show how osmotic pressure measurements on\ndilute solutions can be used to determine Mn and the second virial coefficient, B. The latter\ngives direct information about the solvent quality: B > 0 corresponds to a good solvent, and\nB < 0 to a poor solvent.\nThermodynamic arguments were also sufficient to show how the complete phase diagram\n(T, (p) for a binary system could be constructed from an expression for the free energy of\nmixing. The key features of the phase diagram are the critical point, the coexistence curve\n(binodal), and the stability limit (spinodal).\n"]], ["block_8", ["103.\n"]], ["block_9", ["Cyclohexane\n102:\n"]], ["block_10", ["101\n1__I_1_1_|_I_I_LI\n1_\n11IIIIII\n1\n1\n|_II||1\n105\n106\nM\n107\n108\n"]], ["block_11", ["The crossover from theta-like behavior (a z 1) to fully developed excluded volume, a >> 1, is\nvery broad. It depends on both N and X (and therefore T). For example, for a given N the chain\nwill swell progressively as T is increased above T: G), and for a given T> G), a larger N\nchain will swell more than a shorter one. Because of this broad crossover, it is very common\nin experiments to find apparent values of the exponent v falling between 1/2 and 3/5. This\nwill become important particularly in the context of the intrinsic viscosity, as we shall see in\nChapter 9.\nAlthough designed for polymer solutions, Equation 7.7.10 actually hints at a very important\nresult for molten polymers. Suppose the solvent were a chain of the same monomer, but with a\ndifferent degree of polymerization, P. Presumably x~ 0, so the chain should swell. However,\n"]], ["block_12", ["V1 P, so if P exceeds \\/_\nN1n length there should be little or no swelling. In other words,\nchains1n their own melt should be Gaussian. Flory made this very important prediction1n the\n19503, but it was not until the advent of small\u2014angle neutron scattering in the early 19703 that\nthis fundamental result could be confirmed.\n"]], ["block_13", [". 'irans-Decalin\n"]]], "page_297": [["block_0", [{"image_0": "297_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where TC is the critical temperature for phase separation. For polystyrene with M \u2019\u201c=\u2018-' 3x106,\nShultz and Flory observed TC values of 68\u00b0C and 84\u00b0C, respectively, for cyclohexanone and\ncyclohexanol. Values of Vl for these solvents are about 108 and 106 cm3/m01, and 81 values\nare 9.9 and 11.4 (cal/cm3)1/2, respectively. Use each of these Tc values to form separate\nestimates of 82 for polystyrene and compare the calculated values with each other and with\nthe value of 82 from Table 7.1. Comment on the agreement or lack thereof for the calculated\nand accepted 6\u2019s in terms of the assumptions inherent in this method. If Equation 7.6.7 was\nused instead of Equation 7.6.6, does the agreement among the 82 values improve?\n3.\nThe term (1/2 X) that appears, for example, in the Flory\u2014Huggins expression for B (Equation\n7.4.18) is sometimes replaced by\n;_.=.(._%\n"]], ["block_2", ["W22 contribution continues to be valid. A slight error in counting is made\u2014to allow for\nsimplification of the resulting function\u2014*but this is a tolerable approximation in concentrated\nsolutions. In dilute solutions the approximation introduces more error, but the model is in\ntrouble in such solutions anyhow, so another approximation makes little difference.\n2.\nShow that\n"]], ["block_3", ["3.\nWe developed an expression for the free energy of mixing based on the Flory\u2014Huggins theory.\nThis is a mean\u2014field theory and reduces to the standard regular solution theory when the degree\nof polymerization of the polymer component is set equal to 1. The numerous assumptions of\nthe model were identified.\n4.\nThe Flory\u2014Huggins model was explored in detail, in terms of its predictions for B and for the\nphase diagram. Comparison with experiments reveals that in some systems the Flory\u2014Huggins\ntheory captures the phenomenology in a qualitative manner, but in others it does not. It does\nnot provide a quantitative description for any dilute polymer solution.\n5.\nThe concept of a theta solvent emerges as a central feature of polymer solutions. It has four\nequivalent Operational definitions: (a) the temperature where B 0, (b) the temperature where\nthe interaction parameter X 1/2, (c) the limit of the critical temperature, Tc, as M\u2014+00, and (d) a\nsolvent in which Rg NMW. Physically, a theta solvent is one in which the polymer\u2014solvent\ninteractions are rather unfavorable, so that the chain shrinks to its random~wa1k dimensions.\nThis contraction cancels the effect of the excluded volume interactions, which otherwise swell\nthe chain to self-avoiding conformations: Rg M35.\n6.\nThe Flory\u2014Huggins parameter X can be related to thermodynamic quantities such as the heat of\nvaporization and the cohesive energy density. Only for systems with very weak intermolecular\ninteractions can one h0pe to estimate X reliably based on tabulated thermodynamic quantities, and\neven then there needs to be a substantial, empirical correction term for polymer solutions. This\nmotivates the use of an effective interaction function that can be used to fit experimental data; the\nresulting Xeff has become a commonly used scheme to describe the excess free energy of mixing.\n"]], ["block_4", ["284\nThermodynamics of Polymer Solutions\n"]], ["block_5", ["Problems\n"]], ["block_6", ["1.\nIn the derivation of Equation 7.3.12 it is assumed that each polymer segment is surrounded by\n"]], ["block_7", [{"image_1": "297_1.png", "coords": [45, 637, 140, 672], "fig_type": "molecule"}]], ["block_8", ["To (2/12) 171 (61 62?\n"]], ["block_9", ["2 sites which are occupied at random by either solvent molecules or polymer segments.\nActually, this is true of only (2 2) of the sites in the coordination sphere\u2014(z 1) for chain\nends\u2014since two of the sites are occupied by polymer segments which are covalently bound to\nother polymer segments. Criticize or defend the following proposition concerning this effect:\nThe kinds of physical interactions that we identify as London or spontaneous dipole\u2014dipole\nattractions can also operate between segments which are covalently bonded together, so the\n"]]], "page_298": [["block_0", [{"image_0": "298_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018C. Bawn, R. Freeman, and A. Kamaliddin, Trans. Faraday Soc, 46, 862 (1950).\nIWR. Krigbaum and DD. Geymer, J. Am. Chem. Soc, 81, 1859 (1959).\n"]], ["block_2", ["The units of c are mg/mL, and II is in g/cmz. (This unit of g/cm2 is a bit old fashioned; to\nconvert H to dyn/cmz, multiply by the acceleration due to gravity, g 980 cm/sz.)\n5.\nWrite down by inspection the Flory\u2014Huggins theory prediction for the free energy of mixing\nof a ternary solution (polymer A, polymer B, and solvent).\n6.\nThe osmotic pressure of solutions of polystyrene in cyclohexane was measured at several\ndifferent temperatures, and the following results were obtained:I\n"]], ["block_3", ["4.\nThe osmotic pressure of polystyrene fractions in toluene and methyl ethyl ketone was\nmeasured by Bawn et al.\u2019r at 25\u00b0C, and the following results were obtained. Make plots of\n11/0 versus 0, and evaluate Mn for the three fractions in an appropriate way. Do the results\nmake sense? (Hint: If they do not, perhaps a more sophisticated analysis would help.) What\ncan you say about the quality of these solvents?\n"]], ["block_4", ["Problems\n285\n"]], ["block_5", [{"image_1": "298_1.png", "coords": [44, 153, 312, 377], "fig_type": "figure"}]], ["block_6", ["Fraction\n6 (g/cm3)\nl'I/RTcx10\u00b0 (mol/g)\n"]], ["block_7", ["II\n0.0976\n8.0\n0.182\n6.0\n0.259\n8.7\n"]], ["block_8", ["Fraction\n6 (g/cm3)\nII/RTC ><10\u20185 (mol/g)\n"]], ["block_9", ["II\n0.0081\n13.3\n0.0201\n14.2\n0.0964\n14.2\n0.180\n18.7\n0.257\n26.2\n"]], ["block_10", ["III\n1.75\n0.31\n1.41\n0.23\n2.85\n0.53\n2.90\n0.48\n4.35\n0.88\n6.24\n1.11\n6.50\n1.49\n8.57\n1.63\n8.85\n2.36\n"]], ["block_11", ["II\n1.55\n0.16\n3.93\n0.40\n2.56\n0.28\n8.08\n0.95\n2.93\n0.32\n10.13\n1.30\n3.80\n0.47\n5.38\n0.77\n7.80\n1.36\n8.68\n1.60\n"]], ["block_12", ["Fraction\nToluene c\nToluene II\nMEK c\nMEK II\n"]], ["block_13", ["I\n4.27\n0.22\n2.67\n0.04\n6.97\n0.58\n6.12\n0.14\n9.00\n1.00\n8.91\n0.31\n10.96\n1.53\n"]], ["block_14", ["where G) is the theta temperature (K) and d; is a new parameter. This substitution has the effect\nof replacing one parameter (x) with two ((1; and 6)). Why might one do this? How do 11/ and 6)\n\ufb01t into the discussion of effective X parameters in Section 7.6.3?\n"]], ["block_15", ["T 24\u00b0C\n"]], ["block_16", ["T 34\u00b0C\n"]]], "page_299": [["block_0", [{"image_0": "299_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["286\n"]], ["block_2", ["11.\n"]], ["block_3", ["10.\n"]], ["block_4", [{"image_1": "299_1.png", "coords": [39, 29, 240, 245], "fig_type": "figure"}]], ["block_5", [{"image_2": "299_2.png", "coords": [43, 437, 244, 492], "fig_type": "molecule"}]], ["block_6", [{"image_3": "299_3.png", "coords": [49, 370, 174, 411], "fig_type": "molecule"}]], ["block_7", [{"image_4": "299_4.png", "coords": [49, 137, 236, 217], "fig_type": "figure"}]], ["block_8", [{"image_5": "299_5.png", "coords": [52, 41, 232, 112], "fig_type": "figure"}]], ["block_9", ["Plot all of these data on a single graph as H/RTC versus 0, connecting the points so as to\npresent as coherent a display of the results as possible. Evaluate the molecular weights of the\ntwo polystyrene fractions. Criticize or defend the following proposition: These data show that\nthe (9 temperature for this system is about 34\u00b0C. As expected, the range of concentrations\nwhich are adequately described by the first two terms of the virial equation is less for sample\nIII than for sample II. Above this range other contributions to nonideality contribute positive\ndeviations from the two-term osmotic pressure equation. It would be interesting to see how\nthis last effect appears for sample III at 24\u00b0C, but this measurement was probably impossible\nto carry out owing to phase separation.\nBy combining Equation 7.5.23 and Problem 3, the following relationship can be obtained\nbetween the critical temperature T0 for phase separation and the degree of polymerization:\n1-1+; _1_+_1_\nT,_\u00ae\nor; f\n2N\n"]], ["block_10", ["Use the graphical method outlined above to evaluate (9,\n1,11, and X for polyisobutylene in\ndiisobutylketone.\nAssume X for\na polymer\u2014solvent system followed Equation 7.6.12.\nThere\nare four\npossible cases according to whether the parameters a: and B were positive or negative.\nWhat possible phase diagrams could be observed in each case, i.e., UCST only, LCST\nonly, both, neither, etc.\nFill in the steps omitted in the text to derive the relations for the critical mole fraction and the\ncritical X for regular solution theory (see Equation 7.5.15).\nFind the Flory\u2014Huggins expression for AM2(\u00a21), and then find the range of X for which there\ncan be two physically meaningful values of (is, for which Aug is the same. (Hint: Take the\nlarge N limit, but at the correct moment.) What is the signi\ufb01cance of this range of X?\nA polymer\u2014solvent mixture has a critical temperature of 300 K. The polymer is monodisperse\nand has a molar volume equal to three times that of the solvent. If the solvent were replaced\n"]], ["block_11", ["Derive this relationship and explain the graphical method it suggests for evaluating (9 and if}.\nThe critical temperatures for precipitation for the data shown in Figure 7. 14b are the following:\n"]], ["block_12", ["Tc (\u00b0C)\n18.2\n45.9\n56.2\n"]], ["block_13", ["M (g/mol)\n22,700\n285,000\n6,000,000\n"]], ["block_14", ["III\n0.0156\n2.46\n0.0482\n2.24\n0.0911\n3.42\n0.126\n4.96\n0.139\n6.05\n"]], ["block_15", ["Fraction\nc (g/cm3)\nH/RTC ><106 (mol/g)\n"]], ["block_16", ["II\n0.0959\n18.6\n0.178\n28.1\n0.255\n40.0\nIII\n0.0478\n5.50\n0. 125\n1 1.0\n0.138\n13.2\n"]], ["block_17", ["T 44\u00b0C\n"]], ["block_18", ["Thermodynamics of Polymer Solutions\n"]]], "page_300": [["block_0", [{"image_0": "300_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["20.\n"]], ["block_2", ["References\n"]], ["block_3", ["Problems\n237\n"]], ["block_4", ["12.\n"]], ["block_5", ["13.\n"]], ["block_6", ["14.\n"]], ["block_7", ["15.\n"]], ["block_8", ["16.\n"]], ["block_9", ["18.\n19.\n"]], ["block_10", ["17.\n"]], ["block_11", [".U\u2018P\u2018PJE\u2018JZ\u2018\n"]], ["block_12", ["Flory, P.J., J. Chem. Phys., 10, 51 (1942).\nHuggins, M.L., J. Am. Chem. 506., 64, 1712 (1942).\nFlory, P..T., Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\nde Gennes, P.G., Scaling Concepts in Polymer Physics, Cornell University Press, Ithaca, NY, 1979.\nKrigbaum, W.R. and Flory, P.J., J. Am. Chem. 506., 75, 1775 (1953).\n"]], ["block_13", ["by its dimer, what does Flory\u2014Huggins theory predict for the new critical temperature?\n(Hints: It is a good idea to \ufb01gure out which way TC should change before you do the details.\nAlso, remember that although X is dimensionless, it does depend on the chosen site volume.)\nSuppose you have just developed a synthesis method for a new, \ufb02exible polymer, capable of\nproducing high-molecular-weight samples. Provide a step-by-step outline of how you would\ngo about finding a theta solvent for this polymer, and then locating the theta temperature\nprecisely. Note that although there are many ways to proceed, some will be much more labor-\nintensive than others. Try to minimize the amount of effort required.\nThe Flory\u2014Huggins theory may be extended to a binary mixture of different polymers, i.e., an\nA/B blend. The resulting free energy of mixing can be written\n"]], ["block_14", ["qbc) in terms of NA and NB. Show that your results reduce to the solution case when NB \u2014> 1.\nCompare the solution and symmetric blend (i.e., NA=NB) results in the limit of in\ufb01nite\nmolecular weight; what is the crucial difference?\nThe results from Problem 13 imply that the critical temperature for a symmetric polymer\nblend should increase linearly with N. Confirm this result. What would the N dependence be\nfor the critical temperature if in fact X followed Equation 7.6.12?\nProve that for a polymer solution in the athermal limit (AI-1m 0), no phase separation can\noccur in the Flory\u2014Huggins model. (Hint: The sign of AGm is not enough.)\nConsider the Flory\u2014Huggins theory as applied to an AB statistical copolymer dissolved in\nsolution. It is possible to apply the same expression for the free energy of mixing as for a\nhomopolymer, if a new effective X parameter is employed. Develop a simple relation for Xeff\nin terms offA, XA, X3, and XAB, wherefA is the volume fraction of A units in the copolymer,\nand the three X\u2019s refer to polymer A\u2014solvent, polymer B\u2014solvent, and polymer A\u2014polymer B\ninteractions, respectively. Use mean-field approximations in the same spirit as in the original\ntheory. The important feature is that there are unfavorable A\u2014B interactions in the pure\nstatistical copolymer that are diminished in dilute solution. If your result is correct, you can\nuse it to show that, in principle, it should be possible to find a solvent that dissolves the\nstatistical copolymer but neither of the constituent homopolymers, at constant M.\nExtend the derivation for the excluded volume exponent (v: 0.6) given in Section 7.7 to a\nchain in a space of dimensionality d. What is v in two dimensions; why is it different from\n0.6? What does the result for four dimensions tell you?\nRepeat Problem 8 for a polymer\u2014polymer blend at high molecular weight.\nFor a polydisperse sample, what average degree of polymerization would apply in the\ncorresponding Flory\u2014Huggins expression for the spinodal? Prove your answer; one way to\ndo this is to find the spinodal for a binary mixture of two degrees of polymerization.\nRecast Equation 7.7.10 in terms of the second virial coefficient, B(T).\n"]], ["block_15", ["where NA and NB are the respective degrees of polymerization. Find the critical point (Xe and\n"]], ["block_16", ["AGm = \u201d5% lnq +\u00a3\u00a7 In 4313 +X\u00a2A\u00a2B}\n"]]], "page_301": [["block_0", [{"image_0": "301_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Flory, P.J., Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\nGraessley, W.W., Polymeric Liquids and Networks: Structure and Properties, Garland Science, New York,\n2003.\nKoningsfeld, R., Stockmayer, W.H., and Nies, E., Polymer Phase Diagrams, Oxford University Press, New\nYork, 2001.\nKurata, M., Thermodynamics of Polymer Solutions, Harwood, New York, 1982.\nRubinstein, M. and Colby, R.H., Polymer Physics, Oxford University Press, New York, 2003.\nTanford, 0, Physical Chemistry of Macromolecules, Wiley, New York, 1961.\nYamakawa, H., Modern Theory of Polymer Solutions, Harper & Row, New York, 1971.\n"]], ["block_2", ["288\nThermodynamics of Polymer Solutions\n"]], ["block_3", ["Further Readings\n"]]], "page_302": [["block_0", [{"image_0": "302_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this chapter we will explore the phenomenon of light scattering from dilute polymer solutions.\nLight scattering is an important experimental technique for polymers for several reasons. First, it\nprovides a direct, absolute measurement of the weight average molecular weight, MW. Second, it\ngives information about polymer\u2014polymer and polymer\u2014solvent interactions, through the second\nvirial coefficient, B (introduced in Chapter 7). Third, under many circumstances light scattering\ncan be used to determine the radius of gyration, Rg (described in Chapter 6) without any prior\nknowledge about the shape of the molecule (e.g., coil, rod, globule). Fourth, the description of\nthe scattering process that we will develop in this chapter may be readily adapted to x-ray and\nneutron scattering, two other techniques in common use in polymer science. In short, the wealth of\ninformation that may be obtained from light scattering more than justi\ufb01es the effort we will need to\nexpend in this chapter to understand how it works. As a \ufb01nal introductory comment, the fact that\nlight scattering can determine MW and B tells us that it is a thermodynamic measurement. The fact\nthat it can measure Rg tells us that it is a structural tool as well.\nA light beam may be described as a traveling electromagnetic wave. For our purposes (i.e.,\npolymers and solvents containing mostly C, H, O, and N atoms), the magnetic component of the\nwave is of no consequence, and so we can represent the wave as\n"]], ["block_2", ["where 50 is the amplitude of the electric field, a) is thefrequency in rad/s, t is time, Fis the position,\nand k is the wavevector. This is illustrated in Figure 8.1. Several comments about Equation 8.1.1\nare appropriate.\n"]], ["block_3", ["1.\nThe amplitude EC, is itself a vector. If we take the wave to be travelling along the x direction,\nEl, lies in the 32\u20142 plane. If E}, lies exclusively along one axis, say 2, then the beam is said to be\nz-polarized or vertically polarized; on the other hand, if 50 is uniformly distributed in the y\u2014z\nplane, then the wave is unpolarized.\n2.\nThe frequency can also be written\n"]], ["block_4", ["where v is the frequency in cycles per second (Hz), 0 is the speed of light in vacuum, and A0 is\nthe wavelength in vacuum.\n3.\nIn the photon picture of light, the energy carried by each photon E 2 hr\u00bb, where h is Planck\u2019s\nconstant (6.63 x 10'34 Is).\n"]], ["block_5", ["8.1\nIntroduction: Light Waves\n"]], ["block_6", ["Ezfocos(wt\u2014\u00a3\u00b0F)\n(8.1.1)\n"]], ["block_7", ["_ \n:2\n_\u2014\n(1)\n77V\nA0\n(8.1.2)\n"]], ["block_8", ["light Scattering by Polymer Solutions\n"]], ["block_9", ["8\n"]], ["block_10", ["289\n"]]], "page_303": [["block_0", [{"image_0": "303_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "303_1.png", "coords": [19, 53, 239, 142], "fig_type": "figure"}]], ["block_2", ["where 5* is the complex conjugate of E, obtained by replacing i ( : \\/\u2014~\u20141) with ~i. The use of the\ncomplex conjugate guarantees that I is a real number, because exp(ix) exp(\u2014\u2014ix) =1. Generally, we\nforget about saying the wave is the real part of the complex form, and just remember that at the end\nof a calculation we convert to intensities by Equation 8.1.5.\n"]], ["block_3", ["Figure 8.1\nSchematic of the electric field component of a z-polarized light wave propagating in the\nx direction.\n"]], ["block_4", ["For visible light )to R: 3500-7000 A, and consequently v c/AO as 1015 Hz (recall 0 3.0x 108 m/s).\nThis means that the oscillations in the electric field amplitude are much too rapid for detectors to\nfollow: photomultiplier tubes, photodiodes, and other photodetectors typically have time constants\non the order of nanoseconds. Instead, the detector actually integrates in time the incident intensity,\nI (energy/area/time), which is proportional to IE EI If12; such detectors are called square-law\ndetectors. We will not worry about the proportionality factors between intensity and electric\nfield squared, because in scattering experiments we will always use the ratio of the scattered\nintensity to the incident intensity, thereby canceling out these prefactors. (Even if we did not do\nthat, the detectors actually generate an electrical current proportional to the incident intensity, and\nit is rare that anyone goes through the trouble of converting this signal into the actual incident\nintensity. Rather, we can calibrate the electrical signal with a reference intensity, as will be\ndescribed in Section 8.7.)\nWe could continue to use the cosine representation of the light wave as in Equation 8.1.3, but it\nturns out the arithmetic is much more convenient if we use complex notation. If you do not recall\nhow complex numbers work, they are reviewed in the Appendix. The advantages stem primarily\nfrom the fact that complex numbers allow us to factor out the temporal and spatial dependences of\nthe wave amplitude, and that squaring the amplitude becomes much easier. Accordingly, the wave\ncan be written as\n"]], ["block_5", ["The essence of the traveling wave is that if we consider a particular instant in time (out \ufb01xed), then\nthe wave oscillates in space with wavelength A; and if we consider a particular point in space (kx\nfixed) the wave oscillates in time with frequency a). It will turn out that it is the spatial dependence,\nI; F or kx, that will play the crucial role in scattering experiments.\n"]], ["block_6", ["4.\nThe phase factor, I: F, is the projection of the wavevector k onto the particular position\nvector of interest. The wavevector points in the direction of propagation, x in this instance, and\n"]], ["block_7", ["where Rel. . .} means the real part of the complex argument. The intensity is proportional to\n"]], ["block_8", ["290\nLight Scattering by Polymer Solutions\n"]], ["block_9", ["_p\n_9\nL50, E\"0{eiwre\u2014iwre\u2014if.Feifof'}| \nI\no Eel\n(3,1,5)\n[Nig*\u00b0\u00a7l:\n"]], ["block_10", ["E Re{l\u00a7\u20180 exp [i(wt i: PM } Re{\u00a7o exp[iwt] exp [\u2014ii: F] }\n(8.1.4)\n"]], ["block_11", ["has amplitude 277M. Thus, our wave could be rewritten as\n"]], ["block_12", ["EOE, t) 2 Ed )2, z) c0s(wt kx) = E; cos (271' [pt 3)\n(8.1.3)\n"]]], "page_304": [["block_0", [{"image_0": "304_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "304_1.png", "coords": [33, 58, 392, 231], "fig_type": "figure"}]], ["block_2", ["Source: From York,1989.\n"]], ["block_3", ["Poly(tetra\ufb02uoroethylene)\n1 .4 1\nMethanol\n1 .326\nPoly(dimethylsiloxane)\n1 .43\nWater\n1 .333\nPoly(ethylene oxide)\n1.46\nEthanol\n1.359\nAtactic polypropylene\n1.47\nEthyl acetate\n1.370\nAmorphous polyethylene\n1.49\nrt-Hexane\n1.372\nPoly(vinyl \n1.49\nTetrahydrofuran\n1.404\nPoly(methyl \n1 .49\nCyclohexane\n1 .424\n"]], ["block_4", ["Table 8.1\nRefractive Indices of Common Polymers and Solvents\n"]], ["block_5", ["1 ,4-Polybutadiene\nl .52\nBromobenzene\n1 .557\nPolystyrene\n1.59\nCarbon disulfide\n1.628\n"]], ["block_6", ["Refractive indices for A, 5393 A, at 20\u00b0C or 25\u00b0C.\n"]], ["block_7", ["Equation 8.1.1 represents a solution to Maxwell\u2019s equations in a homogeneous medium. The\nfrequency of the wave (and also the energy of the equivalent photon, hp) is independent of the\nmedium, but the wavelength is not. The wavelength in the material, A, relative to the wavelength in\nvacuum, A0, is determined by a material property called the refractive index n:\n"]], ["block_8", ["1,4-Polyisoprene\n1 .50\nChloroform\n1 .444\nPolyisobutylene\nl .5 l\nToluene\nl .494\n"]], ["block_9", ["where v is the speed of light in the material. Thus the amplitude of the wavevector, k = Mil, is often\nwritten as 27m/A0. The refractive index is determined, in turn, by the polarizability of the\nconstituent molecules or and their spatial arrangement (density and orientation). Qualitatively,\nthe polarizability re\ufb02ects the ability of the incident electric field to distort the electronic distribu-\ntion within a molecule, and this distortion, in turn, reduces c to v. Consequently, more polarizable\nchemical moieties, such as aromatic rings, generally lead to higher refractive indices than less\npolarizable groups, such as \u2014CH3 or ~CH2~. In the liquid state, all orientations of the individual\nmolecules are equally probable, and we say the liquid is isotropic. (In this case we need not worry\nabout the fact that most molecules are actually anisotropic: the polarizability is different along\ndifferent molecular axes). The magnitude of n ranges from about 1.3 to 1.6 for common polymers\nand aqueous or organic solvents as can be seen from the representative values in Table 8.1. The\nrelevant equation that relates the material property, n, to the molecular property, oz, is a version of\nthe Lorentz\u2014Lorenz equation for an ideal gas (also known as the Clausius\u2014Mosotti equation):\n"]], ["block_10", ["Polymer\nn\nSolvent\nr1\n"]], ["block_11", ["where n is the refractive index of the solution, it, is the refractive index of the pure solvent, and 1/\u20181\u2019\nis the number of solute particles per unit volume.\n"]], ["block_12", ["Basic Concepts of Scattering\n291\n"]], ["block_13", ["Scattering is the reradiation of a traveling wave due to a change in the character of the medium in\nwhich the wave is propagating. The mathematical description of scattering could be largely\ndeveloped without specifying whether we are talking about light waves, sound waves, electrical\nsignals, or waves on a lake, as there is a great deal of commonality to all these phenomena. For\nlight, scattering will be caused by local changes in refractive index or polarizability, due to, for\n"]], ["block_14", ["8.2\nBasic Concepts of Scattering\n"]], ["block_15", [{"image_2": "304_2.png", "coords": [37, 508, 115, 545], "fig_type": "molecule"}]], ["block_16", ["n A\u20140 3\n(8.1.6)\n)1\nv\n"]], ["block_17", [{"image_3": "304_3.png", "coords": [46, 318, 96, 353], "fig_type": "molecule"}]], ["block_18", ["477\n\u2018If\n(8.1.7)\n(If-3\n"]], ["block_19", ["2\n2\ntt\u2014nS\n"]]], "page_305": [["block_0", [{"image_0": "305_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1 A apart, and we shine A 5000 A light on them (Figure 8.2a). What will the scattering look like?\nEach atom will radiate a scattered wave with the same amplitude and wavelength, but the net\nscattered intensity will be zero. Why? If we select any particular scatterer and particular detection\nangle, we can always find another scatterer that is exactly M2 further away from the detector. The\ntwo waves from these scatterers will cancel each other. We can do this because the array is\nperfectly regular, and because the array is essentially a continuum (1<<5000 A). The only\ndirection where this argument does not apply is forward, i.e., the scattering angle 6:00. Here\nthe phase shift of the scattered waves between two atoms is exactly canceled by the phase shift\nbetween the incident waves arriving at the two atoms. Thus the light beam propagates happily\nstraight through the material. The important lesson is this: there is no scattering from a perfectly\nhomogeneous material.\n"]], ["block_2", ["for the case of molecules in solution. We anticipate that bigger molecules will scatter propor\u2014\ntionally more than smaller ones. On the other hand, suppose we scatter light from objects that are\nconnected to one another, e.g., monomers within one polymer. Now the phases of the waves scattered\nfrom different monomers should be related, because the distance rjk between any two monomersj and\nIt has some preferred value or range of values. This will lead to some interference between\nthese waves and a net loss of scattered intensity. Recall that interference between two waves is\nconstructive only if the difference in distance that the two waves travel to the detector is some integral\nnumber of wavelengths, m; in other words, when krzmh. The interference will be completely\ndestructive if the path difference is an integral number plus one half of A: kr (m + 1/2)/\\. We can\nexpect, therefore, that in order to see significant destructive interference we need the distance\nbetween scatterers to be a significant fraction of A, say a few percent. This will turn out to be\nthe case. For visible light with A0 re 5000 A, only for polymers that are larger than at least 100 A do\nwe have to worry about this kind of interference. You might recall from Example 6.2 that this size\ntypically corresponds to molecular weights in excess of 105 g/mol.\n"]], ["block_3", ["8.2.2\nScattering from a Perfect Crystal\n"]], ["block_4", ["Imagine we have an absolutely perfect, regular array of scatterers (atoms, if you like) placed every\n"]], ["block_5", ["Imagine that we have scattering from randomly placed, noninteracting objects; we should get a\ntotal scattered intensity, [3, proportional to the number of scattering objects:\n"]], ["block_6", ["8.2.1\nScattering from Randomly Placed Objects\n"]], ["block_7", ["example, a dust particle in the air or to a polymer in the solvent. We will only consider\nnonabsorbing media here, so the incident light intensity will either be transmitted or scattered,\nbut not absorbed. In general, the scattered wave propagates in all directions, i.e., it is a spherical\nwave in three dimensions or a circular wave in two. If the incident and scattered frequencies are\nthe same, i.e., no energy is exchanged between the medium and the wave, the scattering process\nis classified as elastic. We will only consider elastic scattering in this chapter. (There is also\nquasielastic scattering, in which very small differences in energy are detected; this is the basis\nof the powerful technique of dynamic light scattering, which will be described brie\ufb02y in\nChapter 9. Raman and Brillouin scattering are examples of inelastic processes, where the\nlight exchanges vibrational or rotational quanta with the molecules in the former, and exchanges\nenergy with traveling density waves or phonons in the latter.) Lastly, scattering is termed\nincoherent if the intensity is independent of the scattering angle, and coherent if it depends\non the scattering angle. The origin of this terminology will become apparent in the following\ndiscussion, but generically coherence implies that there exists some particular phase relationship\namong different waves.\n"]], ["block_8", ["292\nLight Scattering by Polymer Solutions\n"]], ["block_9", ["15 (number of scatterers) >< (scattering power of each object)\n(8.2.1)\n"]]], "page_306": [["block_0", [{"image_0": "306_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["There are two ways by which we could generate scattering from our hypothetical array. One would\nbe to remove a few scatterers at random (Figure 8.2b). Then the \u201cpairing\u2014off\u201d argument fails\nbecause each atom we remove used to cancel a scattered wave from some other atom, but now it\ncannot. Thus we conclude that random \ufb02uctuations in an otherwise homogeneous medium give\nrise to scattering. This scattering is incoherent: because the \ufb02uctuations are random by construc-\ntion, on average there can be no phase relation between them.\nThe second way of obtaining scattering from our hypothetical array would be to make A close to\nthe distance between the scatterers. For our example, if we use x\u2014rays, where )t :1.54 A is a\ntypical value, then the pairing-off argument fails again. There will be particular angles in which\nplanes of atoms are separated by integral multiples of )t and the scattering from different atoms will\nbe in phase, i.e., coherent. This process is called Bragg dz\ufb01\u2018ractz\u2019on, as illustrated in Figure 8.3 and\ndescribed below. The point we want to emphasize now is that there will be coherent scattering\nwhenever there is a spatial correlation between scattering objects on a distance scale comparable to )t.\nBragg diffraction from atomic crystals gives sharp scattering peaks because the spatial coherence\nis very high; the crystal structure is regular over large distances (see Chapter 13). In scattering\nfrom polymers, we will find that the Spatial coherence is lower, but not vanishing; for example,\n"]], ["block_2", ["Figure 8.2\n(a) A light wave propagating through a perfectly regular array of scatterers will not be scattered\nif the distance between neighbors is much less than )t. (b) Random \ufb02uctuations in the perfect array lead to\nincoherent scattering.\n"]], ["block_3", ["8.2.3\nOrigins of Incoherent and Coherent Scattering\n"]], ["block_4", ["(a)\n+++++++++++\n+++++++++++\n+++++++++++\n+++o+++++++\n+++++++++++\n++++++++o++\n+++++++++++\n+++++++++++\n+o+++++++++\n"]], ["block_5", ["m\n+++++++++++\n"]], ["block_6", ["Basic Concepts \n293\n"]], ["block_7", ["+++++++++++\n+++++++++++\n+++++++++++\n+++++++\nt\nm\u2018lll lll'lIl \u201cill\u201c \u201cIll\" \"N\"\n"]], ["block_8", ["III.IHIIMII\nIn...\" ......\n++++++\n+++++++++++\n+++++++++++\n"]], ["block_9", ["l\nl\n|\n"]], ["block_10", ["I\n"]]], "page_307": [["block_0", [{"image_0": "307_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "307_1.png", "coords": [14, 48, 256, 217], "fig_type": "figure"}]], ["block_2", ["294\nLight Scattering by Polymer Solutions\n"]], ["block_3", ["Figure 8.3\nIllustration of Bragg\u2019s law. (a) The incident light wave and the diffracted light wave each makes\nan angle 6/2 with each plane in the crystal. The planes are separated by a distance D, and the direction of the\nscattering vector c? is indicated. (b) The bold line segments correspond to the extra distance the wave travels in\nbeing scattered from the second plane. All four indicated angles are equal to 6/2, and thus each of the bold line\nsegments has length D sin(6/2).\n"]], ["block_4", ["within a coil, the likelihood of two monomers being separated by a certain distance might be given\nby the Gaussian distribution (Equation 6.7.12). This lower degree of spatial coherence will lead to\nscattering that is a smooth function of the scattering angle, rather than sharp peaks, but the\nunderlying phenomenon is the same.\nThe main relationship we will derive in this chapter, known as the Zimm equation (Equation\n8.5.18), re\ufb02ects these two sources of scattering. A polymer solution would be a homogeneous\nmedium, except for two things. There are random \ufb02uctuations in concentration that give rise to\nincoherent scattering, with intensity proportional to the product of concentration and molecular\nweight (Equation 8.4.24b). Then, large polymers have monomers with partially correlated spatial\nseparations that are a significant fraction of A. This correlation leads to interference and coherent\nscattering, which can be used to determine Rg (Equation 8.5.18). The development will proceed in\nthree stages. First we will consider scattering from a single isolated atom or molecule, a result\ngenerally attributed to Lord Rayleigh. In the second stage we will apply this result to a dilute\nsolution of small polymers, to obtain an expression for the incoherent scattering. Finally, we will\nconsider larger molecules and the resulting coherent contribution. To conclude this overview\nsection, we develop Bragg\u2019s equation and define the scattering vector \u00e9\u2019.\n"]], ["block_5", ["Consider a wave incident on a series of equally spaced, parallel planes of scatterers (which could\nbe atoms, but need not be). The planes are separated by a distance D, and the angle between the\nincident wave direction and the scattering planes is 6/2, as shown in Figure 8.3a. What is the\nrelation among D, )t, and 6 for there to be scattering (or diffraction, as it is called in this context) in\nthe direction an angle 6 away from the incident direction? The distance the light travels from the\nsource to the detector is the same for all atoms in a given planej, so there is no problem there. What\nis essential is that the distance traveled by waves scattered from the next further planej + 1 be m\nlarger, where m is an integer. Then the waves scattered from one plane will be in phase with the\nwaves scattered from the next, and there will be constructive interference at the detector. By\nextension, waves from planej + 2 will travel 2m/\u2018t further than from plane j, and therefore will still\nbe in phase, and so will waves from all the planes in the array. This condition is the basis of\nBragg\u2019s law. To state this as an equation, consider the enlarged diagram in Figure 8.3b. By\n"]], ["block_6", ["8.2.4\nBragg's Law and the Scattering Vector\n"]], ["block_7", ["(a)\n(b)\n"]], ["block_8", ["mm\n"]], ["block_9", ["M.\u201c\n9/2\nj+1\n"]], ["block_10", ["j+2\n"]], ["block_11", [{"image_2": "307_2.png", "coords": [245, 44, 419, 197], "fig_type": "figure"}]]], "page_308": [["block_0", [{"image_0": "308_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "308_1.png", "coords": [32, 522, 203, 640], "fig_type": "figure"}]], ["block_2", ["constructing the indicated perpendiculars, it is possible to see that there are two extra segments that\nthe wave must travel when it is scattered off plane j + 1. The length of each of these segments is D\nsin(9/2). Thus Bragg\u2019s law can be written as\n"]], ["block_3", ["we can see that Bragg\u2019s law (Equation 8.2.2) is automatically satisfied. In other words, the\ncriterion for Bragg diffraction is that the scattering vector (determined by the apparatus)\ncoincides with a reciprocal lattice vector (determined by the material).\n3.\nThe magnitude q has dimensions of inverse length. As will become clearer in Section 8.5 and\nSection 8.6, the scattering will be sensitive to structure in the solution on the length scale l/q,\nso the choice of q (which is fixed by choice of )t and 6) ultimately determines what kind of\nstructural information a given scattering experiment can provide.\n"]], ["block_4", ["Figure 8.4\nThe length of the scattering vector is calculated in terms of the scattering angle 6 and the\nmagnitude of the incident and scattered wavevectors, 27r/A.\n"]], ["block_5", ["What is the signi\ufb01cance of (.7? There are three aspects worth bringing out now:\n"]], ["block_6", ["This is illustrated in Figure 8.4, where again 6 is the angle between the scattered and incident\nwaves.\nFor elastic\nscattering,\nboth\nincident\nand\nscattered\nwave\nvectors\nhave\nmagnitude\n27r/)t 27m/)t0, and thus the magnitude of c? can be obtained (see Figure 8.4):\n"]], ["block_7", ["I.\nIt is also known as the momentum transfer vector. Although no energy is exchanged in elastic\nscattering, photons carry momentum and there must be a change in momentum when the\npropagation direction changes. The law of conservation tells us that momentum is transferred\ninto the medium along the direction of \u00a7. In fact, at sufficiently high intensity the resulting\n\u201cradiation pressure\u201d can cause particles to move.\n2.\nThe direction of a\u2019, shown in Figure 8.3a, corresponds to the normal to the planes of scatterers.\nA vector pointing in this direction with amplitude 27r/D is called a reciprocal lattice vector\n(reciprocal because D is in the denominator), and when the magnitudes of these two vectors\ncoincide\n"]], ["block_8", ["A central quantity in any scattering process is the scattering vector, defined by the difference\n"]], ["block_9", ["between the incident and scattered wave vectors:\n"]], ["block_10", ["Basic Concepts of Scattering\n295\n"]], ["block_11", [{"image_2": "308_2.png", "coords": [44, 225, 229, 261], "fig_type": "molecule"}]], ["block_12", ["m)t 2D sin (3)\n(8.2.2)\n"]], ["block_13", ["528\u20148\nman\n"]], ["block_14", ["_,\n2\n,\n6\n4\n,\n6\n|q| E q = 2 (77:) sm (2) 7?.\u2014 srn (2)\n(8.2.4)\n"]], ["block_15", ["271'\n471'\n,\n6\n\u20185 T \n(825)\n"]], ["block_16", ["(2am) sin(9/2)\n"]]], "page_309": [["block_0", [{"image_0": "309_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "309_1.png", "coords": [34, 439, 196, 536], "fig_type": "molecule"}]], ["block_2", ["In the next two sections we will be concerned with incoherent scattering, and will not need to be\nconcerned with the role of q, but then in Section 8.5 and Section 8.6 it will be of central\nimportance.\n"]], ["block_3", ["The derivation of this equation comes from basic electromagnetism, but it involves some rather\nhairy vector calculus, so we omit it; it can be found in many introductory physics texts. However,\nwe can understand the important inverse-dependence on 1. The total energy of the scattered wave\nshould be\nconserved\nFrom Equation 83.2, we see that the intensity will be proportional to\n|E: 0E3 | r 2.The total energy will be proportional to the intensity integrated qverjhesurface\nof a sphere of radius r. As we go further away from the dipole and integrate [ES\n\u00b0 ESI over the\nsurface of the sphere, the area will1ncrease as r2,and thus the total energy will be independent of r.\nNote that (11'/6) has the dimensions of length, since the induced dipole is essentially the product\nof a charge and the distance of charge separation. Therefore we can write the magnitude of the\nacceleration as\n"]], ["block_4", ["The induced dipole moment oscillates at the same frequency w as the field. An oscillating dipole\ninvolves an accelerating charge, and will therefore radiate an oscillating electric field that we will\ncall the scattered wave ES. The magnitude of this wave depends on the direction of observation\n(through an angle ()6 de\ufb01ned below), the speed of light, the distance from the dipole (r), the electron\ncharge (6), and the acceleration of the charge in the dipole (51'):\n"]], ["block_5", ["The ratio of the scattered intensity, 1,, to the incident intensity, 10, is given by\n"]], ["block_6", ["8.3\nScattering by an Isolated Small Molecule\n"]], ["block_7", ["We begin with the fact that an incident light wave Ei will induce a dipole moment 11\u2019 in an atom or\nmolecule, where\n"]], ["block_8", ["296\nLight Scattering by Polymer Solutions\n"]], ["block_9", ["qb l, and there is no scattered wave in the vertical direction 2 (sin qb 0). If we use vertically\npolarized light and detect scattering in the horizontal plane (currently the most commonly\nemployed geometry), then sin qb 1 and life is easy. The scattered intensity, normalized to the\nincident vertically polarized intensity, is given by\n"]], ["block_10", ["where we have used to 277v: Zia-\u20ac010, and as before the dot product is required in taking the\nsquare of the electric field vector.\nIf the incident wave travels along x, and is polarized along 2, then d) is the angle of detection\nrelative to the z-axis (see Figure 8.5a). There is no angular dependence in the x \u2014y plane since sin\n"]], ["block_11", [{"image_2": "309_2.png", "coords": [36, 433, 220, 528], "fig_type": "figure"}]], ["block_12", ["[S___\nEg\u2018ogs\n_62|a|2\nSiab\n[0\nE\"? ii_\nm2\nlEi*l.Ei I\n"]], ["block_13", ["[12 a]; aE\u2019, exp[i(wt 15' F)]\n(8.3.1)\n"]], ["block_14", ["|\u00a7,| =:3}; sins!)\n(8.3.2)\n"]], ["block_15", ["_16ar4a\n_ \u201dT4\nsin (8\n(8.3.4)\n"]], ["block_16", ["_ 6,\n(8:21:04)\n"]], ["block_17", ["dt\n6\n"]], ["block_18", ["d2\n*\n"]], ["block_19", ["|E\u00a7\u201c\u00b0Ei| sin (b\ne2\n#1585)\n"]], ["block_20", ["ad s\n:d\u2014,25i= lEil\n(8.3.3)\n"]]], "page_310": [["block_0", [{"image_0": "310_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "310_1.png", "coords": [14, 26, 281, 212], "fig_type": "figure"}]], ["block_2", [{"image_2": "310_2.png", "coords": [17, 53, 287, 145], "fig_type": "figure"}]], ["block_3", [{"image_3": "310_3.png", "coords": [22, 158, 267, 354], "fig_type": "figure"}]], ["block_4", [{"image_4": "310_4.png", "coords": [24, 234, 278, 334], "fig_type": "figure"}]], ["block_5", [{"image_5": "310_5.png", "coords": [32, 213, 171, 325], "fig_type": "figure"}]], ["block_6", ["Suppose, however, that we are dealing with unpolarized incident light, which is sometimes the\ncase. This we can View as equal parts of vertically and horizontally polarized light. Now we define\nthe scattering angle, 6, between the direction of the transmitted (unscattered) wave (x in this case)\nand the scattered wave to the detector in the x\u2014y plane. For the vertically polarized component of\nthe incident beam, the scattered wave is the same for all 6. For the horizontally polarized part,\nwhich is y-polarized, there will be no scattering along y. Figure 8.5b shows that the sin d) factor\nbecomes sin(7r/2 6) =cos 6 in our nomenclature. Thus the scattered field from the horizontal part\nvaries as cos 6, and the scattered intensity as cos2 6, whereas for the vertical part it is constant. The\nincident intensity was 50% of each, so the net intensity varies as (1 + cos2 6)/2, and we insert this in\nEquation 8.3.4 to obtain\n"]], ["block_7", ["Figure 8.5\nIllustration of the effect of incident beam polarization for a light wave propagating along x, and\nwith scattering being detected in the x\u2014y plane. (a) Vertically (z) polarized beam; q!) is the angle of scattering\nwith respect to z, and sin qb \n1 for all 6 in the x\u2014y plane. (b) Horizontally (y) polarized beam; gb is the angle\nof scattering with respect to y, and therefore sin ch sin(7r/2 6) cos 6 in the x\u2014y plane. The dependence\nof the magnitude of sin qb on 6 is also illustrated for both cases.\n"]], ["block_8", ["(a)\n"]], ["block_9", ["Scattering \n297\n"]], ["block_10", ["(b)\n"]], ["block_11", ["Io,u\n"]], ["block_12", [{"image_6": "310_6.png", "coords": [47, 52, 180, 144], "fig_type": "figure"}]], ["block_13", ["1, _ 8774a2(1 + c032 6)\n"]], ["block_14", ["\u2014\nr2166\n(8.3.6)\n"]], ["block_15", [{"image_7": "310_7.png", "coords": [68, 328, 239, 421], "fig_type": "figure"}]], ["block_16", [{"image_8": "310_8.png", "coords": [82, 340, 215, 414], "fig_type": "molecule"}]], ["block_17", [{"image_9": "310_9.png", "coords": [91, 141, 243, 240], "fig_type": "figure"}]], ["block_18", ["cos2 6\n/ 6\n"]], ["block_19", ["sin2\u00a2=1\\f\nx;\nQ\n\"\n"]], ["block_20", [{"image_10": "310_10.png", "coords": [106, 154, 229, 232], "fig_type": "molecule"}]], ["block_21", ["y\n"]], ["block_22", ["z\nL?\n"]], ["block_23", ["z\ny a\n/\\9\nx\n"]], ["block_24", ["x\n"]], ["block_25", ["y\n"]]], "page_311": [["block_0", [{"image_0": "311_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this section we adapt the Rayleigh scattering equation for an isolated object of polarizability a,\nEquation 8.3.5, to the case of a dilute, nonabsorbing polymer solution. Our vertically polarized\nincident light beam will illuminate some region of the sample solution, and the detection optics\n"]], ["block_2", ["Taking 1 Us + (4.1 x 10\u2014 19) J/photon =2.5 ><1018 photons/s. This is a lot of photons, when we bear\nin mind that sensitive photodetectors can count just a few photons per second; but it clearly is not\nenough: 2.5><1018><7.5><10_32 g 2><10\u201413 is still a very small number. Furthermore, the calcula\u2014\ntion gives the total scattered intensity over the surface of a sphere of radius\n1 m, whereas our\ndetector would only collect some small fraction of that. On the other hand, it is worth remembering\nthat if our laser beam encountered a mole of water molecules, the story would be quite different.\nAnd, in particular, sunlight traverses great distances through the atmosphere and encounters many\nmoles of gas phase species, so atmospheric scattering is far from negligible.\n"]], ["block_3", ["after taking care to put all the lengths in centimeters. This is a very small number, which does not\nlook promising. But, even though it is a very small fraction, how much light is going in? Here it is\neasiest to adopt the photon picture. A 1 W laser emits\n1 J/s; how many photons is that? Here we\nrecall the discussion following Equation 8.1.2, and find that 1 photon at this wavelength gives\n"]], ["block_4", ["8.4\nScattering from a Dilute Polymer Solution\n"]], ["block_5", ["where the subscript u denotes unpolarized incident light. This equation is associated with the name\nof Lord Rayleigh, and such scattering from independent polarizable objects is called Rayleigh\nscattering [2]. One interesting feature of Equation 8.3.5 and Equation 8.3.6 is the dependence on\nA54. This strong dependence means that shorter wavelength, higher frequency, or higher energy\nwaves are scattered considerably more. In the atmosphere, molecules and particles scatter blue\nlight down to the earth preferentially over red light, and the sky overhead appears blue. On the\nother hand, when the sun is low in the sky, the blue light is scattered away from the observer and\nthe remaining, transmitted light has a reddish hue.\nWe will use Equation 8.3.5 throughout the rest of this chapter, but it is important to remember\nthat we are assuming vertically polarized incident light, and that the scattering is detected in the\nhorizontal plane.\n"]], ["block_6", ["We\nbegin\nby\napplying\nRayleigh\u2019s\nresult,\nEquation\n8.3.5,\nassuming\npolarized\nlight for\nsimplicity. We need to know the distance to the detector, r, the wavelength of light, A0, and the\npolarizability of water, a. Let us assume we place our detector 1 m from the molecule, and that we\nare using a\n1 W argon laser with A0 2488 nm. The average polarizability of water is given as\n1.65 ><10_24 cm3 [3]. However, it is also instructive to estimate it via Equation 8.1.7, the Lorentz\u2014\nLorenz equation. The refractive index of water is about 1.333 (see Table 8.1), its molecular weight\nis 18 g/mol, and its density is about 1.0 g/cm3, so Equation 8.1.7 gives\n"]], ["block_7", ["which is pretty close to the tabulated value. Using the tabulated value in Equation 8.3.5 we have\n5 g 16 x\n(31:04\nx (1.65 x\n10:4)2 \nlo\n(100)\nx (488 x 10\u20147)\n"]], ["block_8", ["Can we measure the light scattering from a single water molecule in empty space?\n"]], ["block_9", ["Solution\n"]], ["block_10", ["298\nLight Scattering by Polymer Solutions\n"]], ["block_11", ["Example 8.1\n"]], ["block_12", ["hc\n6.63 x 10\u201834 x 3 x 108\n_19\n5\u2014)?0_\n488x10\u20189\n_4.1>< 10\nJ/photon\n"]], ["block_13", ["a X\n18\n.\u2014\n4><3.14\n1.0x6x1023\n"]], ["block_14", ["2 1.86 x 10'24 cm3\n"]]], "page_312": [["block_0", [{"image_0": "312_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "312_1.png", "coords": [28, 428, 248, 475], "fig_type": "molecule"}]], ["block_2", ["The average could be either the time average for one cell, or the ensemble average for the scattering\nvolume, as these two should be equivalent (recall from Chapter 6 that such a system is said to be\nergodic). In the actual measurement, we will record the signal for a finite amount of time, therefore\nperforming a time average, and we will record scattering from all the cells in the scattering volume,\ntherefore performing an ensemble average as well. From Equation 8.3.5, we know that the average\nscattering from any cell will depend on the average of the squared polarizability:\n"]], ["block_3", ["and we simplify this through a very important assumption, that the scattering from \ufb02uctuations in\npressure (6p) and temperature (57) is the same in the neat solvent as in the dilute solution.\nConsequently, when we consider the excess scattered intensity, [ex 130mm\u201c ISOIVCm, only the\nconcentration \ufb02uctuations matter. With these developments Equation 8.3.5 can be transformed into\n"]], ["block_4", ["Therefore the concentration derivative is\nas J}: 92)\n(8.6,\n(9c 1p\u201c 277\n66\nTip\n"]], ["block_5", ["These assumptions are not too restrictive, except that they collectively imply that Rg (and therefore\nthe molecular weight) polymer is not too big. We will deal with the large molecule case in\nthe subsequent sections. Now each cell will have an instantaneous polarizability, which can be\nexpressed the sum of an average value, (0:), plus a \ufb02uctuation, 8a:\n"]], ["block_6", ["There are three terms on the right hand side of Equation 8.4.2. The first will not contribute to the\nnet scattering from the solution, because it is the same for every cell. By the argument given in\nSection 8.2, a completely uniform material does not scatter. The second term is identically zero,\nbecause by definition (50:) :0; the \ufb02uctuations are equally likely to be positive or negative.\nConsequently, we reach the very important conclusion that the scattering is determined entirely by\nthe mean\u2014square \ufb02uctuations in polarizability,\n(80:)2\n.\nNow we need to relate 5a to \ufb02uctuations in t e thermodynamic variables p, T, and c:\n"]], ["block_7", ["will be arranged to collect from some portion of the illuminated region; we call this portion\nthe scattering volume. Now we divide the scattering volume into a large number of imaginary\n\u201ccells\u201d of volumecriteria:\n"]], ["block_8", ["where 1/\u20181\u2019 is the number of cells per unit volume. At this stage we need to work on two terms, (acr/\n8c)\" and ((802). The former is transformed using the Lorentz\u2014Lorenz equation (Equation 8.1.7):\n"]], ["block_9", ["3.\nThe cells are statistically independent, i.e., the \ufb02uctuations in any one cell are uncorrelated\nwith those in any other.\n"]], ["block_10", ["Scattering \n299\n"]], ["block_11", ["1.\n\\l\u2019m\u2019 <.\nEach cell contains many monomers, with a concentration c subject to statistical \ufb02uctuations.\n"]], ["block_15", ["1,,\n16774\n2\n1\n16774\n(Barf\n"]], ["block_16", ["a (a) + \n(8.4.1)\n"]], ["block_17", ["8a\n80:\n6o:\n5\n=\n\u2014\u2014\n\u2014\n5T\n\u2014\n8.4.\na\n(ap)T,c6p\n+\n(8T)p,c\n+\n(6C)T,p66\n(\n3)\n"]], ["block_18", ["or = x1!\n(8.4.5)\n"]], ["block_19", ["[0\nr2/\\,3\n<(\n)\n\u20181;\n72/),3\n(9C\nTip\n(\n)\n\u20181,\n(\n"]], ["block_20", ["(a2) <((a) + 5002)\n: <(a)2> + 2(a)(5a) + <(5a)2>\n(8.4.2)\n"]], ["block_21", [{"image_4": "312_4.png", "coords": [65, 523, 267, 557], "fig_type": "molecule"}]], ["block_22", [":22 113\n"]], ["block_23", ["471'\n"]], ["block_24", ["2\n1\n"]], ["block_25", ["'\n'\n"]]], "page_313": [["block_0", [{"image_0": "313_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "313_1.png", "coords": [30, 107, 168, 165], "fig_type": "molecule"}]], ["block_2", [{"image_2": "313_2.png", "coords": [31, 365, 153, 401], "fig_type": "molecule"}]], ["block_3", ["BZG/c'\ufb01c2 to a virial expansion appropriate for dilute solutions (see Section 7.4). A slight compli\u2014\ncation arises because we have been working with C as the concentration variable, and we need to\nreturn to numbers of moles, Hg, in order to handle the chemical potential. The solution volume, V,\ncan be written in terms of the partial molar volumes (see Equation 7.1.8):\n"]], ["block_4", ["by incorporating Equation 8.4.6 and Equation 8.4.10. The last transformation we need is to relate\n"]], ["block_5", ["which relates the concentration \ufb02uctuations to the associated free\u2014energy penalty. The larger the\ncost in free energy, the smaller the average \ufb02uctuations will be. (Recall from Section 7.5 that for a\nsystem at equilibrium, i.e., one that is stable, 32G/862 must be positive.) However, if the solution\napproaches a stability limit or spinodal, the denominator tends to zero and the \ufb02uctuations, and\ntherefore the scattering, can get very big indeed. The numerator of Equation 8.4.10 just acknow-\nledges the fact that the more thermal energy is available, the larger the \ufb02uctuations will tend to be.\nWe can now recast Equation 8.4.4 as\n"]], ["block_6", ["where P(SC) is the probability of a given \ufb02uctuation Sc. As positive and negative \ufb02uctuations are\nequally probable, P must be symmetric about zero (just like the Gaussian distribution for the end-\nto-end vector in Equation 6.7.1). The size of a \ufb02uctuation is related to the associated \ufb02uctuation in\nfree energy, 5G, through the Boltzmann factor:\n"]], ["block_7", ["The key part of this relation is the so-called refractive index increment, art/ac (where we drop the\nreminder about constant T and p from now on), which can either be measured precisely using a\ndifferential refractometer, or looked up in tables, as will be discussed in Section 8.7.\nNow we return to the concentration \ufb02uctuation term, ((502). We can say that\n"]], ["block_8", ["Note that the first term in Equation 8.4.9 is not symmetric about 5r: = 0, and therefore does not\ncontribute. If we insert the remaining 620/862 term into the exponential and perform the integrals\nin Equation 8.4.7 using the formula cited in Section 6.7, we arrive at the very simple relation\n(see also Problem 3)\n"]], ["block_9", ["and because we need not worry about \ufb02uctuations in volume, dV= 0, and therefore\n"]], ["block_10", ["300\nLight Scattering by Polymer Solutions\n"]], ["block_11", ["and we can expand 5G as a Taylor series (see Appendix):\n"]], ["block_12", [{"image_3": "313_3.png", "coords": [35, 485, 199, 538], "fig_type": "molecule"}]], ["block_13", [{"image_4": "313_4.png", "coords": [44, 277, 227, 310], "fig_type": "molecule"}]], ["block_14", ["2\n2\n2\nE1=W(QE) \n(8.4.11)\n[0\nrZAO\n36\n(azG/acz)\u201c,\n"]], ["block_15", ["\u20145G\nP(6(:) A exp [77,\u2014]\n(8.4.8)\n"]], ["block_16", ["V\ncm, :3a\n(8.4.12b)\nV1\n"]], ["block_17", ["66\n1\n62G\n2\nso (Elma +\n5\u2014! \n(8.4.9)\n"]], ["block_18", ["V \u201dH71 \u2018l' HQVQ\n(8.4.123)\n"]], ["block_19", ["00\nI(5\u20ac)2P(5C) (156\n((56)2 ,0\n(8.4.7)\nIP(56) dSC\n"]], ["block_20", ["2\n_\nH\n((50) > \u2014W\n(8.4.10)\n"]]], "page_314": [["block_0", [{"image_0": "314_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "314_1.png", "coords": [30, 387, 300, 449], "fig_type": "molecule"}]], ["block_2", [{"image_2": "314_2.png", "coords": [32, 283, 168, 326], "fig_type": "molecule"}]], ["block_3", ["where we have used Equation 7.4.7, the virial expansion for H/RT. To conclude,\n"]], ["block_4", ["From Equation 7.4.2 for the osmotic pressure,\n"]], ["block_5", ["wehave\n"]], ["block_6", ["The Gibbs\u2014Duhem relation n1 dul + n; dug :0 at \ufb01xed T and p gives\n%=_\ufb02%\n60\n112\n8C\n"]], ["block_7", ["30\n"]], ["block_8", ["and\n"]], ["block_9", ["Inserting 8.4.13b into 8.4.14 find\n"]], ["block_10", ["At constant T and 7.1.1)\n"]], ["block_11", ["then\n"]], ["block_12", ["Since\n"]], ["block_13", ["Scattering \n"]], ["block_14", [{"image_3": "314_3.png", "coords": [41, 484, 125, 524], "fig_type": "molecule"}]], ["block_15", [{"image_4": "314_4.png", "coords": [44, 231, 154, 270], "fig_type": "molecule"}]], ["block_16", ["32_G__Z 3+2 %__Z\n\u201d1171+\u201d2V'2 %\n862\n\u2014\nM\nV1\n60\n\u2014\nM\nr2217]\n8C\n_:K %\n\u201cCV1\n8c\n"]], ["block_17", ["3,11,]\n_3l'I\n_\n1\n\u2014=\u2014V\u2014=-\u2014VRT\u2014\u2014\n2B\n8c\n13c\n1\n(M+\n6+\n)\n"]], ["block_18", ["n = (M)\nV1\n"]], ["block_19", ["\u00a72_G_ %_Y2% X\n86'2\u2014\n30\nV1\n86\nM\n"]], ["block_20", ["66\nV;\nv\n#2\nM2\u2018TM1\n'\u2014\n66\nV]\nM\n"]], ["block_21", ["V2\nd6 #:1a + 11.2a (M2 71\u2014111)a\n"]], ["block_22", ["dnz _ V\nFIE\u2014M\n"]], ["block_23", ["nzM\n6\u20141\u2014\u2014\nV\n"]], ["block_24", ["kT\n\u2014kT\nckT\n(am/3.22) \n"]], ["block_25", ["c\n1\n\u2018VNa. (1/M+ZBc+-~)\n"]], ["block_26", ["(VN716) (%)\n"]], ["block_27", ["= VRT(l/M + 23.: + . . -)\n"]], ["block_28", ["(8.4.15b)\n"]], ["block_29", ["(8.4.13b)\n"]], ["block_30", ["(8.4.15a)\n"]], ["block_31", ["(8.4.13a)\n"]], ["block_32", ["(8.4.18)\n"]], ["block_33", ["(8.4.19)\n"]], ["block_34", ["(8.4.14)\n"]], ["block_35", ["(8.4.16)\n"]], ["block_36", ["(8.4.17)\n"]], ["block_37", ["301\n"]]], "page_315": [["block_0", [{"image_0": "315_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This expression is usually rearranged in either of two ways, as follows:\n"]], ["block_2", ["1.\nIn the limit of low concentration Equation 8.4.24b reduces to Ie,~(number of scatter-\ners)><(size of scatterer)~cM, just as advertised for incoherent scattering in Section 8.2.\nNote that Equation 8.4.24a, somewhat perversely, has the experimental signal-the scattered\nintensity\u2014in the denominator, so this result is not so obvious.\n2.\nIn the limit of infinite dilution, light scattering measures M, as does the osmotic pressure.\nThese two experiments are intimately related, because the light scattering is determined by\nconcentration \ufb02uctuations, and their amplitude is related to the associated osmotic cost. In\nother words, one could imagine a semipermeable membrane around each of our fictional cells.\nAny extra polymer in a particular cell will drive up II, whereas a lower-than-average\nconcentration would necessarily cause II to increase in some other cells. Thermal energy\ndrives random \ufb02uctuations, but the osmotic compressibility resists them.\n3.\nThe virial coefficient is obtained from the concentration dependence of the scattering, just as\nin the osmotic pressure experiment. Equation 8.4.24b shows directly that in a good solvent,\nwhen B > 0, the intensity will first increase linearly with c, but the rate of increase will drop\n"]], ["block_3", ["where the second version is obtained by recalling that 1/(1 +x) 1 \u2014x+x2+ -. The former\nversion is more often used when plotting data, because the right hand side should be linear in 6,\nwhereas the latter is more transparent in terms of the physical content. These equations tell us\nseveral important things.\n"]], ["block_4", ["or\n"]], ["block_5", ["to reach the result\n"]], ["block_6", ["where R9 is the so\u2014called Rayleigh ratio. It is the normalized excess scattered intensity per unit\nvolume, with the purely geometrical quantity r factored out; therefore R9 should depend only on\nthe solution and A0, and not the instrument used to measure it. Note that R9 has units of cm\u2018l,\nbecause 1.,x is the scattered intensity per unit volume, whereas I0 is just the incident intensity.\nSimilarly, the purely optical factors can be grouped\n"]], ["block_7", ["and as we have been working with the \ufb02uctuations in a cell of volume \u20181\u2019, we set V \u20181\u2019 in Equation\n8.4.11. The \ufb01nal result is therefore\n"]], ["block_8", ["302\nLight Scattering by Polymer Solutions\n"]], ["block_9", ["where the volume of the fictional cell, \u20181\u2019, has happily disappeared.\nWe now regroup some terms:\n"]], ["block_10", ["KC\n1\n._=._\n23\n8.4.24\nR9\nM \nc +\n(\na)\n"]], ["block_11", ["4\n2\n2\n2\nK E\n77 \u201d (an/ac)\n(3.4.22)\nAONW\n"]], ["block_12", ["R9 KcM{l 230M }\n(8.4.24b)\n"]], ["block_13", ["K\nR9 T\u2014C\u2014\u2014\n(8.4.23)\n__\n23\n. . .\nM \nc +\n"]], ["block_14", ["[ex\n2\n1r E R9\n(8.4.21)\n"]], ["block_15", ["ex\n4\n2\n2\n2\nI\u2014 \nw \" (\u00a7\u201d/3C)\n1\nC\n(8.4.20)\n1.,\nrZAONa,\n(M + 236 + - - .)\n"]]], "page_316": [["block_0", [{"image_0": "316_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "316_1.png", "coords": [29, 395, 174, 429], "fig_type": "molecule"}]], ["block_2", ["_\u2014\nc (g/mL)\n0.0011\n0.0026\n0.0039\n0.0052\n0.0067\n0.0089\n"]], ["block_3", ["1'1.,><105 (cm\u2014l)\n1.87\n3.75\n5.09\n6.06\n7.02\n8.04\n"]], ["block_4", ["Thus light scattering measures the absolute weight average molecular weight, whereas the osmotic\npressure experiment gives the number average (recall Equation 7.4.11). The reason for this\ndifference is that the intensity scattered from an individual polymer is proportional to M, so that\nalthough the \ufb02uctuations are determined by the number of molecules per unit volume, the\nscattering signal is weighted by an additional factor of M.\n"]], ["block_5", ["The following light scattering data were obtained on solutions of polystyrene in toluene at 25\u00b0C.\nA polarized Helium\u2014Neon laser was used (AD: 633 nm); 871/80 was found to be 0.108 mL/g, and\n71: 1.494. Calculate the weight average molecular weight and the second virial coefficient.\n"]], ["block_6", ["when the B term contributes appreciably. On the other hand, in a poor solvent, with B m 0 or\neven negative, the intensity will increase more rapidly with further increase in c. This is\nillustrated schematically in Figure 8.6.\n4.\nWe reiterate that Equation 8.4.24a and Equation 8.4.24b are valid for \u201csmall\u201d polymers only;\nwe will explore this restriction more quantitatively in the following section.\n5.\nAlthough we have stressed the intimate connection between light scattering and osmotic\npressure, there is one important difference. For a dilute but polydisperse sample, we can\n"]], ["block_7", ["Figure 8.6\nThe excess scattered intensity as a function of concentration for different values of the second\nvirial coefficient, B.\n"]], ["block_8", ["and so\n"]], ["block_9", ["Scattering \n303\n"]], ["block_10", ["Example 8.2\n"]], ["block_11", ["q;\n.I'\n.IIII\".\n:\n--.\n|'\nJ\n-\n:\nI'I'J\u2019H\u2019r\n\u2018\n5 :-\n\u201951:?\"\n13> 0\n\u2018:\n"]], ["block_12", ["(t;\n:\noil.\" I J\u201c:\n"]], ["block_13", [":2\n:\nIII-l\n-\n"]], ["block_14", [{"image_2": "316_2.png", "coords": [40, 55, 284, 216], "fig_type": "figure"}]], ["block_15", ["10 :'\nB < 0\ni \n.\nI.\n_f\n\"'\n"]], ["block_16", [{"image_3": "316_3.png", "coords": [43, 445, 211, 489], "fig_type": "molecule"}]], ["block_17", ["15\n_\nI\n_l\u2014I_\n]\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\u2014\nI\n_\nI\n:1\n"]], ["block_18", ["see that\n"]], ["block_19", ["61\u201413(1) R9\n[(26n\n2%:\nMw\n"]], ["block_20", ["\u00b0\n=\n=\n-\n,-\n.42\n1133129\nKcM\n[(20,114\n(8\n5)\n"]], ["block_21", ["715M;\nKC\nKEG; \u201427\n\u20141\u2014\u2014\n(8.4.26)\n1\u2018\n_:_________=\n"]], ["block_22", ["-\nL\n_l\nI\nL\nI\ni.\nI\ni\nI _I_\nI\nI_I_\nI\nI\ni\nI\nI\nI\n-\n00\n2\n4\n8\n10\n"]], ["block_23", ["E\na?\"\n2\n"]], ["block_24", ["C, a.u.\n"]]], "page_317": [["block_0", [{"image_0": "317_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "317_1.png", "coords": [20, 58, 324, 245], "fig_type": "figure"}]], ["block_2", ["We now take into account the finite size of a polymer. Unfortunately, the mathematics is a little\ntedious, but there is no shortcut. On the other hand, the main result will turn out to be relatively\nsimple. We will restrict ourselves to the limit of a very dilute solution, i.e., we only consider one\npolymer at a time. The basic idea was outlined at the beginning of the chapter, namely that if the\ndistance, rjk, between two monomers j and k on a chain is a significant fraction of the radiation\nwavelength, it, then the waves scattered from each monomer will have a phase difference at the\ndetector. This leads to some destructive interference and a net reduction in the scattered intensity.\nThis phase difference will depend on the scattering angle, 9, and therefore the intensity will also\ndepend on 9; thus we are dealing with coherent scattering. Operationally, we can define a form\nfactor for a single polymer, P(9), as\n"]], ["block_3", ["where the Rayleigh scattering is given by Equation 8.4.24. From this relation we can see that\n0\u00a7P(9) _<__ 1. In general, we should worry about interference between waves scattered from\n"]], ["block_4", ["Then we plot KC/Rg versus 0, as shown in Figure 8.7. Linear regression gives an intercept of\n5.46x10\u20186 mol/g and a slope of 7.00><10\"4 mL moi/g2. Thus we obtain\n"]], ["block_5", ["First we need to calculate K by Equation 8.4.22:\n"]], ["block_6", ["8.5\nThe Form Factor and the Zimm Equation\n"]], ["block_7", ["304\nLight Scattering by Polymer Solutions\n"]], ["block_8", ["Figure 8.7\nPlot of data for Example 8.2 according to Equation 8.42a.\n"]], ["block_9", ["Solution\n"]], ["block_10", ["a?\nE\nE\n\u201c:3\n_\n_\n5\u2018\n5 X 10_5 \nlntercept = 5.46 X 10'5\n_-\n"]], ["block_11", ["5 ><10'5 :\u2014\n-:\n9.?\n-\n_\n\u20185\n-\n-\nE\n-\n_\n"]], ["block_12", ["1'5 X 10\u20145\n_\nI\nI\nI\nI\nl\nI\nI\nl\nr\n1\nI\nI\nr\nI\nI\nI\nl\nr\n_l\nJ\n\u2018\n1\n\u2014-\n-I\n"]], ["block_13", ["Actual 16,49)\nP 0 =\n.5.1\n( )\nRayleigh I...(0)\n(8\n)\n"]], ["block_14", ["B 7.00 x 10-4/2 3.5 x 10\u20144 mL moi/g2\n"]], ["block_15", ["M... 1/(5.46 x 10\u20146) 183,000 g/mol\n"]], ["block_16", ["_ ._..\n\u20147\n2\n2\n_\n(6.33 x 10\u20145)4 x 6.02 x 1023 T 1\u201906 x 10\ncm mOI/g\n"]], ["block_17", ["0\n"]], ["block_18", ["0\n0.002\n0.004\n0.006\n0.008\n0.01\n"]], ["block_19", ["E\nSlope = 7.00 x 10-4\nI\n"]], ["block_20", ["E\nE\n"]], ["block_21", ["|-\n"]], ["block_22", ["1\nI\nl\nI\nI\nL\n|\nI\nI\nI\nl\nl\nI\nl\nl\nI\nl\n|\nI\n"]], ["block_23", ["c, g/mL\n"]], ["block_24", ["\u2014\n"]]], "page_318": [["block_0", [{"image_0": "318_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "318_1.png", "coords": [33, 409, 285, 476], "fig_type": "molecule"}]], ["block_2", ["To develop a mathematically precise de\ufb01nition of P09) we consider each monomer to be a\nRayleigh scatterer. electric field scattered by the molecule, Ema, is given by the superposition\nof the fields \n"]], ["block_3", ["where 50,, is the amplitude of the field scattered from each monomer and5}- is the phase of the\nwave from choose to use\nsums overj and k to avoid confusion with i \\/\u2014\u20141.) In Equation 8.5.2 we have assumed that each\nmonomer is identical in scattering power (i.e., a homopolymer). The scattered intensity requires\nthe squaring of the total field:\n"]], ["block_4", ["The Form \n305\n"]], ["block_5", ["different \nbecome the operational of what called the solution structure factor, S(9). However,\n"]], ["block_6", ["because \nthe positions of different polymers and the intermolecular interference terms do not contribute.\n"]], ["block_7", ["8.5.1\nMathematical Expression for the Form Factor\n"]], ["block_8", ["For Rayleigh scattering we assumed that each polymer was very small compared to A, and\ntherefore the phase is the same for each monomer along the chain: 5,- 8k. With this simplification\nwe can write\n"]], ["block_9", ["We have also taken the average (el1(51\u20145))because thatrs the experimentally important quantity.\nEquation 8.5.4 is an alternative definition of P(6), and again we emphasize that it is the single\nchain form factor; the double summation is over the N monomers of one chain. For the general\ncase, we would just extend the double summation over all pairs of monomers in the scattering\nvolume, and then Equation 8.5.4 would define the structure factor.\nThe next step is to develop an expression for 5k 5}, the phase difference between the waves\nscattered from any pair of monomers k and j. Here the scattering vector comes to the rescue,\nbecause this phase difference is very simply expressed as the dot product of \u00a7 with the vector\nseparating monomersj and k:\n"]], ["block_10", ["If we recall the derivation of Bragg\u2019s law in Section 8.2, two particles in a given lattice plane\nautomatically scatter waves in phase with one another. In this case \u00e9\u2019 \u00b0 I}, = 0, because 5 is\nperpendicular to the lattice planes. In other words, the phase difference is determined solely by\nthe component of rj\u2014k that is parallel to the scattering vector.\n"]], ["block_11", [{"image_2": "318_2.png", "coords": [35, 177, 309, 223], "fig_type": "molecule"}]], ["block_12", [{"image_3": "318_3.png", "coords": [41, 319, 168, 362], "fig_type": "molecule"}]], ["block_13", ["5k 5} I? FIk\n(8.55)\n"]], ["block_14", ["N\nN\n_'\nT\n\"'\n\u2014iwt\n-i8-\n_'\niwt\ni5\nIs IEsftot \ns.totl N (53?t\nE\n6\nJ\n\u00b0 (Ems e\nE\ne\n1\u2018\n"]], ["block_15", ["N\nN\n_.\n2\n.\n= IEo.si\nE\nE\n61(8r6\nj=l k=l\n"]], ["block_16", ["M2\nM2\n_.056\u201c\u201d:65\u2019\n(8.5.2)\nj=l\nj=l\n"]], ["block_17", [{"image_4": "318_4.png", "coords": [115, 277, 329, 341], "fig_type": "molecule"}]], ["block_18", ["1\n"]], ["block_19", ["\u2018\nz\n1%\n\n(8.5.4)\n"]], ["block_20", [{"image_5": "318_5.png", "coords": [164, 287, 329, 325], "fig_type": "molecule"}]], ["block_21", ["j=1\n"]], ["block_22", ["(8.5.3)\n"]]], "page_319": [["block_0", [{"image_0": "319_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "319_1.png", "coords": [19, 465, 461, 602], "fig_type": "figure"}]], ["block_2", [{"image_2": "319_2.png", "coords": [26, 624, 236, 675], "fig_type": "molecule"}]], ["block_3", ["Then we can write\n"]], ["block_4", ["where we have looked up the integrals for sin 6 cosk 6. The resulting series in square brackets is, in\nfact, nothing more than\n"]], ["block_5", ["using the transformation to spherical coordinates (if this is unfamiliar, see the Appendix). Then we\neliminate the dot product:\n"]], ["block_6", ["in which q (47ml) sin(6/2) from Equation 8.2.4. If we expand the exponential as a power series\nthis becomes\n"]], ["block_7", ["Now, if we restrict our attention to isotropic samples, such that in spherical coordinates (r,6,(;b) P(FJ}k)\nhas no 6 or d) dependence, then we can get rid of the vectors altogether. First, we recognize that\n"]], ["block_8", ["where P073) is the probability of monomersj and It being separated by a vector i}. For the case of a\nchain in a theta solvent (recall Chapter 6 and Chapter 7), we already know this probability\nfunction: it is a Gaussian function (Equation 6.7.1).\n"]], ["block_9", ["8.5.2\nForm Factor for Isotropic Solutions\n"]], ["block_10", ["This equation represents the standard definition of the form factor. It may appear rather compli-\ncated, because it involves a complex number and two vectors. However, remember that the\ncomplex numbers are just a convenience and they will disappear when we calculate anything\nobservable. Furthermore, the dot product of two vectors is a scalar, so we should be able to get rid\nof the vectors as well. The average in Equation 8.5.6 is\n"]], ["block_11", ["306\nLight Scattering by Polymer Solutions\n"]], ["block_12", [{"image_3": "319_3.png", "coords": [45, 74, 220, 113], "fig_type": "molecule"}]], ["block_13", ["(67302\n(67004\nsin qr k\n2 1\u2014\n+\n+\n_\u20142\u2014\u2014\u2014\n1\n(8.5.11)\n3!\n5!\nqr};c\n"]], ["block_14", ["H\n..\n'\nizqzrf\u2018k cos2 6\nexp[1q\n- rjk] : 1 +1qrjkcos6 +\n2'\n"]], ["block_15", ["exp[iq\u2018 Fir] exp[iqrjk cos 9]\n(8.5.9)\n"]], ["block_16", ["1\nN\n_.\n..\nP(6)=:\n-\n"]], ["block_17", ["dfj-k ax), am dz), :rj, sin 6 d6 dqb drjk\n(8.5.8)\n"]], ["block_18", ["Now we insert this result into Equation 8.5.4:\n"]], ["block_19", ["273'\n\u20196'\n(alpha 8.1) = mam; dry]. Jdd) d6 sm6 + iqrj, cosg \n"]], ["block_20", ["\u2014\u2014]\u2014V-\u20142\u2014 \n(exp[iq Uri)\n(8.5.6)\n"]], ["block_22", ["ll\n"]], ["block_23", ["$62\nq4\ufb01z\n2!3\n 4!5 \nll\n"]], ["block_24", ["0=\u2014-\u2014-\u2014,8\n"]], ["block_25", ["(ah\u2014,8\n"]], ["block_26", ["rah\u2014.58\n"]], ["block_27", ["P(rJ,-k)rj2k eqrJ:(sin 6 + iqrjk cos 6 sin 6 +\n"]], ["block_28", ["P(rjk)2wr-k drjk\n(8.5.10)\n"]], ["block_29", ["0\n"]], ["block_30", ["0\n"]], ["block_31", [{"image_4": "319_4.png", "coords": [198, 558, 342, 597], "fig_type": "molecule"}]], ["block_32", ["-222\n2\n-\n1\nq r- cos 6s1n6+-\u00ab-\n1\"\n2\n)d6\n"]]], "page_320": [["block_0", [{"image_0": "320_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "320_1.png", "coords": [27, 311, 148, 355], "fig_type": "molecule"}]], ["block_2", ["The Form Factor and the Zimm Equation\n307\n"]], ["block_3", ["We now return to the end of the previous section, and insert P(q)=P(9) into the scattering\nequation (Equation 8.4.24a):\n"]], ["block_4", ["Equat10n6.5.8:\n1:11:21\n"]], ["block_5", ["so in the end,\n"]], ["block_6", ["where P(rjk) z 4771'; P\ufb01k).\nThis assumption of isotropy is not at all restrictive for a dilute polymer solution at rest. Note that\nit is not an assumption that the shape of the molecule is spherical, only that on average the\nintrarnolecular bond vectors point equally in all directions. Thus, even a solution of rod\u2014like\nparticles can be isotropic in this sense. We can use Equation 8.5.12 whenever we have information\nabout P(rjk). However, it turns out we can also extract something very useful even if we do not\nknow anything particular about this distribution.\n"]], ["block_7", ["because 1/(1 x) =1 + x + x2 + -.We now use this result to obtain the Zimm equation:\n"]], ["block_8", ["The power series in qr 1,given in Equation 8.5.11, combined with Equation 85.6, gives for P(q):\n"]], ["block_9", ["N\nN\nBut, the average 2 Z is directly related to (82>(2 R2) for any shape, as shown \n"]], ["block_10", ["8.5.4\nZimm Equation\n"]], ["block_11", ["independent of the shape of the particle. Thus if the experiment is designed such that qE < 1, the\nhigher order terms in the expansion Equation 8.5.15 can be neglected, and Rg can be determined\nwithout any prior knowledge of the average conformation. It turns out that for \ufb02exible and semi\u2014\n\ufb02exible polymers this condition is quite often satis\ufb01ed (see Example 8.4 and Problem 6).\n"]], ["block_12", ["in the limit of c \u2014\u2014> 0. Now it is traditional to manipulate 1/P(9) as follows:\n"]], ["block_13", ["8.5.3\nForm Factor as n \u2014> 0\n"]], ["block_14", ["This is the fundamental result for light scattering from dilute polymer solutions [4].\n"]], ["block_15", ["and therefore we have the important result that\n"]], ["block_16", [{"image_2": "320_2.png", "coords": [37, 66, 279, 121], "fig_type": "molecule"}]], ["block_17", [{"image_3": "320_3.png", "coords": [40, 618, 217, 650], "fig_type": "molecule"}]], ["block_18", ["P(q)= A1\u2014,ZZ{1\n<13? >+%...}\n(8.5.13)\n\u20181_\n"]], ["block_19", ["2\nP(q)=1\u2014%R\u00a7+-~\n(8.5.15)\n"]], ["block_20", ["KC\n1\nl\n_ z \u2014 \u2014\u2014\n28\n8.5.16\nR9\nMW 19(0) +\nc +-\n(\n)\n"]], ["block_21", ["1\n1\n1\n2\n:1\n_\n2R2\n8.5.17\n13(9)\nqz\n2\n+ 3 q\ng \n(\n)\n"]], ["block_22", ["Kc\n1\nt]2\n2\n__:__\n\u2014~\u2014~R\n2\n8.5.18\nR9\nMw(1+3\ng+\n)+Bc+\n(\n)\n"]], ["block_23", ["__,\n_,\nsin\nr-\nsin\nr1\n1\nN\nN\n2\nz W ZZ <51)\n(8.5.14)\n)\u2018=1 k=1\n"]], ["block_26", [{"image_5": "320_5.png", "coords": [77, 77, 265, 110], "fig_type": "molecule"}]], ["block_27", ["1\n\u2014\u2014\n313g + - --\n"]], ["block_28", ["1= 1 k:\n6\n"]]], "page_321": [["block_0", [{"image_0": "321_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Equation 8.5.18 provides the basis for a particular method to analyze light-scattering data, the so-\ncalled Zimm plot. We need to perform two extrapolations, to zero scattering angle (6 0\u00b0 or q =0)\nand to zero concentration (23c=0), and the result will be 1 /Mw. Furthermore, the slopes of the\nangle extrapolation and the concentration extrapolation, respectively, will provide values of Rg and\nB, assuming that the range of the independent variable is such that only the first term of the\nrelevant expansion is important. Accordingly, we would plot KC/Rg versus sin2(6/2) + 7c, where -y\nis an arbitrary constant; the sin2(6/2) term is proportional to ([2. The value of -y is chosen to spread\nthe data out. An example is shown in Figure 8.8 for a solution of methylcellulose in water. One\n"]], ["block_2", ["C.\u2014I., and Lodge, T.P., Macromolecules, 32, 7070, 1999. With permission.)\n"]], ["block_3", ["Figure 8.8\nZimm plot for a sample of methylcellulose in water. (Reproduced from Kobayashi, K., Huang,\n"]], ["block_4", ["You should confirm that if you start from Equation 8.5.18, and transform it appropriately you will\nrecover Equation 8.5.19. It is important to point out a subtlety here. You might well ask, if c is high\nenough that polymer\u2014polymer terms contribute to the concentration \ufb02uctuations (i.e., through B),\nwhy don\u2019t we have to worry about interference effects between monomers on nearby chains? The\nanswer is that actually we do, and that is the source of the second P(6) term multiplying B in Equation\n8.5.19. In the simplest approach (which is not simple), the single contact approximation introduced\nby Zimm [4] (Equation 8.5.19) is the result. We should also point out that several texts write the right\nhand side of Equation 8.5.18 incorrectly, as (l/MW + 23c + -) (1 + q2/3R\u00e9 + ).\n"]], ["block_5", ["308\nLight Scattering by Polymer Solutions\n"]], ["block_6", ["8.5.5\nZimm Plot\n"]], ["block_7", ["E\nl\u2018\n.\ngrit-Fir...\u201d-\nI\nJ\n"]], ["block_8", [{"image_1": "321_1.png", "coords": [43, 368, 334, 611], "fig_type": "figure"}]], ["block_9", ["If we recast this relation from Equation 8.4.24b, we obtain\n"]], ["block_10", ["R9 KcMwP(6){1 ZBcMwP(6) . . .}\n2\n2\n(8.5.19)\n_\nq\n2\nq\n2\n\u2014KCMw(1\u2014\u20183\u2014Rg\"'){l_ZBCMw(1_\u00a7'Rgv--)\n...}\n"]], ["block_11", [{"image_2": "321_2.png", "coords": [50, 395, 295, 532], "fig_type": "figure"}]], ["block_12", [{"image_3": "321_3.png", "coords": [52, 72, 315, 127], "fig_type": "molecule"}]], ["block_13", ["10x10\u20145\nl\n1\nI\nI\nl\u2014\nr\nI\u2014I\nT]\nI\nI\nI\n'l\u2014l__'l_\u2018\u2014l'\nFl\nI\nI\nI\nI\u2014\u2018l\nI\nI\nI\nI\n"]], ["block_14", ["_\n,9\n_1\n\u2018e'\nE\nl\n_1\nH\na\n'\nE\nE.\n\u201cg\nE\u201d\n-\n-...__\no)\n-\nE)\nE\no\n3))\n2x106 -\n0\nL0.\n0\n\u2014\n"]], ["block_15", ["G\n.\nI\n'\n\u201dI\n,'\n.\n19:30\"\nj\n4X10F6\n\u2014\ng\n"]], ["block_16", ["r\nwe\" .\n-\n.\nI\n6x10\u20186 \n-\n_ \n.\n,\n\u2014\n"]], ["block_17", ["erg-\u201dff\u201d,\n'\n8X10\u20146 \n.\nO\n.\n.\n...|\n"]], ["block_18", ["ub>0ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nub>Lub>\nub>ub>nub>ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nub>Lub>J\nub>a.ub>\nub>tub>\nub>ub>nub>ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nub>1ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nl J_ub>Lub>\nu\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>\nub>ub>ub>ub>ub>Iub>ub>ub>ub>ub>. l\nub>ub>nub>ub>\nub>Lub>\nr\nr\n"]], ["block_19", ["O\n1\n2\n3\n"]], ["block_20", ["_\nf\n.0\nI\nI\ni\n"]], ["block_21", ["'-\n'0',\u201d \n"]], ["block_22", ["-\n\ufb02\n.1\n641500\n"]], ["block_23", [".\nKi\n\u201c'03\n(I?\nn\n'\n(.3\n\u2018\ufb01\u2019\nn\no\no\no\n"]], ["block_24", ["amp-(912) + 30000\n"]]], "page_322": [["block_0", [{"image_0": "322_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "322_1.png", "coords": [25, 438, 341, 577], "fig_type": "table"}]], ["block_2", [{"image_2": "322_2.png", "coords": [25, 409, 381, 611], "fig_type": "figure"}]], ["block_3", ["Solution\n"]], ["block_4", ["The \n309\n"]], ["block_5", ["The scattering data for the methylcellulose sample in Figure 8.8 are given in the following\ntable. The data were taken at 20\u00b0C, with an argon ion laser operating at 488 nm. Under these\nconditions, 72 for water is 1.33, and Biz/ac was determined to be 0.137 mL/g. Calculate MW, Rg, and B.\n"]], ["block_6", ["choice of is l/Ac,Ac is the difference between successive concentrations.\nIn this case all of the angle data for a given concentration \nof the angle concentration. In any event, the choice \ncosmetic. regression on each set of data,\nshifted concentration be fit to a straight line against sin2(6/2), =0\u00b0\nintercept \nangle 6, Kc/Re should be fit to a straight line against 0, and the c=0 intercept recorded (and\nplotted). Then, should be fit to a straight line against 0, and the c 0 intercepts\nto a straight line against sin2(6/2). Both of these lines should meet on the KC/Rg axis, and the\nresulting slopes of these lines are proportional to R: and B, respectively, \nthe proportionality depending on exactly how the data were treated. This process may be under-\nstood in detail by working through Problem 10, or the following example.\n"]], ["block_7", ["The horizontal axis of the Zimm plot requires a choice of the shift factor 7. As the different\nconcentrations are separated by Ac 9:: 0.1 or 0.2 mg/mL, a reasonable choice for 37 would be in the\nrange l/Ac % 5000\u201410,000; in fact, a value of 3000 was selected for Figure 8.8. The next step is to\nperform the first extrapolation to zero angle for each concentration. This may be done directly from\nthe data in the table, using sin2(6/2) as the x axis, by linear regression (\u201cleast squares\u201d). The\nresulting 6=0\u00b0 data are listed below in the second table, and these extrapolations are shown\nin Figure 8.9a, with the 620\u00b0 data highlighted as solid circles. Note that to plot the data in\nFigure 8.9a, each value of sin2(6/2) has to be added to the appropriate value of 7c.\n"]], ["block_8", ["6 (\u00b0).\n311120372)\nKc/R9x10\u00b0\n"]], ["block_9", ["6 (mg/mL)\n0.21\n0.30\n0.50\n0.59\nKc/R9x10\u00b0,6=0\u00b0\n3.42\n3.62\n4.03\n4.13\n"]], ["block_10", ["The second extrapolation is to c=0 for each angle. The resulting linear regression curves and\nintercepts are shown in Figure 8%, and the c = 0 values are listed in the following table.\n"]], ["block_11", ["Example 8.3\n"]], ["block_12", ["30\n0.0670\n3.61\n3.87\n4.21\n4.29\n40\n0.117\n3.86\n4.16\n4.57\n4.71\n50\n0.179\n4.25\n4.43\n4.89\n5.07\n60\n0.250\n4.71\n4.95\n5.38\n5.48\n70\n0.329\n5.14\n5.34\n5.76\n5.84\n80\n0.413\n5.60\n5.78\n6.20\n6.29\n90\n0.500\n6.01\n6.08\n6.59\n6.72\n100\n0.587\n6.37\n6.43\n7.02\n7.14\n110\n0.671\n6.83\n6.89\n7.35\n7.59\n120\n0.750\n6.96\n7.23\n7.66\n7.88\n130\n0.821\n7.49\n7.66\n8.14\n8.25\n140\n0.883\n7.66\n8.00\n8.42\n8.57\n150\n0.933\n7.92\n8.20\n8.65\n8.82\n"]], ["block_13", ["0.21 mg/mL\n0.30 mg/mL\n0.50 mg/mL\n0.59 mg/mL\n"]]], "page_323": [["block_0", [{"image_0": "323_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "323_1.png", "coords": [31, 305, 278, 529], "fig_type": "figure"}]], ["block_2", ["Finally, these two sets of data should be extrapolated to 6 =0 and 6=O\u00b0, respectively. The\nresults are shown in Figure 8.90. From these two straight lines, the desired information can be\nextracted as follows:\n"]], ["block_3", ["Figure 8.9\nConstruction of the Zimm plot of Figure 8.8 as developed in Example 8.3. (a) The extrapolation\nto 9 O for each concentration. (b) The extrapolation to c =0 for each angle.\n(continued)\n"]], ["block_4", ["310\nLight Scattering by Polymer Solutions\n"]], ["block_5", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n.I\nI\nI\nI\nI.\nI\nI\nI\nI\nI\nI\nI\nI.\nI.\nI\nI\nI\nC)0\n1\n2\n3\n(a)\nsin2(9/2) + 30000\n"]], ["block_6", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nL\nI\nI\n.I.\nl\nI\nI\nI\nI\nO0\n1\n2\n3\n(b)\nsin2(9/2) + 30000\n"]], ["block_7", [{"image_2": "323_2.png", "coords": [37, 36, 269, 262], "fig_type": "figure"}]], ["block_8", ["a\n6 \n-\n"]], ["block_9", ["\u00abg\n6 \ufb02\n-\n"]], ["block_10", ["e\n-\n_\n\u00a7\n4 \n-\n"]], ["block_11", ["x\nI\nI\nit?\n-\n.\n\u00a7\n4 \n-\n"]], ["block_12", [{"image_3": "323_3.png", "coords": [44, 334, 270, 455], "fig_type": "figure"}]], ["block_13", ["10\n"]], ["block_14", ["10\n'\n'\n'\n'\nI\n'\n'\n'\n'\nI\n'\n'\n'\n'\nI\n'\n'\n'\n'\nI\n.\n'\n.\n'\nI\n'\n'\n'\n-\n_\nJ\n"]], ["block_15", ["2\n_\n_\n"]], ["block_16", ["8\n..\n_\n"]], ["block_17", ["2\n\u2014\n_\n"]], ["block_18", ["8 \n%\n-\n"]], ["block_19", ["r\nJ\n"]], ["block_20", ["-\n4\n"]], ["block_21", [{"image_4": "323_4.png", "coords": [77, 63, 265, 195], "fig_type": "figure"}]]], "page_324": [["block_0", [{"image_0": "324_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "324_1.png", "coords": [29, 593, 300, 642], "fig_type": "molecule"}]], ["block_2", ["sin2(9/2)\n0.0670\n0.117\n0.179\n0.250\n0.329\n0.413\n0.500\nKc/Rgx 106 6\u2018 =0\n3.28\n3.45\n3.79\n4.31\n4.76\n5.22\n5.54\nsin2(6/2)\n0.587\n0.671\n0.750\n0.821\n0.883\n0.933\nKC/R9X106 I: 0\n5.85\n6.33\n6.48\n7.04\n7.23\n7.47\n"]], ["block_3", ["The angle extrapolation is\n"]], ["block_4", ["therefore\n"]], ["block_5", ["The \n311\n"]], ["block_6", ["Both fits give\na common c=0, 620\u00b0 intercept of 3.03x10\u20186. Thus MW is obtained\nas\n1/(3.03 x10'6)= 3.3 ><105 g/mol.\nThe concentration extrapolation is\n"]], ["block_7", ["Consequently,\n"]], ["block_8", ["and Rg 64 nm.\n"]], ["block_9", ["Figure 8.9 (continued)\n(c) The extrapolations of the 0 0 data to 6 =0, and the 6 =0 data to 0 =0.\n"]], ["block_10", ["0o\n6 \n-\n"]], ["block_11", [{"image_2": "324_2.png", "coords": [38, 528, 300, 587], "fig_type": "molecule"}]], ["block_12", ["x\ni\n_\n9f\n-\n_\n"]], ["block_13", ["(X)\n4//\n"]], ["block_14", [{"image_3": "324_3.png", "coords": [40, 48, 262, 230], "fig_type": "figure"}]], ["block_15", ["I\nI\nI\nI\nl\nI\n_I_\nI\nI\nI\nI\nI\nJ\nI\nI\nI\nI\nO0\n1\n2\n(c)\nsin2(e/2) + 30000\n"]], ["block_16", ["11\u20192 \n3\n=\n0\n2\n8\n3.03 x 10-6 \nX (4 x 3.14 x 1.33)\n4 73 \u201cm\n"]], ["block_17", ["K\n1\n_C 2 _ + 236 = 3.03 x 10-6 + 0.64 x 10-6 (3000c)\n"]], ["block_18", ["B (1/2) x 3000 x 0.64 x 10-6 9.6 x 10-4 mo] cnII3/g2\n"]], ["block_19", ["Kc\n1\n1\n1\n1 166-2712\nR9\nMw( +3q\n3)\nMW< +3\n33\n3111(9/ ) g\n"]], ["block_20", ["1o\n-\n-\n-\n-\nI\n.\n-\n-\n-\nI\n- =\u2014-\n-\nI\n"]], ["block_21", ["8 \ny: 3.03 + 4.82X\n_\n"]], ["block_22", ["y: 3.03 + 0.64x\nj\n24\n-\n"]], ["block_23", ["= 3.03 x 10-6 + 4.82 x 10-6 sin2(6/2)\n"]], ["block_24", ["4.82 x 10-6\n488\n2\n"]], ["block_25", ["_\n"]]], "page_325": [["block_0", [{"image_0": "325_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We need to calculate Rg for each polymer, and the minimum and maximum q values of the\ninstrument. In Example 6.2, we calculated the unperturbed Rg for polystyrene with M\u2014\n\u2014 105 to be\n8.5 nm. As Rg varies with \\/_\nMwe can estimate R3 for 104 to be \\/_Otirnes smaller, i.e., 2.7 nm, and\nRg for 10\u20185 to3 be \\/_Otimes bigger, or 27 nm. For cyclohexane, n: 1.424 from Table 8.1 (at a\n"]], ["block_2", ["Estimate which of the four regimes would be accessed by light scattering on solutions of\npolystyrene in cyclohexane at the theta temperature (345\u00b0C). Assume M values of 104, 105, and\n106 g/rnol, and that the angular range of the instrument is 25\u00b0\u2014150\u00b0.\n"]], ["block_3", ["qr\u2014I is essentially a \u201cruler\u201d in the sample. The choice of q dictates that the scattering experiment\nwill explore \ufb02uctuations on the length scale cfl. In Bragg diffraction, intensity is found at a\nparticular angle when the relation m}t 20 sin 6/2 is satisfied (Equation 8.2.2), where D is the\nspacing between planes of atoms in the crystal and m is an integer. The Bragg equation just\naccounts for the fact that the phase shift is an integral number of wavelengths when moving from\none lattice plane to the next, and thus the interference is constructive. The Bragg equation for first\norder (m 1) can be rewritten D\u20181 (2/)1) sin(6/2) q/2ar. In other words, scattering is exactly the\nsame process as Bragg diffraction. The only difference is that in scattering, we look for spontan-\neous \ufb02uctuations that just happen to have the right orientation and spacing (2dr/q) to scatter to the\ndetector, whereas in diffraction there are particular lattice planes. Now we can reinterpret the four\nregimes above in terms of the relationship between the spacing of the lines and the size of the\nmolecule. This is illustrated in Figure 8.10, where a set of lines spaced at 277/q is shown with four\ndifferent polymers, illustrating the four size regimes.\n"]], ["block_4", ["1.\nIfn <<1, which means either small molecules or very small scattering angles, then P(q) 2:: 1,\nThus the molecules may be considered as point scatterers, and there is no information on chain\ndimensions. In other words, we may call this the Rayleigh regime, and Equation 8.4.24a and\nEquation 8.4.24b apply.\n2.\nIf n < 1, then only the next term in the power series matters: P(q)~\nW 1 -\u2014 (q2/3)R2. In this\nregime one can obtain the value of Rg without any knowledge about the shape of the\nmolecule. A plot of 1/1ex versus 4]2 should be linear, with slope R2/3 Note that the intensity\nneed not be calibrated to obtain Rg; the units of the intercept cancel out when determining\nthe slope. Alternatively, P(q)~~ exp(\u2014- q2/3R2), and a plot of ln Iex versus i]2 should be linear\nwith slope -\u2014R2/3 This latter format1s termed a Guinier plot, and this regime the Guinier\nregime [5].\n3.\nIf 1 g n _<_ 10, more terms in the power series expansion of P(q) become important, and these\ndepend on the specific shape of the molecule. Accordingly, in this regime the mathematical\nform of P(q) is helpful in distinguishing different molecular shapes. We will identify some\nspecific functional forms below.\n4.\nIf n >> 1, the scattering is dominated by the internal structure of the molecule, and one can\nextract no information about Rg.\n"]], ["block_5", ["A more physical appreciation of these four regimes can be gained from the following viewpoint.\nRecalling the discussion in Section 8.2, the scattering vector has dimensions of inverse length, and\n"]], ["block_6", ["Solution\n"]], ["block_7", ["We continue this chapter with some further discussion of form factors. We return to the expansion\nof Equation 8.5.13, where P(q) is a power series in even powers of gig-k. We can delineate four\ngeneral regimes of behavior, depending on the approximate magnitude of n.\n"]], ["block_8", ["Example 8.4\n"]], ["block_9", ["312\nLight Scattering by Polymer Solutions\n"]], ["block_10", ["8.6\nScattering Regimes and Particular Form Factors\n"]]], "page_326": [["block_0", [{"image_0": "326_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "326_1.png", "coords": [22, 44, 285, 143], "fig_type": "figure"}]], ["block_2", ["R3, to the scattering length scale, I/q.\n"]], ["block_3", ["Therefore for M 104, n ranges from 0.0079x2.7=0.021 to 0.035x2.7=0.095. This falls\nentirely into the Rayleigh regime, and we would not be able to determine R3. For M 105, n\nranges from 0.067 to 0.30. This extends into the Guinier regime, and with precise measurements Rg\ncould be determined. Finally, for M 106, n ranges from 0.21 to 0.95. Although this appears to\nbe in the Guinier regime, in fact the next term in the expansion of P(q) may contribute at larger\nangles, so the lower angle data should be emphasized in the determination of R3.\nThere are two important conclusions to draw from these numbers. First, one can only change\nq by a factor of about 5 in a typical instrument. Second, for \ufb02exible polymers and typical molecular\nweights it is hard to get much beyond the Guinier regime. To access structural details at smaller\nlength scales, it is necessary to increase q substantially. This is done by utilizing x-ray or neutron\nscattering, where the wavelength is only a few angstroms; recall q 1/)\\.\n"]], ["block_4", ["Scattering \n313\n"]], ["block_5", ["slightly different wavelength and temperature, but we are only estimating here). Assuming we have\n"]], ["block_6", ["an argon laser (A0 488 nm), then\n"]], ["block_7", ["The form factors associated with particular distribution functions, P(rjk). have been derived for\na variety of shapes. We will just state the results for three particularly important ones: the Gaussian\ncoil, the rigid rod, and the hard sphere.\n"]], ["block_8", ["Figure 8.10\nIllustration of four regimes of scattering behavior, depending on the ratio of the polymer size,\n"]], ["block_9", ["1.\nFor the Gaussian coil, the form factor is known as the Debye function after it was \ufb01rst\ndeveloped by Debye [6]. It may be written as\n"]], ["block_10", ["231'\n\u20185?\n"]], ["block_11", ["q\n - \n.\nrd\n"]], ["block_12", ["4\n.\n. 24\n2\nX 3 14 X\n1 4\nsin (\u20145) 0.0079 nm\u20181\nqmi\u201c \n488\n2\n4 x 3.14 x 1.424\n.\n150\n_,\nQmax \n488\n3111(7) 0.035 nm\n"]], ["block_13", [{"image_2": "326_2.png", "coords": [48, 151, 288, 254], "fig_type": "figure"}]], ["block_14", ["2\nP(q) = Foe-x\n\u2014 1 +x),\nx a q:\n(8.6.1)\n"]], ["block_15", ["This function applies to chains in a theta solvent, and in the melt. It is important to realize that\nit no longer applies when q\u2018l becomes comparable to the persistence length or statistical\n"]], ["block_16", [{"image_3": "326_3.png", "coords": [52, 153, 162, 257], "fig_type": "molecule"}]], ["block_17", [{"image_4": "326_4.png", "coords": [53, 151, 163, 257], "fig_type": "figure"}]], ["block_18", [".\n/\nmy)\nI\nk)\n(I\n\\\n\\qj\n/\n1q s10\nq>>1\n"]], ["block_19", ["qg \u00ab1\nqt? S1\n"]]], "page_327": [["block_0", [{"image_0": "327_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "327_1.png", "coords": [33, 386, 288, 629], "fig_type": "figure"}]], ["block_2", ["Figure 8.11\nForm factors P(q) for Gaussian coils, hard spheres, and very thin rods (3) as a function\nof (mg)?\n(continued)\n"]], ["block_3", ["The technique of light scattering was first applied to polymer solutions in the 1940s. Early\ninstruments were homemade, and limited by the quality of the available components, such as\nphotodetectors. Two early commercial instruments, the Bryce-Phoenix and the So\ufb01ca, remained in\n"]], ["block_4", ["314\nLight Scattering by Polymer Solutions\n"]], ["block_5", ["segment length; in that regime the chain conformation no longer follows the Gaussian\ndistribution.\n2.\nFor a rigid rod of length L and zero width, the form factor is [7]\n"]], ["block_6", ["where the definite integral is tabulated in most mathematical handbooks.\n3.\nFor a hard sphere of radius R the result (also due to Lord Rayleigh [8]) is\n"]], ["block_7", ["8.7\nExperimental Aspects of Light Scattering\n"]], ["block_8", ["(a)\n(wig)2\n"]], ["block_9", ["1-2lil'llllTllllllllll\n"]], ["block_10", [{"image_2": "327_2.png", "coords": [51, 102, 232, 137], "fig_type": "molecule"}]], ["block_11", ["1/P(q) is plotted versus q\ufb01, i.e., in the form anticipated by the Zirnm plot. In this format all three\nhave the same small q limiting slopes, as they must, but they diverge beyond n a: 1. Note that\nRg :L/m for the rod, and x / 3/5R for the sphere (see Table 6.3), so the different definitions of\nx in Equation 8.6.1 through Equation 8.6.3 need to be handled carefully. In Figure 8.11c, P(q) is\nplotted on a logarithmic axis against n; note the oscillations in P(q) for a sphere. It is clear that\nall three P(q) look similar forn < 1, as expected; this corresponds to regions 1 (\u201cRayleigh\u201d) and\n2 (\u201cGuinier\u201d) above. Beyond that they begin to diverge, corresponding to regimes 3 and 4.\n"]], ["block_12", ["3\n2\n,\n2\nP(q) \n\u20143\n(smx xcosx)\n,\nx E qR\n(8.6.3)\nx\n"]], ["block_13", ["1\nsrnz\nsinx\n2\nqL\nP(q) J\u2014 dz (\u2014>\n,\nx E \n(8.62)\nx\nx\n2\n"]], ["block_14", ["All three P(q) expressions are plotted in Figure 8.11a as a function of (n)2. In Figure 8.11b,\n"]], ["block_15", ["O\nN\n43-\n0)...\n('30\nS\n"]], ["block_16", ["\u2014_\n"]], ["block_17", [".\u2014.\n"]], ["block_18", ["\u2014..\u2014\n"]], ["block_19", ["\u2014-\n"]], ["block_20", ["u\u2014_\n"]], ["block_21", ["\u2014-\n"]], ["block_22", ["\u2014\u2014\n"]], ["block_23", ["I|I_\u2019_LI_IJ_I_|_]_Li_111i11[iI'II\n"]]], "page_328": [["block_0", [{"image_0": "328_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["service in some laboratories for decades. Even with the advent of lasers, high-quality photomulti-\npliers, and sensitive photodiodes, light scattering was largely conducted on custom instrumentation,\nwhich limited its applicability. Two trends in the last two decades have reinvigorated the field,\nhowever. One was the emergence of dynamic light scattering (DLS, see Section 9.5) as an important\ncharacterization technique for polymer, biopolymer, and colloidal solutions. A DLS instrument can\nbe used to measure [(q) as well, and therefore commercial DLS apparatuses now often serve a dual\nrole. The second trend was the emergence of light scattering as an absolute molecular weight\ndetector for size-exclusion chromatography (SEC, see Section 9.7). In this context light scattering\nis finally approaching the status of a \u201croutine\u201d characterization tool for polymer solutions.\n"]], ["block_2", ["Figure 8.11 (continued)\n(b) Inverse form factors as in (a), showing the convergence to a common slope\nof 1/3 in the small n limit. (c) Form factors plotted on a logarithmic scale, showing the oscillations for the\nhard sphere.\n"]], ["block_3", ["(b)\n(CIFI\u2019g)2\n"]], ["block_4", ["(C)\nqFI'g\n"]], ["block_5", ["Experimental of Light Scattering\n315\n"]], ["block_6", ["COII\n\u2019\u2014\n5:\nL.\nI\n._\n"]], ["block_7", ["3\n\u2019\u2014\nI \n_\n4\n-\u2014\u2014\n_\n"]], ["block_8", ["\u20196'-\n_\n_\n"]], ["block_9", ["a: 10 \nE\u2014\n\\\nCo\u201c\n\u2014E\n"]], ["block_10", ["E\n[\u2014\u2014\n/\n"]], ["block_11", [{"image_1": "328_1.png", "coords": [38, 253, 286, 495], "fig_type": "figure"}]], ["block_12", [{"image_2": "328_2.png", "coords": [43, 31, 280, 260], "fig_type": "figure"}]], ["block_13", ["10\nE\nSphere\n\\/ \n_E\n"]], ["block_14", ["10\u20144\n"]], ["block_15", ["10\u20141\n2\u2014\nN \\\n\u20185\nE\n\\\n\u2018 x \nE\nr\n\u2018~~_:\n"]], ["block_16", ["10\n"]], ["block_17", ["100\nl\nI\nl\nI\nI\nI\n|\nIT\n1\u2014 rI\u2014I\n\u2014r T I\u2014I\u2014r\n"]], ["block_18", ["O\n"]], ["block_19", ["6\n_\n"]], ["block_20", ["2 t-\n_\n"]], ["block_21", ["8 l\u2014\n__\n"]], ["block_22", ["_3\n,\u2014\n"]], ["block_23", ["0\n2\n4\n6\n8\n10\n"]], ["block_24", ["E\nI\n"]], ["block_25", ["L\u2014\n_\n"]], ["block_26", ["l-\n_I\n"]], ["block_27", ["I\"\n_\n"]], ["block_28", ["L.\n_.\n"]], ["block_29", ["l_\n_\n"]], ["block_30", ["\u2019\u2014\n"]], ["block_31", ["E\nSphere\n/\n\u2014\n"]], ["block_32", ["0\n2\n4\n6\n8\n10\n"]], ["block_33", ["E\nf \\\nE\n"]], ["block_34", ["E\nE\nI\n\\ f\u2014\n"]], ["block_35", ["I\nI\nl\nI\n|\nI\nI _L _I_ l_l_I\n_I\nL\nl\nI\nI\nI\nI\n"]], ["block_36", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_37", ["I\nI\n1\n\u2014r T 7\u2014I\nl\nI\nI7\nI\nl\n_l\nT\nI_\nI\nl\nl\n"]], ["block_38", ["llllllI\n/ /.\n"]], ["block_39", [".\n_\n"]], ["block_40", ["\u2014\n"]], ["block_41", ["l\nI\n"]]], "page_329": [["block_0", [{"image_0": "329_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "329_1.png", "coords": [19, 474, 339, 630], "fig_type": "figure"}]], ["block_2", ["I\nSource\n:\nin,\n"]], ["block_3", ["where nglass is typically about 1.5 and nah 1.0. This gives a re\ufb02ectivity of about 4% per air\u2014glass\ninterface. The larger the diameter of the sample cell, the further this source of stray light will be\nfrom the scattering volume, and therefore the easier it is to eliminate. Empirically, for a\n1 cm\nsample cell it is very difficult to get to scattering angles below about 25\u00b0 because of this problem.\nNote that there will also be two glass\u2014sample solution interfaces, where for some solutions\n(aqueous ones, for example) there will also be substantial re\ufb02ection.\nThe re\ufb02ection problem is often mitigated, but not completely eliminated, by use of an index\u2014\nmatching \ufb02uid. In this case the sample cell is suspended in a much larger container filled with a\n\ufb02uid of refractive index near that of the glass (silicon oil and toluene are two common examples).\nBy this expedient the main air-glass interfaces can be moved several centimeters from the\nscattering volume. Furthermore, the index-matching bath can provide an excellent heat-transfer\nmedium for controlling the sample temperature. Good temperature stability is important for precise\nmeasurements.\nThe scattered light is collected by some combination of lenses and apertures, and directed onto\nthe active surface of the detector. In routine applications it is usually assumed that the scattered\n"]], ["block_4", ["The basic components of a light scattering instrument are illustrated in Figure 8.12. The typical\nlight source is a laser, although the unique features of a laser source (e.g., temporal and spatial\ncoherence, collimation, high monochromaticity) are not required. The most desirable feature of\nthe source is stability. The source may be vertically polarized or unpolarized, as discussed in\nSection 8.3, but it is important to establish that you have one or the other, and not a partially\npolarized beam. The beam should be collimated before traversing the sample cell. The solution\nis usually contained in a glass cell, and the glass should be carefully selected for clarity and lack\nof imperfections. All other things being equal, the larger the diameter of the sample cell, the\nbetter. The reason is that at each air\u2014glass interface a significant portion of the incident light\nwill be reflected, rather than transmitted. This re\ufb02ected light may find its way to the detector as\nstray light, or back into the scattering volume at a different angle, leading to erroneous\nscattering signals. The re\ufb02ected fraction of the intensity, R, for near\u2014normal incidence is\ngiven roughly by\n"]], ["block_5", ["Figure 8.12\nSchematic diagram of a light scattering photometer.\n"]], ["block_6", ["8.7.1\nInstrumentation\n"]], ["block_7", ["316\nLight Scattering by Polymer Solutions\n"]], ["block_8", ["mM\n(8.7.1)\n(\u201dglass + \u201dair)\n"]], ["block_9", ["Io monitor\n"]], ["block_10", ["I\n\u2018\n"]], ["block_11", ["I\nI, monitor\n"]], ["block_12", ["43/\n/\nI5 detector\n"]]], "page_330": [["block_0", [{"image_0": "330_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["used \nsome situations, such examining rod-like particles, signi\ufb01cant depolarization of the\nscattered polarizer front of detector becomes necessity. Three\ngeneral detection schemes can be envisioned. The simplest is to have a single detector at some\nfixed of \ninformation, detector \nprocess changes in time. The next option is to have a single detector that\ncan rotate which runs through the center of the sample cell. This provides a\ncontinuously variable 9, which is very desirable. In recent years, it has become popular to use a\nmovable Optical collect the light, and keep the actual detector fixed in space. In this\nconfiguration, usuallya photomultiplier, which can be extremely sensitive (capable\nof counting if need be) and which has a wide linear range (output current\nproportional to incident intensity over many orders of magnitude). The third approach is to arrange\nmultiple detectors at fixed angles, which allows for simultaneous detection with the attendant\nincrease in overall signal-to-noise (the \u201cmultiplex advantage\u201d). This technique has only recently\nbecome practical with improvements in photodiode detectors, such that small, inexpensive units\nare sufficiently sensitive and linear. The multiangle light scattering technique is now the basis of a\npopular SEC detector, which can give real-time information about M and Rg for each slice of the\nchromatogram (see Section 9.7).\nThe final instrumental issue is the use of an incident intensity monitor, to follow \ufb02uctuations\nand drift in the source output. Often a small portion of the incident beam is split off, for example\nby inserting a piece of glass in the incident beam at 45\u00b0 to the prOpagation direction, and directed\nto a separate detector. This permits the scattered signal to be normalized to the incident intensity\nin real time, which ultimately leads to a much more reliable R9. For example, when measuring a\ndilute polymer solution the scattered intensity may be less than twice that of the solvent alone.\nThe solvent and solution would be measured at different times, so small but uncontrolled\nvariations in source intensity can severely compromise the reliability of the excess intensity. It\nis also desirable to place a detector to monitor the transmitted beam. This could serve as an\nincident intensity monitor, but by having detectors both prior to and after the sample it is possible\nto detect changes in the total sample scattering, and thereby to assess whether absorption or\nmultiple scattering is a problem.\n"]], ["block_2", ["Referring back to Equation 8.4.22, we see that in order to extract a value of MW, we will need to\nknow the refractive index of the solvent, art/60, 0, A0, and r. All of these are straightforward to\ndetermine, except r. In addition, and most importantly, the Rayleigh ratio involves the ratio of\nthe scattered intensity per unit volume to the incident intensity, and as we noted in the beginning\nof the chapter, we rarely determine a true intensity. Furthermore, we would not know the active\narea of the photodetector, or its quantum efficiency (the fraction of incident photons actually\ndetected).\nThe resolution of these difficulties is actually rather simple. We assume that the detector\nproduces an output signal S (current, voltage, or \u201ccounts\u201d) proportional to the incident intensity:\n"]], ["block_3", ["where the proportionality factor B is independent of the magnitude of S over the relevant range\n(i.e., linear response), and the subscripts denotes \u201cscattering.\u201d As long as we maintain the detector at\na fixed distance r from the scattering volume as the scattering angle is varied, then the detector\n"]], ["block_4", ["8.7.2\nCalibration\n"]], ["block_5", ["light is entirely polarized vertically (for vertically polarized incident light), and a polarizer is not\n"]], ["block_6", ["Experimental Aspects of Light Scattering\n317\n"]], ["block_7", ["SS BSIS\n(8.12)\n"]]], "page_331": [["block_0", [{"image_0": "331_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "331_1.png", "coords": [18, 484, 256, 575], "fig_type": "figure"}]], ["block_2", [{"image_2": "331_2.png", "coords": [18, 465, 282, 649], "fig_type": "figure"}]], ["block_3", ["should collect light over the same range of solid angle fraction at each 6. A similar equation\ndescribes the response of the incident intensity monitor:\n"]], ["block_4", ["We are now in a position to obtain the Rayleigh ratio (units of cm\u201d) as follows:\n"]], ["block_5", ["where there is a single unknown proportionality factor 7. This we obtain by measuring a pure\nsolvent for which the absolute Rayleigh ratio has been measured (by someone who took a lot of\ncare) at the same wavelength and temperature. A second calibration step is required for instruments\nwith multiple detectors, as each detector will have its own value of B (and also a different r and\ncollection solid angle). This is best accomplished by using a solution with relatively high scattering,\nbut one for which the scattering is entirely incoherent. A solution of a moderate molecular weight\npolymer can serve this purpose. Then the signal from each detector can be normalized to a single\nreference detector (usually chosen at 6 90\u00b0) by a multiplicative factor. It should be noted that\nEquation 8.7.2 and Equation 8.7.3 assume that there is no background signal (i.e., a finite 5 when\nthe source is turned off) or stray light (i.e., contributions to S not from sample scattering at angle 6).\nIn fact, all photodetectors have some dark current (output in the absence of input), but this is very\nsmall. (If it is a significant fraction of the scattering signal, then your experiment is in trouble.)\nOther sources of background (e.g., room light) and stray light can be minimized by appropriate\ninstrumental design. However, the simple fact is that the detector is indiscriminate; all photons that\nreach it will be counted as scattering, no matter what their origins, and so great care must be taken to\neliminate background and stray light.\nThere is an additional correction step that must be performed when using a single detector that\nrotates to various 9. Refen\u2018ing to Figure 8.13, the scattering volume is determined by the\nintersection of the incident beam and the collected beam (the latter being set by the geometry of\nthe detection system). Assuming that both are square in cross section, and that the latter is\ncomparable in diameter to the former, we can see that as 6 deviates from 90\u00b0 the volume of\n"]], ["block_6", ["318\nLight Scattering by Polymer Solutions\n"]], ["block_7", ["Figure 8.13\nIllustration of the change in scattering volume with scattering angle, 6, for incident and\nscattered beams with width d.\n"]], ["block_8", ["So 3010\n(8.13)\n"]], ["block_9", ["R9:\n"]], ["block_10", ["Vol d3/sin a\n"]], ["block_11", [": \n__ Siolvent)\n(874)\nSo\n"]], ["block_12", ["\u2014 \n3050\n"]], ["block_13", ["rZUSsolution Igolvent)\nIo\n"]], ["block_14", ["\u20184\n4\n"]], ["block_15", ["d\n"]]], "page_332": [["block_0", [{"image_0": "332_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Usually one is interested in concentrations on the order of 10\u20182 10\u20144 g/mL, and arr/ac is of\norder 0.1, so it is not simply a matter of measuring n directly, because the change in n will only\nappear in the third or fourth place after the decimal. Rather, a di\ufb01ferenrial refractometer should be\nused. A variety of designs are available, either commercially or by relatively simple assembly in\nthe laboratory, but the principle behind the simplest version is illustrated in Figure 8.14. The light\n"]], ["block_2", ["As is evident from the preceding discussion, (Tin/8c\u201c plays a central role in the magnitude of the\nscattered intensity. It may be defined as follows:\n"]], ["block_3", ["a refractive index detector serves as a direct concentration monitor. There are two standard\ndiagnostics for the presence of dust in the sample. The first is to examine the temporal \ufb02uctuations\nin 53. These should be random, and have a root\u2014mean\u2014square amplitude close to x/ (55). If some\ndust is present, the signal is likely to increase suddenly and then decrease suddenly some seconds\nlater, as dust particles drift in and out of the scattering volume. The scattering volume can also be\nexamined visually; the tell-tale bright \ufb02ashes of dust are often easily seen. The second diagnostic is\nto examine 1/109) versus sin2(6/2) (or qz). According to the Zimm equation this should produce a\nstraight line. Dust will increase 1(6) selectively at low 6, resulting in a characteristic downturn in\nthe plot. (Note that this test should only involve the range of 6 deemed to be free of re\ufb02ections.)\n"]], ["block_4", ["where n is the refractive index of the solution, it, the refractive index of the pure solvent, and c is\nthe concentration in g/mL. The initial slope of n(c) versus 6 gives Bit/ac, but is better obtained as\n"]], ["block_5", ["There are two main issues here. First is the choice of solvent, and second is the preparation of\n\u201cdust\u2014free\u201d samples. One may not have the freedom to choose the solvent, but all other things\nbeing equal it is desirable to have |8nl6c| as large as possible, as the intensity is proportional to\n(an/36f. It may also be helpful to choose a solvent with a relatively small R9 of its own, so that the\npolymer contribution to the excess scattering is larger. (You might be wondering \u201cWhy does the\npure solvent scatter at all, given that we emphasized that homogeneous materials do not scatter?\u201d\nThe answer is density \ufb02uctuations, and the scattered intensity is determined by the isothermal\ncompressibility of the solvent, K, in parallel to Equation 8.4.11, the intensity is proportional to\nkTK). Finally, some solvents are easier to make dust-free than others; for example, more polar\nsolvents such as water and THF are often trickier to clean than toluene or cyclohexane. The\npreparation of dust-free samples takes some care and experience. It is essential to remove dust, as\nstray particles that are significantly larger than the polymer molecules will scatter strongly. The\ntwo standard options are filtration and centrifugation, and the fonner is usually preferred. Both are\nless than ideal, in that they may change the concentration of the solution. The use of light scattering\n"]], ["block_6", ["a functioning is to plot 53(6) sin 6 versus 9 for an incoherent scattering solution. This\nproduct should be independent of 6. Ultimately, at high and low 9 the data will deviate from the\nconstant due to re\ufb02ections at the various interfaces discussed above. This measurement will\ntherefore provide a guide as to the reliable range of 6 for the instrument.\n"]], ["block_7", [{"image_1": "332_1.png", "coords": [34, 574, 192, 607], "fig_type": "molecule"}]], ["block_8", ["intersection increases by a factor of l/sin 6. Thus, the signal at any angle should be multiplied by a\nfactor of sin 6 in order to account for this variation in scattering volume. A very good indication of\n"]], ["block_9", ["8.7.4\nRefractive Index Increment\n"]], ["block_10", ["as an SEC detector has a particular advantage here in that the column acts as an excellent filter, and\n"]], ["block_11", ["8.7.3\nSamples and Solutions\n"]], ["block_12", ["Experimental Aspects of Light Scattering\n319\n"]], ["block_13", ["an.\n,\nn n5\n__\n_\nAn,\nE 3313\u201c\nc\n) \u2018l11.%(7)\n(8.7.6)\n"]], ["block_14", ["nzns+\nc+acz+~~\n(8.7.5)\n"]], ["block_15", ["9:\u201d:\n"]], ["block_16", ["BC\n"]]], "page_333": [["block_0", [{"image_0": "333_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "333_1.png", "coords": [33, 23, 280, 234], "fig_type": "figure"}]], ["block_2", ["beam is incident on a divided cell that contains the solvent in one compartment and the solution in\nthe other (or solutions of two different concentrations). The dividing glass surface is angled, so that\nthe beam direction in compartment 2 depends on the refractive index of the \ufb02uid in compartment 2,\nby Snell\u2019s law. The beam is then re\ufb02ected back through both compartments, with a final direction\nthat depends on An. For small An it can be shown that the displacement of the re\ufb02ected beam, d, is\nproportional to An.\nThe value of an/ac depends on T and )t, so it should be measured under the same conditions as the\nlight scattering itself. It is independent of molecular weight for large M, but will develop a significant\nM dependence at low M, so this may need to be taken into account in, for example, SEC detection (see\nSection 9.7). The main source of the M dependence of an/ac lies in the M dependence of the polymer\nrefractive index np, which in tum depends directly on density. The M dependence of the density re\ufb02ects\nprimarily the concentration ofchain ends, and art/ac is therefore usually linear in 1/M,,. It is tempting to\ntry and relate {in/36 to n], n, in some simple way, but this is risky. There is no general, rigorous\nexpression for n(c) that depends only on n3, np, and c, and in fact n need not even vary monotonically\nwith 6. However, it is usually the case that art/ac will increase with the magnitude of up n3, so this\nbecomes a reasonable starting point when choosing a solvent to maximize an/ac. Note that an/ac can\nbe positive, negative, or zero; as it appears squared in the scattered intensity, the sign does not matter. If\none chooses an isorefractive solvent, such that (an/ac z 0, that polymer will be invisible in the\nscattering experiment. This can be used to advantage in examining polymer mixtures and copolymers,\nin order to highlight the behavior of one particular component.\n"]], ["block_3", ["In this chapter we have developed in some detail the equations governing the scattering of light\nfrom dilute polymer solutions. Light scattering is a very powerful experimental tool, providing\nboth thermodynamic and structural information. Recent developments in commercial instrumen-\ntation have accelerated the use of light scattering in routine characterization. The main concepts in\nthe development are the following:\n"]], ["block_4", ["1.\nCompletely uniform materials do not scatter. Scattering in polymer solutions arises from\nrandom \ufb02uctuations in concentration. This scattering is incoherent, which means that the\nintensity is independent of scattering angle.\n2.\nThe incoherent intensity is determined by the mean-square concentration \ufb02uctuations, which\nin turn are set by the ratio of the thermal driving force, H\", to the free energy penalty for a\ngiven \ufb02uctuation, (326/862. It is through this relationship that the scattering experiment\nprovides a measurement of Mw and B.\n"]], ["block_5", ["8.8\nChapter Summary\n"]], ["block_6", ["320\nLight Scattering by Polymer Solutions\n/\n._\n"]], ["block_7", ["Figure 8.14\nIllustration of a refractometer based on a split cell prism.\n"]], ["block_8", ["\u20185\n"]], ["block_9", ["d\n"]]], "page_334": [["block_0", [{"image_0": "334_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["*B.H. Zimm J. Chem. Phys. 16, 1093, 1099 (1943).\nI1.]. Blum and M.F. Morales Arch. Biochem. Biophys. 43, 208 (1953).\n"]], ["block_2", ["3.\nCarry though the integration of Equation 8.4.7 to obtain Equation 8.4.10.\n4.\nAn estimate of RE can be obtained rather simply from the so-called dissymmetry ratio, defined\n"]], ["block_3", ["Draw a plot in polar coordinates of the scattering envelope in the x y plane. How would the\nenvelope of a Rayleigh scatterer compare with this plot? By interpolation, evaluate the\ndissymmetry ratio and Rg. What are some practical and theoretical objections to this procedure\nfor estimating Rg?\n5.\nThe effect of adenosine triphosphate (ATP) on the muscle protein myosin was studied by light\nscattering in an attempt to resolve con\ufb02icting interpretations of viscosity and ultracentrifuge\ndata. The controversy hinged on whether the myosin dissociated or changed molecular shape\nby interaction with ATP. Blum and MoralesI reported the following values of (Kc/R6)C:0\nversus sin2(9/2) for myosin in 0.6 M Kc1 at pH 7.0. Which of the two models for the mode of\n"]], ["block_4", ["1.\nExplain in your own words (i) why the permeation of solvent through a membrane into a\npolymer-rich phase and the amount of light scattered by a polymer solution are related, (ii)\nwhy even though they are closely related, the osmotic pressure and light scattering experi\u2014\nments measure different moments of the molecular weight distribution, and (iii) why the\nosmotic pressure measurement becomes increasingly difficult for degrees of polymerization\n>103 while light scattering becomes harder for degrees of polymerization \u00a3103.\n2.\nEstimate the largest and smallest molecular weight polystyrenes for which one could reason\u2014\nably expect to measure Rg reliably in THF, using an Argon laser at 488 nm, and a usable\nangular range of 30\u00b0 < 6 < 150\u00b0; THF is a good solvent for polystyrene.\n"]], ["block_5", ["3.\nFor polymers that are large enough, a significant phase difference can arise between portions\nof the incident wave that are scattered from different monomers on the same chain. This leads\nto angle-dependent, or coherent, scattering. The underlying process is very similar to Bragg\ndiffraction from crystals, with the key difference being that in polymer solutions there are only\naverage correlations in the positions of the various monomers, rather than a permanent lattice.\n4.\nThe description of coherent scattering is built around the scattering wavevector, (f. The\nmagnitude of this vector depends on wavelength and scattering angle, and has units of inverse\nlength. Depending on the magnitude of the dimensionless product n, the coherent scattering\ncan give information about the internal structure or the overall size of the polymer. It is often\npossible to use light scattering to measure Rg in a completely model-independent way.\n"]], ["block_6", ["Problems\n321\n"]], ["block_7", ["Problems\n"]], ["block_8", ["as Iex(6=450)/Iex(6= 135\u00b0). Explain how this works. Zimm has reported the intensity of\nscattered light ()1 = 364 nm) at various angles of observation for polystyrene in toluene at a\nconcentration of 2><10_4 g/mLT. The following results were obtained (values marked were\nestimated and not measured):\n"]], ["block_9", ["0\n4.29\u201d\u201c\n25.8\n3.49\n36.9\n2.89\n53.0\n2.18\n66.4\n1.74\n90.0\n1.22\n113.6\n0.952\n143.1\n0.763\n180\n0.70*\n"]], ["block_10", ["9 (o)\nIS (a.u.)\n"]]], "page_335": [["block_0", [{"image_0": "335_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["IRE. Steiner and K. Laki, Arch. Biochem. Bt'ophys. 34, 24 (1951).\n1P. Outer, C.I. Carr, and 13.11. Zimm, J. Chem. Phys. 18, 839 (1950).\n"]], ["block_2", ["322\nLight Scattering by Polymer Solutions\n"]], ["block_3", [{"image_1": "335_1.png", "coords": [37, 313, 419, 438], "fig_type": "figure"}]], ["block_4", ["For poly(n-hexyl isocyanate) and poly(methyl methacrylate), estimate the range of molecular\nweights for which the Guinier regime (n < 1) can be accessed. The data in Figure 6.10 and\nTable 6.1 may be useful.\nAggregation of fibrinogen molecules is involved in the clotting of blood. To learn something\nabout the mechanism of this process, Steiner and LakiJr used light scattering to evaluate M and\nthe length of these rod\u2014shaped molecules as a function of time after a change from stable\nconditions. The stable molecule has a molecular weight of 540,000 g mol and a length of\n840 151. The accompanying table shows the average molecular weight and average length at\nseveral times for two different conditions of pH and ionic strength 11. Criticize or defend the\nfollowing proposition: The apparent degree of aggregation x at various times can be obtained in\nterms of either the molecular weight or length. The ratio of the value of x based on M to that\nbased on length equals unity exclusively for end-to-end aggregation and increases from unity as\nthe proportion of edge\u2014to\u2014edge aggregation increases. In the higher pH\u2014lower p. experiment\nthere is considerably less end\u2014to\u2014end aggregation in the early stages of the process than in the\nlower pH\u2014higher p. experiment.\n"]], ["block_5", ["Zimm plots at 546 nm were prepared for a particular polystyrene at two temperatures and in\nthree solvents. The following summarizes the various slopes and intercepts obtainedi:\n"]], ["block_6", ["Slope\nIntercept 9:0\nMethyl ethyl ketone\n0.608\n260\nDichloroethane\n1. 16\n900\nToluene\n1. 14\n1060\n"]], ["block_7", ["Solvent\nIntercept 6:0\nIntercept 9:0\nMethyl ethyl ketone\n0.551\n230\nDichloroethane\n1 .05\n870\nToluene\n1.09\n800\n"]], ["block_8", ["Solvent\n"]], ["block_9", ["ATP interaction with myosin do these data support? Explain your answer by quantitative\ninterpretation of the light\u2014scattering data.\n"]], ["block_10", ["T 67\u00b0C\n"]], ["block_11", ["1 150\n1.63\n1600\n1000\n2.0\n1200\n1670\n2.20\n1900\n2350\n3.30\n2200\n"]], ["block_12", ["sin2(6/2)\n0.15\n0.21\n0.29\n0.37\n0.50\n0.85\nKc/Rgx 107 (Before ATP)\n0.9\n1.1\n1.5\n1.8\n2.2\n2.7\nKc/R9x107 (With ATP)\n1.9\n2.8\n3.7\n4.6\n6.0\n6.8\n"]], ["block_13", ["T:22\u00b0C\n"]], ["block_14", ["1(3)\nMx 10 (g mol\n\u2014 1)\nLength (A)\nr (s)\nMx 10\n\u2014 6 (g mol\n\u2014 1)\nLength (A)\n"]], ["block_15", ["650\n1.10\n1300\n900\n1.10\n1100\n"]], ["block_16", [{"image_2": "335_2.png", "coords": [58, 471, 341, 556], "fig_type": "figure"}]], ["block_17", ["pH 8.40 and p. 0.35 M\npH 6.35 and p. 0.48 M\n"]], ["block_18", [{"image_3": "335_3.png", "coords": [91, 90, 381, 142], "fig_type": "molecule"}]], ["block_19", ["Intercept 6:0\n"]], ["block_20", ["Slope\n"]], ["block_21", ["Slope\nSlope\n"]]], "page_336": [["block_0", [{"image_0": "336_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n323\n"]], ["block_2", ["10.\n"]], ["block_3", ["11.\n"]], ["block_4", ["12.\n"]], ["block_5", ["13.\n"]], ["block_6", ["14.\n"]], ["block_7", [{"image_1": "336_1.png", "coords": [50, 325, 187, 457], "fig_type": "figure"}]], ["block_8", ["Evaluate MW, Rg, and B from each piece of pertinent data and comment on (a) the agreement\nbetween MW values and (b) the correlation of Rg and B with solvent quality.\nPlot the light scattered intensity (arbitrary units) versus scattering angle that you would\nexpect to see for a very dilute solution of polystyrene with M =4,000,000 in cyclohexane at\n35\u00b0C. Assume the instrument has an angular range of 20\u00b0\u2014150\u00b0. On the same axes show the\nangular dependence of the intensity if the scattering object were a hard sphere with the same Rg.\nFor polystyrene in butanone at 67\u00b0C the following values of Kc/Rgx 10\u00b0 were measured at the\nindicated concentrations and angles. Construct a Zimm plot from the data below using\n7: 100 mL/g for the graphing constant. Evaluate M, B, and R3 from the results. In this\nexperiment A0 546 nm and n 1.359 for butanone.\n"]], ["block_9", ["The slope\u2014intercept ratios have units of cubic centimeters per gram, and the intercepts are\nc/n, where the subscript v indicates vertically polarized light. The following values of n\nand 3n/3c can be used to evaluate K:\n"]], ["block_10", ["Methyl ethyl ketone\n1.378\n0.221\n1.359\n0.230\nDichloroethane\n1 .444\n0.158\n1 .423\n0. 167\nToluene\n1.496\n0. 108\n1.472\n0. 1 18\n"]], ["block_11", ["Draw a sketch of a complete Zimm plot for a high molecular weight polymer in a theta\nsolvent, assuming data were acquired at eight angles and at four concentrations. In a different\ncolor pen, indicate the effect on the data if a small amount of dust were present in the\nsolution. Similarly, in a different color indicate the effect on the data of raising the solution\ntemperature substantially. State any assumptions you make.\nWhen a dilute solution of block copolymers undergoes micellization, i.e., some numbers of\nchains aggregate into a (usually spherical) assembly to shield one block from the solvent it\ndoes not like, the light scattering intensity increases. In fact, if micellization is induced by\nchanging temperature at a fixed concentration, the ratio of the intensity after micellization to\nthat before micellization is a good estimate of the average aggregation number of the\nmicelles. Explain this observation.\nShow that the form factors for the Gaussian coil and the hard sphere do indeed reduce to the\nexpected 1\n\u2014 q\ufb01/3 at low q, using the appropriate series expansions.\nImagine that you perform light scattering measurements on a diblock copolymer, and you\ngenerate a Zimm plot in the standard way. However, one of the blocks is isorefractive with\nthe solvent (an/ac 0), so it does not contribute directly to the scattering signal. Comment\non how the apparent values of MW, Rg, and B (i.e., those obtained by assuming you are\n"]], ["block_12", ["25.8\n\u2014\n1.48\n36.9\n1.84\n1.50\n53.0\n1.93\n1.58\n66.4\n1.98\n1.62\n90.0\n2.10\n1.74\n113.6\n2.23\n1.87\n143.1\n2.34\n1.98\n"]], ["block_13", ["0 (\u00b0)\n1.9x 10'3\n3.8x 10'4\n"]], ["block_14", ["c (g/mL)\n"]], ["block_15", ["72\nan/ac\nn\nan/ac\n"]], ["block_16", ["T = 20\u00b0C\nT 67\u00b0C\n"]]], "page_337": [["block_0", [{"image_0": "337_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["324\n"]], ["block_2", ["15.\n"]], ["block_3", ["16.\n"]], ["block_4", ["17.\n"]], ["block_5", ["18.\n"]], ["block_6", ["19.\n"]], ["block_7", ["looking at a homopolymer) might relate to the values associated with the copolymer as a\nwhole, or those of the two blocks.\nImagine dissolving dilute quantities of two different polymers, A and B, in a solvent and\nmaking light scattering measurements. What is the appropriate equation to describe the\nresults, analogous to the Zimm equation? You may assume that the chains are sufficiently\nsmall that P(q) 1 for both A and B. You will need to consider two concentrations, CA and CB,\ntwo molecular weights MA and MB, two values of one important optical parameter, and more\nthan one second virial coef\ufb01cient. You do not need to do a complicated derivation to get the\nanswer, but it is necessary to consider how the light scattering signal responds to concentra-\ntion \ufb02uctuations, and how many different \u201ckinds\u201d of concentration \ufb02uctuations there are in a\nthree-component mixture.\nSuppose you had a new polymer, in which you knew the monomer structure but not the shape\nof the polymer in solution (e.g., rod, coil, spherical globule, etc.). A good light scattering\nmeasurement would give you MW, Rg, and B. However, from this information alone, you\ncould also infer the shape; explain how, perhaps with numerical estimates of the relevant\nparameters. Note that fitting to the form factor is not the answer; all form factors are the same\nin the small n limit.\nImagine making light scattering measurements on a statistical copolymer of styrene and\nmethyl methacrylate, in a very dilute solution, and at angles such that n << 1. The sample\nhas a mean styrene composition of f (0 c,\u2014, (an/80,,\n(where the subscript i allows for polydispersity in M), and the usual optical constants. In\nreality, the measured intensity may well be larger than this value; the reason lies in the fact\nthat each chain will have a composition that may differ from f. Making reasonable\nassumptions about the M andf dependences of (an/86),, the refractive index increment of\nchain i, develop an expression for the observed intensity, in terms of averages involving 8f,-,\nwhere 5f, f, f, the difference in composition between chain\n1' and the sample average.\nYour result should show the interesting (and correct) prediction that there can be excess\nscattered intensity from a statistical copolymer solution, even when a solvent is chosen\nsuch that (an/8c) 0.\nAn analytical expression for the form factor of a Kratky-Porod worm\u2014like chain (WLC,\nChapter 6.4.2) is not available (although numerical approximations valid in various regimes\nhave been developed). Nevertheless, with a little thought you should be able to make a\nreliable sketch of what P(q) must look like. Take a WLC with L 100 nm and 6p 10 nm.\nUsing a reasonable plotting program, generate a plot of q2P(q) versus q (a so\u2014called \u201cKratky\u201d\nplot) for a Gaussian coil with the same Rg as this WLC. Run the q axis from 0 out to 0.5\naI. Do the same thing for a rigid rid with L 100 nm, and plot it on the same axes. Finally,\nby considering the low and high q asymptotic behavior of the various structures, draw by\nhand a smooth curve representing the WLC.\nThere are often differences in practice between various analysis schemes that are otherwise\nequivalent in principle. For example, use a computer to generate P(q) \u201cdata\u201d for a high\nmolecular weight Gaussian coil from n 0 out to n 2. Add \u201cnoise\u201d to the data with a\nreasonable amplitude using a random number generator. Then, fit the data to a straight line\naccording to the Zimm approach (1/P(q) versus (n)2) and to the Guinier approach (1n\nP(q) versus (n)2. Vary the n range of the data included in the fit at the high n end\n(why is this appropriate?) Which fitting approach is more robust, in terms of returning the\ncorrect Rg value?\n"]], ["block_8", ["Light Scattering by Polymer Solutions\n"]]], "page_338": [["block_0", [{"image_0": "338_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Brown, W. (Ed.), Light Scattering: Principles and Development, Clarendon Press, Oxford, 1996.\nHuglin, M.B. (Ed.), Light Scattering from Polymer Solutions, Academic Press, New York, 1972.\nJohnson, C.S. Jr. and Gabriel, D.A., Laser Light Scattering, Dover Publications, New York, 1994.\nTanford, C., Physical Chemistry of Macromolecules, Wiley, New York, 1961.\nYamakawa, H., Modern Theory of Polymer Solutions, Harper & Row, New York, 1971.\nWyatt, P.J., Analytica Chimica Acta, 272, 1, 1993.\n"]], ["block_2", ["Brandrup, J. and Immergut, E.H. (Eds), Polymer Handbook, 3rd ed., Wiley, New York, 1989.\nStrutt, J.W. (Lord Rayleigh), Phil. Mag. 41, 177, 447 (1871).\nMiller, T.H., CRC Handbook ofChemistry and Physics, 8 1 st ed., pp. 10\u2014160, CRC Press, Cleveland, 2001.\nZimm, B.H., J. Chem. Phys. 16, 1099 (1948).\nGuinier, A., Ann. Phys. 12, 161 (1939).\nDebye, P., J. Appl. Phys. 15, 338 (1944).\nNeugebauer, T., Ann. Physik 42, 509 (1943).\nLord Rayleigh, Proc. Roy. Soc. A84, 25 (1910).\nPOTJQ\u2018P\u2018PP\u2019PT\u2018\n"]], ["block_3", ["Further Readings\n325\n"]], ["block_4", ["References\n"]], ["block_5", ["Further Readings\n"]]], "page_339": [["block_0", [{"image_0": "339_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This chapter contains one of the more diverse assortments of topics of any chapter in this volume.\nIn this chapter we discuss the viscosity of polymer solutions, the diffusion of polymer molecules,\nthe technique of dynamic light scattering, the phenomenon of hydrodynamic interaction, and the\nseparation and analysis of polymers by size exclusion chromatography (SEC). At first glance these\nseem to be rather unrelated topics, but all are important to molecular weight determination in\nsolution. Furthermore, all share a crucial dependence on the spatial extent of the molecules. In\nChapter 8 we considered in detail how light scattering can provide a direct measurement of the\nradius of gyration. In this chapter the measure of size turns out to be roughly proportional to, but\nnot numerically equal to, the radius of gyration. As the chapter heading suggests, we now consider\nfor the first time in the book the time-dependent properties of polymers and particularly the rate at\nwhich polymer molecules move through a solvent. By emphasizing dilute solutions, the properties\nof individual polymer molecules are highlighted; in Chapter 11 we will consider the dynamic\nproperties of more concentrated solutions and melts.\nA parameter that plays an important role in unifying the concepts of viscosity and diffusion is the\nfriction factor. We will initially define the moiecular friction factor, f, by a thought experiment.\nImagine a polymer molecule dissolved in a solvent. Further imagine that we have a way of pulling\nthis molecule gently, but persistently, in one direction. After an initial start-up period, or transient\nresponse, we would find that if we apply a constant force I? the molecule would move with a constant\nvelocity \u20187 in the direction of the force. The proportionality factor between the applied force and the\nresulting velocity isf = |13|/|v\u2019| The units offare therefore (g cm/s2)/(cm/s) = g/s in the cgs system\n(the SI unit is kg/s), and a typical value for a polymer dissolved in water might fall between 10\u20187 and\n10\u201c5 g/s. The next few sections are concerned with two general questions: how can we gain\nexperimental access to f, and what canf tell us about the polymers themselves? Our starting point\nwill be to consider a single hard sphere. Although this idealization might appear at first to resemble\nthe ideal gas or the ideal solution in thermodynamics\u2014a simple model to use in learning the r0pes,\nbut of limited practical relevancewit is not so. For a rather profound reason, which we shall explore\nin Section 9.6, even \ufb02oppy random coils have friction factors very similar to those of hard spheres of\ncomparable size. However, we will get to that in due course; let us start at the beginning. Before\ndiscussing hard spheres, we need to introduce the viscosity of a \ufb02uid in more concrete terms.\nAs a place to begin, we visualize the \ufb02uid as a set of infinitesimally thin layers moving parallel\nto each other, each with a characteristic velocity. In addition, we stipulate that those \ufb02uid\nlayers that are adjacent to non\ufb02owing surfaces have the same velocity as the rigid surface. This\nis another way of stating that there is no slip at the interface between the stationary and \ufb02owing\nphases, which is a good approximation for the slow \ufb02ows of immediate interest. Now suppose we\nconsider a sample of \ufb02uid that is maintained at constant temperature, sandwiched between two\nrigid parallel plates of area A as shown in Figure 9.1. If a force F in the x\u2014direction is applied to the\ntop plate that plate and the layer of \ufb02uid adjacent to it will accelerate until a steady x\u2014velocity is\nreached (although force and velocity are vectors, we will drop the arrows for the remainder of\nthis chapter). As long as the deforming force continues to be applied, this velocity is unchanged.\n"]], ["block_2", ["9.1\nIntroduction: Friction and Viscosity\n"]], ["block_3", ["Dynamics of Dilute Polymer Solutions\n"]], ["block_4", ["9\n"]], ["block_5", ["327\n"]]], "page_340": [["block_0", [{"image_0": "340_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "340_1.png", "coords": [21, 542, 149, 578], "fig_type": "molecule"}]], ["block_2", [{"image_2": "340_2.png", "coords": [34, 48, 236, 155], "fig_type": "figure"}]], ["block_3", ["This time\u2014independent behavior is called steady \ufb02ow and will be our primary concern. During the\nacceleration that precedes the stationary state, the velocity is a function of time. For our purposes,\nwe shall simply wait until the stationary state is reached and not even question how long it will\ntake. Force per unit area is called stress, and in shear \ufb02ow is given the symbol 0'.\nIn the experiment described in Figure 9.1, the bottom plate remains in place and the nonslip\ncondition stipulated above requires that the layer of \ufb02uid adjacent to the bottom plate also has zero\nvelocity. This situation clearly requires that the velocity of the \ufb02uid varies from layer to layer\nacross the gap between the two rigid plates. To formulate this mathematically, we write that the top\nand bottom of these imaginary \ufb02uid layers are separated by a distance Ay and differ from each\nother in velocity by Av. The ratio Av/Ay has units of reciprocal time and is called either the\nvelocity gradient or the rate of shear. The former name is self-explanatory and the latter may be\nunderstood by considering the actual deformation the sample undergoes under the shearing force.\nDuring a short time interval At, the top layer moves a distance Ax relative to the bottom layer.\nAccordingly, Av may be written as Av :Ax/At and the velocity gradient may be expressed as\nAv/Ay (Ax/At)/Ay (Ax/Ay)/At. The shear displacement Ax divided by the distance over which\nit vanishes to zero, in this case Ay, is called the shear strain, which we represent by 7/. These\nrelationships show that the ratio Av/Ay also describes the rate at which the shear strain develops,\nor, more simply, the rate of shear A'y/At, or 7'1. In summary,\n"]], ["block_4", ["where 7] is called the coef\ufb01cient of viscosity of the \ufb02uid or, more simply, the viscosity. Equation\n9.1.2 implies that the velocity gradient is exactly the same throughout the liquid. As this may not\nbe the case over macroscopic distances, our best assurance of generality is to consider the limiting\ncase, in which Ay and therefore Av approach zero. In the limit of these infinitesimal increments Av\nand Ay become dv and dy, respectively, so Equation 9.1.2 becomes\n"]], ["block_5", ["Figure 9.1\nThe relationship between the applied force F per unit area A and the velocity v used in the\nde\ufb01nition of viscosity.\n"]], ["block_6", ["Now let us invoke our experience with liquids of different viscosities, say, water and molasses, and\nimagine the magnitude of the shearing force that would be required to induce the same velocity\ngradient in separate experiments involving these two liquids. Our experience suggests that more\nforce is required for the more viscous \ufb02uid. Since the area of the solid plates in this liquid sandwich\nis also involved, we can summarize this argument by writing a proportionality relation between the\nshear force per unit area, a, and the velocity gradient\n"]], ["block_7", ["328\nDynamics of Dilute Polymer Solutions\n"]], ["block_8", [{"image_3": "340_3.png", "coords": [38, 428, 124, 472], "fig_type": "molecule"}]], ["block_9", [{"image_4": "340_4.png", "coords": [44, 634, 115, 677], "fig_type": "molecule"}]], ["block_10", [{"image_5": "340_5.png", "coords": [45, 547, 130, 573], "fig_type": "molecule"}]], ["block_11", ["\u00e9i_.__Ax/Ay\nAy_yu\nAt\n(9.1.1)\n"]], ["block_12", ["02nd\u2014yzn'jr\n(9.1.3)\n"]], ["block_13", ["y\nA\n"]], ["block_14", ["F\nAv\n,\n"]], ["block_15", ["dv\n"]], ["block_16", ["l\n\u20191\u2019\nI\n"]], ["block_17", ["I\nA\nxH\u2014+ \n"]], ["block_18", ["*\u2018W\u2019\nX\n"]], ["block_19", ["1\u2014\u00bb!\n"]], ["block_20", ["\u2014v__-}J\nI;\n"]]], "page_341": [["block_0", [{"image_0": "341_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "341_1.png", "coords": [33, 342, 161, 384], "fig_type": "molecule"}]], ["block_2", ["To appreciate this result, we remember that Av as introduced in Figure 9.1 is actually Ax/Ar. Thus\nthe product F(Av/Ay) can be written as F(Ax/At)/Ay, and F(Av/Ay)/A becomes FAx/AAyAt. The\nproduct of a force and the distance through which it Operates equals an energy AE and the product\nof A and Ay equals the volume element of AV, upon which the shearing force described in\n"]], ["block_3", ["Equation 9.1.3 is called Newton\u2019s law of viscosity, and those \ufb02uids that follow it (with\n7?\nindependent of the magnitude of the shear rate) are said to be Newtonian.\nEquation 9.1.3 describes a straight line with zero intercept if 0\u2018 is plotted versus the velocity\ngradient; such a plot is shown in Figure 9.2. Since the coefficient of viscosity is the slope of this\nline, this quantity has a single value for Newtonian liquids. Liquids of low molecular weight\ncompounds solutions are generally Newtonian, but quite a few different variations from\nthis behavior are also observed. We shall not attempt to catalog all of these variations, but shall\nonly mention the other pattern of behavior shown in Figure 9.2. This example of non\u2014Newtonian\nbehavior is described as shear thinning, and is often observed when the material under study is a\npolymer solution or melt. Since Equation 9.1.3 defines the coefficient of viscosity as the slope of a\nplot of 0\u2018 versus velocity gradient, it is clear from Figure 9.2 that shear\u2014thinning substances are not\ncharacterized by a single viscosity. The viscosity at a particular velocity gradient is given by the\nratio a/(dv/dy); inspection of Figure 9.2 reveals that shear-thinning materials appear less viscous at\nhigh rates of shear than at low rates. For the purposes of this chapter and the next, we are only\nconcerned with Newtonian response. For shear-thinning \ufb02uids, Newtonian response can be\nachieved by reducing the shear rate sufficiently to access the linear portion near the origin in\nFigure 9.2. This may be formalized by defining the zero shear viscosity, 7103\n"]], ["block_4", ["In most viscometers it is possible to vary '5! in such a way as to ascertain whether 17 is constant. For\nthe remainder of this book, the symbol 7? refers to the zero shear viscosity, unless explicitly stated\notherwise.\nTo see another interpretation of viscosity, we multiply both sides of Equation 9.1.2 by Av/Ay:\n"]], ["block_5", ["Introduction: and Viscosity\n329\n"]], ["block_6", ["Figure 9.2\nComparison of shear stress versus shear rate for Newtonian and shear\u2014thinning behavior.\n"]], ["block_7", ["PM or\n/\n"]], ["block_8", [{"image_2": "341_2.png", "coords": [40, 442, 278, 647], "fig_type": "figure"}]], ["block_9", ["Av\nFAv\ns\n_=__:\n_\u2014\n-1-\n\u201cA, My\n\"(A33\n(9\n5)\n"]], ["block_10", ["770\nEll/1:161)\n(9.1.4)\n"]], ["block_11", ["Newtonian (high 17)\n"]], ["block_12", [",\nShear thinning\n"]], ["block_13", ["Zero shear limit\n"]], ["block_14", ["dv/dy 7\"\n"]], ["block_15", ["Newtonian (low 17)\n"]]], "page_342": [["block_0", [{"image_0": "342_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["does not tell the whole story, however. Other forces are also operative: external forces of, say,\ngravitational or mechanical origin are responsible for the motion in the \ufb01rst place, and pressure forces\nare associated with the velocity gradient. In general, things are not limited to one direction in space,\nbut the forces, gradients, and velocities have x, y, and 2 components. It is possible to bring these\nconsiderations together in a very general form by adding together all of the forces acting on a volume\nelement of liquid, including viscous forces, and using the net force and the mass of the\nvolume element in Equation 9.2.2. The resulting expression is called the equation ofmotion and is\n"]], ["block_2", ["The deforming forces that induce \ufb02ow in \ufb02uids are not recovered when these forces are removed.\nThese forces impart kinetic energy to the molecules, which is dissipated within the \ufb02uid as heat.\nEquation 9.1.7 implies that at sufficiently high shear rates the amount of heat generated could lead\nto measurable temperature increases; this phenomenon of viscous heating is indeed a concern in\nhigh shear-rate measurements and applications.\nWe conclude this section with a consideration of the units required for r] by Equation 9.1.3. To\ndo this, we rewrite these equations in terms of the units of all quantities except 1). The units of 7\u20191\nmust make the expressions dimensionally correct. Force has units of mass times acceleration, or\ng cm/s2 in cgs units. Since the viscosity gradient has the units 3\u20181 and area is cmz, the dimensional\nstatement of Equation 9.1.3 is\n"]], ["block_3", ["The shearing force F that is part of the definition of viscosity can also be analyzed in terms of\nNewton\u2019s second law and written as\n"]], ["block_4", ["In order to satisfy this equation, 11 must have units g/cm/s, which is defined to be 1 poise (1 P). In\nthe SI system the units are kg/m/s, or pascal seconds (Pa 3). At room temperature, water has a\nviscosity of about 0.01 P or 0.001 Pa 5 and other low molecular weight liquids have comparable\nviscosities. The viscosity of a polymer liquid depends very much on the concentration and\nmolecular weight of the polymer, as we shall see in Section 9.3 and in Chapter 11, and it can be\nmany orders of magnitude larger than 0.01 P.\n"]], ["block_5", ["When the force is divided by the area of a shearing plane, A, to obtain a stress, we would also write\n"]], ["block_6", ["Figure 9.1 operates. Therefore FAx/AAyAt is the same as AE/AVAI. Defining the increment in\nshear energy dissipated per unit volume by the symbol AW, we obtain\n"]], ["block_7", ["If it were complete, Equation 9.2.2 would be a differential equation whose solution would give v, the\nvelocity of the \ufb02owing liquid, as a function of time and position within the sample. Equation 9.2.2\n"]], ["block_8", ["9.2\nStokes\u2019 Law and Einstein's Law\n"]], ["block_9", ["330\nDynamics of Dilute Polymer Solutions\n"]], ["block_10", ["As in the parallel case of going from Equation 9.1.2 to Equation 9.1.3, we take the limit of\ninfinitesimal increments and write\n"]], ["block_11", [{"image_1": "342_1.png", "coords": [39, 150, 117, 189], "fig_type": "molecule"}]], ["block_12", [{"image_2": "342_2.png", "coords": [40, 83, 120, 123], "fig_type": "molecule"}]], ["block_13", [{"image_3": "342_3.png", "coords": [40, 520, 129, 558], "fig_type": "molecule"}]], ["block_14", ["F\nmdv\ndv\n_=__:\n._\n.22\nA\nA dt\nndy\n(9\n)\n"]], ["block_15", ["dv\nF:\n= _\n9.2.1\nma\nmdr\n(\n)\n"]], ["block_16", ["AW\nAv\n2\n37: \n(9.1.6)\n"]], ["block_17", ["dW\ndv\n2\n"]], ["block_18", ["g cm s\u2018z/cm2 (n) 5*1\n"]], ["block_19", ["dt \n(9.1.7)\n"]]], "page_343": [["block_0", [{"image_0": "343_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "343_1.png", "coords": [20, 437, 264, 652], "fig_type": "figure"}]], ["block_2", ["The first of these problems involves relative motion between a rigid sphere and a liquid, as\nanalyzed by Stokes around 1850 [1]. The results apply equally to liquid \ufb02owing past a stationary\nsphere with a steady-state (subscript \u201c0\u201d) velocity v0, or to a sphere moving through\na\nstationary liquid with a velocity \u2014v0; the relative motion is the same in both cases. If the relative\nmotion is in the vertical direction, we may visualize the slices of liquid described above as\nconsisting of a bundle of layers, some of which are shown schematically in Figure 9.3.\nIn the horizontal plane containing the center of the sphere a limiting velocity v0 is reached as the\ndistance r from the center of the sphere becomes large. This would be the observable settling\nvelocity of such a spherical particle, for example. In an infinitesimally thin layer adjacent to the\nsurface, the tangential component of velocity would be that of the solid sphere. This is the no slip\ncondition as it applies to this problem. This means that a velocity gradient exists that is described in\nterms of the distance from the center of the sphere and the component of velocity perpendicular to\nr. The velocity gradient dv/dr should certainly be proportional to v0 and then we need a quantity\nwith units of length to be dimensionally correct. The only relevant length we have is R, the radius\n"]], ["block_3", ["Figure 9.3\nDistortion of flow streamlines around a Spherical particle of radius R. The relative velocity in\nthe plane containing the center of the sphere equals v0 as r \u2014> oo.\n"]], ["block_4", ["the cornerstone of \ufb02uid mechanics. The equation of motion is generally written in terms of vector\noperators and takes on a variety of forms, depending on the system of coordinates and the vector\nidentities that have been employed. If the equation of motion is complicated even in writing, things\nare even worse in the solving. Accordingly, we will not reproduce a full treatment here, but rather\noutline the essential points.\nAs with any differential equation, an important part of solving the equation of motion is\ndefining the boundary conditions. An important boundary condition was introduced above, namely\nthat no slip occurs at the boundary between a moving \ufb02uid and a rigid wall. The \u201cno slip\u201d or\n\u201cstick\u201d condition means that the \ufb02uid layer immediately adjacent to a stationary wall has a\nvelocity of zero, with successive layers away from the wall possessing larger increments of\nvelocity. At a sufficiently large distance from the wall, a net velocity is attained, which is\nunperturbed by the presence of the surface. In Section 9.4 we shall apply this idea to the \ufb02ow of\na liquid through a capillary tube. For now we consider two classic problems involving the effect of\nrigid spheres on the \ufb02ow behavior of a liquid.\n"]], ["block_5", ["9.2.1\nViscous Forces on Rigid Spheres\n"]], ["block_6", ["Stokes\u2019 Law and Einstein's Law\n331\n"]], ["block_7", ["..\nIv.\n"]]], "page_344": [["block_0", [{"image_0": "344_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The second classic problem arises in describing the viscosity of a suspension of spherical particles.\nThis problem was analyzed by Einstein in 1906, with some corrections appearing in 1911 [2]. As\nwith Stokes\u2019 law, we shall only present qualitative arguments to give plausibility to the final form.\nThe fact that it took Einstein five years to work out the \u201cbugs\u201d in this theory is an indication of the\ncomplexity of the formal analysis. Derivations of both the Stokes and Einstein equations that do\nnot require vector calculus have been presented by Lauffer [3]; the latter derivations are at about\nthe same level of difficulty as most of the mathematics in this book, but are lengthy.\nWe return to Figure 9.1 with the stipulation that the volume of \ufb02uid sandwiched between the\ntwo plates is a unit of volume. This unit is defined by a unit of contact area with the walls and a unit\nof separation between the two walls. Next we consider a shearing force acting on this cube of \ufb02uid\nto induce a unit velocity gradient. According to Equation 9.1.7, the rate of energy dissipation per\nunit volume from viscous forces dW/dt is proportional to the square of the velocity gradient, with\nn, the factor of proportionality. Thus, to maintain a unit gradient, a volume rate of energy\ndissipation equal to n, is required.\nNext we consider replacing the sandwiched fluid with the same liquid, but in which solid\nspheres are suspended at a volume fraction qb. Since we are examining a unit volume of liquid\u2014-\u2014a\nsuspension of spheres in this case\u2014the total volume of the spheres is also qb. We begin by\nconsidering the velocity gradient if the velocity of the t0p surface is to have the same value as\nin the case of the pure liquid. Being rigid objects, the suspended spheres contribute nothing to the\nvelocity gradient. As far as the gradient is concerned, the Spheres might as well be allowed to settle\nto the bottom and then be fused to the lower, stationary wall. The equivalency of the suspended\nspheres and a uniform layer of the same volume are illustrated schematically in Figure 9.4a and\nFigure 9.4b, respectively. Since the unit volume has a unit cross-sectional area, a volume qb fused\nto the base will raise the stationary surface by a distance qS and leave a liquid of thickness 1 d) to\ndevelop the gradient. These dimensions are also shown in Figure 9.4b. If the velocity of the top\nlayer is required to be the same in this case as for the pure solvent, then the gradient in the liquid\nneed only be the fraction 1/(1 (b) of that for the pure liquid. Of course, since (p is less than unity,\nthis \u201cfraction\u201d is greater than unity.\nNow we return to consider the energy that must be dissipated in a unit volume of suspension to\nproduce a unit gradient, as we did above with the pure solvent. The same fraction applied to the\nshearing force will produce the unit gradient and the same fraction also describes the volume rate\n"]], ["block_2", ["This famous equation is Stokes\u2019 law for rigid spheres. We emphasize that the viscosity in Equation\n9.2.3 is that of the medium surrounding the sphere by labeling it with the subscript \u201cs\u201d for solvent,\nand thus\nf :67mg}?\n(9.2.4)\n"]], ["block_3", ["of the sphere, and thus we anticipate that dv/dr is proportional to vo/R. Formal analysis of the\nproblem via the equation of motion verifies this argument from dimensional analysis and provides\nthe necessary proportionality factors as well.\nFrom Equation 9.2.2 we see that the total viscous force associated with this motion equals\n"]], ["block_4", ["for Spherical particles of radius R.\n"]], ["block_5", ["7) x (dv/dr) >< (area), where the pertinent area is proportional to the surface of the sphere and\ntherefore varies as R2. This qualitative argument suggests that the viscous force opposing the\nrelative motion of the liquid and the sphere is proportional to [n(vo/R)] (R2). The complete solution\nto this problem reveals that this is correct; both pressure and shear forces arising from the motion\nare proportional to nRvO, and the total force of viscous resistance is given by\n"]], ["block_6", ["9.2.2\nSuspension of Spheres\n"]], ["block_7", ["332\nDynamics of Dilute Polymer Solutions\n"]], ["block_8", ["Fvis 67rnRv0 =fv0\n(9.2.3)\n"]]], "page_345": [["block_0", [{"image_0": "345_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "345_1.png", "coords": [17, 186, 242, 288], "fig_type": "figure"}]], ["block_2", [{"image_2": "345_2.png", "coords": [26, 54, 220, 305], "fig_type": "figure"}]], ["block_3", [{"image_3": "345_3.png", "coords": [32, 49, 263, 166], "fig_type": "figure"}]], ["block_4", ["d) is small. This being the case, 1/(1\u20149\u20185)2 can be replaced by the leading terms of the series\nexpansion (1+ (,5 + (p2 + -)2 to give\n7,: n,(1+%\u00a2)(1+\u00a2+~)2: n.(1+2.5\u00a2+4\u00a22+~)\n(9-2-7)\n"]], ["block_5", ["This is Einstein\u2019s famous viscosity equation; the following observations are pertinent:\n"]], ["block_6", ["Again, since qS < 1, 17 > 773.\nThis is only one of the contributions to the total volume rate of energy dissipation; a second\nterm that arises from explicit consideration of the individual spheres must also be included. This\nsecond effect can be shown to equal 1.5 (\ufb01ns/(1 q5)2; therefore the full theory gives a value for 7),\nthe viscosity of the suspension:\n"]], ["block_7", ["1.\n1) is the viscosity of the suspension as a whole; 17, is the viscosity of the solvent; and (pi) is the\nvolume fraction occupied by the spheres.\n2.\nThe validity of the derivation is limited to small values of qt\u00bb, 30 Equation 9.2.7 is generally\ntruncated after the first two terms on the right\u2014hand side.\n3.\nThe viscosity does not depend on the radius of the spheres, only on their total volume fraction.\n"]], ["block_8", ["One additional assumption that underlies the derivation of the second term in Equation 9.2.6 is that\n"]], ["block_9", ["Figure 9.4\n(a) Schematic representation of a unit cube containing a suspension of spherical particles at\nvolume fraction d). (b) The volume equivalent to the spheres in (a) is fused to the base, leaving 1 d) as the\nthickness of the liquid layer.\n"]], ["block_10", ["of energy dissipation compared to the situation described above for pure solvent. Since the latter\nwas 77, we write for the suspension, in the case of dv/dy 1,\n"]], ["block_11", ["(a)\n43), Volume occupied by spheres\n"]], ["block_12", ["Stokes' Law and Einstein's Law\n333\n"]], ["block_13", [{"image_4": "345_4.png", "coords": [40, 474, 208, 511], "fig_type": "molecule"}]], ["block_14", ["E;5523E5335g?5533555335553E5255555E;Egigi55555555;E553555;225555355325552355335355-135555 :=-'\nequivalent\nto spheres = q)\n"]], ["block_15", [",..A...............-----..1...;_|_n_n_.........\n"]], ["block_16", ["dW\n1\n\u2014 Z\nZ \n9.2.5\n"]], ["block_17", ["1.5\n1\n.5\nits\naims _\n+0\n(i5\n(92.6)\nn=1_\u00a2+(1_\u00a2)z\u2014m(l_\u00a2)2\n"]], ["block_18", ["<33\n"]], ["block_19", ["Q\nUnit area\n"]], ["block_20", ["I\nl\n"]], ["block_21", ["..\ni Thickness of\n\u2018\n\u2014i \n"]], ["block_22", ["0\nthickness\n"]], ["block_23", ["Thickness of\nImmdz1\u2014a\n"]], ["block_24", [".\n__\n"]], ["block_25", ["Unit\n"]]], "page_346": [["block_0", [{"image_0": "346_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where we have introduced the intrinsic viscosity, [1)], in place of a\u2019. Note that by dimensional\nanalysis we can see that [17] must have units of inverse concentration, typically mL/g. Conse-\nquently it is not actually a viscosity; rather, it is a coefficient that quantifies the rate at which the\nsolution viscosity increases per g/mL of added solute, when 6 is small. As 6 increases further, the\n"]], ["block_2", ["The intrinsic viscosity [1)], a quantity that will be defined formally below, is a measure of the\nability of added polymer to increase the viscosity of the solution over that of the solvent. It turns\nout that [n] is directly related to the size of the polymer in solution, and before SEC (to be\ndiscussed in Section 9.8) it was the most commonly employed method for determining molecular\nweight. It is still useful in this regard, but it will also provide us with a basis for understanding\nmuch about the behavior of polymers in solution. We begin by proposing that the dilute solution\nviscosity can be written as a power series in concentration, by analogy with the osmotic pressure\n(Equation 7.4.7) and the Einstein relation (Equation 9.2.7):\n"]], ["block_3", ["where a and b are unspeci\ufb01ed coefficients and 773 is the solvent viscosity. We can rearrange\nEquation 9.3.1 by factoring out 17,:\n"]], ["block_4", ["9.3.1\nGeneral Considerations\n"]], ["block_5", ["1.\nThe liquid medium is assumed to be continuous. This makes the results suspect when applied\nto spheres, which are so small that the molecular nature of the solvent cannot be ignored.\n2.\nBoth relationships have been repeatedly verified for a variety of systems and for spheres with a\nwide range of diameters. Despite item (1), both Equation 9.2.4 and Equation 9.2.7 have often\nbeen applied to individual molecules, for which they often work surprisingly well.\n3.\nThe spherical geometry assumed in the Stokes and Einstein derivations gives the highly\nsymmetrical boundary conditions favored by theoreticians. For ellipsoids of revolution having\naxial ratio a/b, friction factors have been derived by Perrin [4] and the coefficient of the first-\norder term in Equation 9.2.7 has been derived by Simha [5]. In both cases, the calculated\nquantities increase as the axial ratio increases above unity. For spheres, of course, a/b :1.\n4.\nIn the derivation of both Equation 9.2.4 and Equation 9.2.7, the disturbance of the \ufb02ow\nstreamlines is assumed to be produced by a single particle. This is the origin of the limitation\nto dilute solutions in the Einstein theory, where the net effect of an array of spheres is treated\nas the sum of the individual nonoverlapping disturbances. When more than one sphere is\ninvolved, the same limitation also applies to Stokes\u2019 law. In both cases, contributions from the\nwalls of the container are also assumed to be absent.\n"]], ["block_6", ["4.\nBy describing the concentration dependence of an observable pr0perty as a power series,\nEquation 9.2.7 plays a comparable role for viscosity as Equation 7.4.7 does for osmotic pressure.\n5.\nThe volume fraction emerges from the Einstein derivation as the natural concentration unit to\ndescribe viscosity. This parallels the way volume fraction arises as a natural thermodynamic\nconcentration unit in the Flory\u2014Huggins theory, as seen in Section 7.3.\n"]], ["block_7", ["We shall make further use of the Stokes equation later in this chapter; for the present, viscosity is\nour primary concern, and the Einstein equation is our point of departure.\n"]], ["block_8", ["9.3\nIntrinsic Viscosity\n"]], ["block_9", ["334\nDynamics of Dilute Polymer Solutions\n"]], ["block_10", ["Both the Stokes and Einstein equations have certain features in common, which arise from the\nhydrodynamic origins they share:\n"]], ["block_11", ["n=ns+ac+b62m\n(9.3.1)\n"]], ["block_12", ["n 173(1 +a'c+b'62~-)\n= 773(1 + Cln] + kh \u20ac2[\ufb0212\u00b0\u00b0)\n(9,3,2)\n"]]], "page_347": [["block_0", [{"image_0": "347_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "347_1.png", "coords": [21, 528, 396, 662], "fig_type": "figure"}]], ["block_2", [{"image_2": "347_2.png", "coords": [32, 359, 180, 390], "fig_type": "molecule"}]], ["block_3", ["Table 9.1\nSummary of Names and De\ufb01nitions of the Various Functions of 17, ns, and c\nin Which Solution Viscosities Are Frequently Discussed\n"]], ["block_4", ["next term, proportional begins contribute. Justthe osmotic \n"]], ["block_5", ["In the literature one encounters a variety of different ways of presenting the solution and\nthese are collected Table 9.1 for reference purposes.\nHaving defined [7}], now we proceed to see what it can tell us about the polymer solute. Equation\n9.3.2, which was just pr0posed on general principles, is analogous to Einstein\u2019s equation (Equation\n9.2.7), which using hydrodynamics. To compare these equations directly, \nconvert the concentration c into a volume fraction, qb. This can be done as follows:\n"]], ["block_6", ["where we have left the solute volume rather vaguely defined. For a rigid sphere, as envisioned by\nEinstein, this volume is simply Avogadro\u2019s number x (417/3)R3. For other shapes, however, or for\n\ufb02exible coils, it is not so clear what volume is appropriate. The molar volume is a thermodynamic\nquantity, but the viscosity measurement is hydrodynamic, not thermodynamic. Accordingly, we\nfinesse this issue by defining a hydrodynamic volume for the molecule, Vh:\n"]], ["block_7", ["which, by comparison with Equation 9.3.2 relates the intrinsic viscosity to the hydrodynamic volume:\n"]], ["block_8", ["Now we make a bold step. We propose that the hydrodynamic volume is proportional to the radius\nof gyration cubed:\n"]], ["block_9", ["the term in re\ufb02ectsinteractions molecules; kind of\nvirial coefficient for viscosity and is known as the Huggins coefficient. We can rearrange Equation\n9.3.2 to obtain a formal definition of [77]:\n"]], ["block_10", ["and we insert Equation 9.3.5 into Equation 9.2.7:\n"]], ["block_11", ["Symbol\nDe\ufb01nition\nCommon name\nIUPAC name\n"]], ["block_12", ["\u201dn1.\n1\nRelative viscosity\nViscosity ratio\n7\u20193\n"]], ["block_13", ["Intrinsic \n335\n"]], ["block_14", ["mp\nI'\u2014 1\nSpecific viscosity\n\u2014\n"]], ["block_15", ["nred\n(1 (\u2014n\u2014 1)\nReduced viscosity\nViscosity number\n773\n"]], ["block_16", ["mm,\n% ln (1)\nInherent viscosity\nLogarithmic viscosity number\n"]], ["block_17", [".\nn\u2014m\n..\n.\n.\n...\n.\n.\n[7}]\n11m)\n\u2014En\u2014\nIntrmsrc v13cosrty\nLimiting Viscosrty number\nc\u2014r\n5\n"]], ["block_18", [{"image_3": "347_3.png", "coords": [46, 463, 96, 498], "fig_type": "molecule"}]], ["block_19", ["a :\ufb01Navl/h\n(9.3.5)\n"]], ["block_20", ["Solute volume\nc\n=\n= \u2014 V\n1\n. .4\n<1\"\nSolution volume\nM(\no ume of a mole of solute)\n(9 3\n)\n"]], ["block_21", ["V1, \u2014R3\n(9.3.8)\n"]], ["block_22", ["5 c\n77: 775(1+\u00a7\u2014M-Navvh+\u201d')\n(9.3.6)\n"]], ["block_23", ["[771211111 (77\u20141173)\n(9.3.3)\nCTO\nC773\n"]], ["block_24", ["[77] 51:3\" g\n(9.37)\n"]], ["block_25", [{"image_4": "347_4.png", "coords": [56, 528, 243, 668], "fig_type": "figure"}]], ["block_26", ["773\n"]]], "page_348": [["block_0", [{"image_0": "348_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "348_1.png", "coords": [24, 194, 160, 234], "fig_type": "molecule"}]], ["block_2", ["Poly(ethylene oxide)\nToluene\n35\n14.5\n0.70\nPoly(ethyleneterephthalate)\nm-Cresol\n25\n0.77\n0.95\n1,4-Polybutadiene\nCyclohexane\n20\n36\n0.70\n1,4\u2014Polyisoprene\nDioxane (0)\n34\n135\n0.50\nPoly(hexamethy1ene adipamide)\nm\u2014Cresol\n25\n240\n0.61\nPoly(dimethy1siloxane)\nToluene\n25\n21.5\n0.65\nPolyethylene\np\u2014Xylene\n75\n135\n0.63\nPolyprOpylene\nCyclohexane\n30\n20.9\n0.76\nPolyisobutylene\nBenzene\n40\n43\n0.60\nPoly(methyl methacrylate)\nChloroform\n25\n4.8\n0.80\nPoly(vinyl chloride)\nChlorobenzene\n30\n71.2\n0.59\nPoly(viny1 acetate)\nEthanol (0)\n56.9\n90\n0.50\nPoly(viny1 alcohol)\nWater\n25\n20\n0.76\nPolystyrene\nToluene\n25\n17\n0.69\nPolystyrene\nCyclohexane (0)\n35\n80\n0.50\nPolyacrylonitrile\ndimethyl formamide\n50\n30\n0.752\nPoly(8-caprolactam)\nm-Cresol\n25\n320\n0.62\n"]], ["block_3", ["Source: From Kurata, M. and Tsunashima, Y., in Polymer Handbook, 3rd ed., Brandrup, J. and Immergut, E.H. (Eds),\nWiley, New York, 1989.\n"]], ["block_4", ["This gives what we were looking for: a direct relation between the intrinsic viscosity and the\nmolecular weight. Using the values of V cited above, we can see that [77] should be independent of\nM for a rigid sphere, increase as Mor MA\"5 for a \ufb02exible chain in a theta solvent or good solvent,\nrespectively, and increase as M2 for a rod. These various possibilities are encompassed by the\ngeneral relation\n"]], ["block_5", ["Table 9.2\nValues for the Mark\u2014Houwink Parameters for a Selection of Polymer\u2014Solvent Systems\nat the Temperatures Noted\n"]], ["block_6", ["a The units for k itself depend on the value of a; the indicated value gives [1]] in mL/g.\n"]], ["block_7", ["We know from Chapter 6, and Equation 6.6.1 in particular, that RgrvM\u201d, where the exponent u\ntakes on various characteristic values (1/3 for a solid sphere, 1/2 for a \ufb02exible coil in a theta solvent\nor in the melt, 3/5 for a \ufb02exible chain in a good solvent, and\n1 for a rigid rod). If we insert this\nrelation into Equation 9.3.8, and the result into Equation 9.3.7, we have\n"]], ["block_8", ["This relationship with a 1 was first proposed by Staudinger [6], but in this more general form it is\nknown as the Mark\u2014Houwiuk equation [7]. The constants k and a are called the Mark\u2014Houwink\nparameters for a system. The numerical values of k and a depend on both the nature of the polymer\nand the nature of the solvent, as well as the temperature; extensive tabulations are available [8,9]\nand Table 9.2 gives a few examples. (Note, however, that the values can vary for a given system\namong different investigators, and that attention must be paid to details such as microstructure,\n"]], ["block_9", ["Polymer\nSolvent\nT (\u00b0C)\nks x 103 (mL/g)El\na\n"]], ["block_10", ["In other words, the volume that matters in the viscosity experiment is not the volume actually\noccupied by the polymer segments (which would be the degree of polymerization times the volume\nof the monomer), but the volume pervaded by the entire molecule. For a random coil this means\nthat we assume that the molecule behaves hydrodynamically like a rigid sphere of radius Rg. This\nmight seem rather far fetched at first glance, but in fact, it is basically correct.\n"]], ["block_11", ["336\nDynamics of Dilute Polymer Solutions\n"]], ["block_12", ["9.3.2\nMark-Houwink Equation\n"]], ["block_13", ["R3\nM31;\n[77] M3\"\u20141\n(939)\n"]], ["block_14", ["[77] kM\u201c\n(9.3.10)\n"]], ["block_15", [{"image_2": "348_2.png", "coords": [96, 419, 453, 638], "fig_type": "figure"}]]], "page_349": [["block_0", [{"image_0": "349_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["molecular weight range, and polydispersity of the samples.) Since viscometer drainage times\n(see Section 9.4) are typically on the order of a few hundred seconds, intrinsic viscosity experiments\nprovide a rapid method for evaluating the molecular weight of a polymer. A possible drawback to\nthe method is that the Mark\u2014Houwink coefficients must be established for the particular system\nunder consideration by calibration with samples of known molecular weight, but given the\nextensive tabulations this is often not a significant limitation.\nThe values of the exponent a in Table 9.2 range between 0.5 and about 1. Although in practice\na m 0.5 in all theta systems, in good solvents the values vary quite a bit. Part of this is a\nconsequence of the slow crossover to good solvent limiting behavior anticipated by the Flory\u2014\nKrigbaum theory (Equation 7.7.10). In other words, even though a solvent might be \u201cbetter-\nthan-theta\u201d for a given polymer, the experimental temperature might not be sufficiently far above\nT 0, or the molecular weight range might not extend to sufficiently high values, to obtain :2 3/5\nand thus a 0.8. In some cases a > 0.8, which can be attributed to a semi\ufb02exible structure (recall\nthat the simple argument above predicts a =2 for a rod). There is another, more complicated\nreason why the exponents can take on a range of values, the phenomenon of hydrodynamic\ninteractions, and this will be explored in Section 9.7.\nEven fractionated polymer samples are generally polydisperse, which means that the molecular\nweight determined from intrinsic viscosity experiments is an average value. The average obtained\nis called the viscosity average molecular weight, MV, which can be derived as follows:\n"]], ["block_2", ["Table 9.3 lists the intrinsic viscosity for a number of polystyrene samples of different molecular\nweights. The M values are weight averages based on light scattering experiments. The values of [n]\nwere measured in cyclohexane at the theta temperature of 35\u00b0C. In the following example we\nconsider the evaluation of Mark\u2014Houwink coefficients from these data.\n"]], ["block_3", ["where n,- is the number of molecules with molecular weight M,- and we assume all species\n1'\nhave the same k and a.\n4.\nCombining Equation 9.3.11 and Equation 9.3.13 we obtain\n"]], ["block_4", ["where the index 2' refers to different molecular weights.\n3.\nWe can now obtain [77] as\n_.\n77\u2014773\n__.\nZQI\u2019UL\u2018\n\u201c'1 \n3936(T)\n\u2014 is(Tl\n_ _ k \nH\nZniMi/V\n_\nEra-M,-\n"]], ["block_5", ["2.\nThe dilute solution viscosity for the polydisperse system can be expressed as\n"]], ["block_6", ["Intrinsic Viscosity\n337\n"]], ["block_7", ["1.\nThe experimental intrinsic viscosity is proportional to some average value, My, raised to the\npower a, according to Equation 9.3.10,\n"]], ["block_8", [{"image_1": "349_1.png", "coords": [38, 415, 218, 485], "fig_type": "molecule"}]], ["block_9", [{"image_2": "349_2.png", "coords": [50, 522, 145, 566], "fig_type": "molecule"}]], ["block_10", ["l/a\nEJliMI-Fa\nMVE\n\u2014\u2019-\u2014\u2018\u2014\n9.3.14\n( $5\"e\n(\n)\n"]], ["block_11", ["For flexible polymers in general, Mn < M,r (MW, and Mv 2MW if a 1. On the basis of this\nlast observation, it can be argued that the Mark\u2014Houwink coefficients should be evaluated\nusing weight average rather than number average molecular weights as calibration standards.\nWe saw in Chapter 8 how MW values can be obtained from light scattering experiments.\n"]], ["block_12", ["n n.(1+ Ze.[n1.\u2014+-~)\n(9.3.12)\n"]], ["block_13", ["[1;] kf\n(9.3.11)\n"]], ["block_14", ["(9.3.13)\n"]]], "page_350": [["block_0", [{"image_0": "350_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "350_1.png", "coords": [0, 59, 343, 321], "fig_type": "figure"}]], ["block_2", ["that corresponds to k=0.088 and (1 =0.497. The exponent agrees with expectation for a theta\nsolvent; the uncertainty in a is at least i0.003. Below M 10,000 the relationship between [17] and\nM becomes more complicated. A full explanation of this dependence is not yet available, but\namong the contributing factors are non-Gaussian conformations for short chains, chain stiffness,\nchain end effects, and modification of the solvent dynamics in the vicinity of the chain.\n"]], ["block_3", ["Table 9.3\nIntrinsic Viscosity as a Function of Molecular Weight for Samples\nof Polystyrene\n"]], ["block_4", ["266\n1.49\n19800\n11.9\n370\n1.92\n44000\n18.0\n474\n2.47\n50500\n19.2\n578\n2.74\n55000\n20.0\n680\n2.93\n76000\n24.5\n904\n3.26\n96200\n26.0\n1,480\n3.70\n1.25 x 105\n29.0\n1,780\n4.08\n1.60 x 105\n34.0\n2,270\n4.50\n1.80 x 105\n35.4\n3,480\n5.43\n2.47 x 105\n42.0\n5,380\n6.59\n3.94 x 105\n54.5\n10,100\n9.00\n4.06 x 105\n55.0\n20,500\n12.3\n5.07 x 105\n60.0\n40,000\n17.2\n6.22 x 105\n66.0\n97,300\n27.3\n8.62 x 105\n78.0\n1.91 x 105\n38.0\n1.05 x 10\u20185\n86.0\n3.59 x105\n51.2\n1.56 x106\n106\n7.32 x 105\n73.4\n1.32 X 106\n98.1\n"]], ["block_5", ["Evaluate the Mark\u2014Houwink coefficients for polystyrene in cyclohexane at 35\u00b0C from the data in\nTable 9.3. How well do the two data sets agree? What is the appropriate set of Mark\u2014Houwink\nparameters for high molecular weight PS in this solvent? Does the exponent agree with expect-\nation? At what molecular weight does this relation break down?\n"]], ["block_6", ["Visually it is clear that the two data sets agree extremely well over the common molecular weight\nrange. Furthermore, for M greater than about 10,000 the data fall on a straight line. A combined\nleast squares fit of the two sets of data in this range gives\n"]], ["block_7", ["The two data sets are plotted in Figure 9.5 in a log\u2014log format. This particular choice is helpful\nbecause adherence to a power law relationship, such as the Mark\u2014Houwink relation, will result in\nthe data falling on a straight line with the exponent as the slope. From Equation 9.3.10\n"]], ["block_8", ["Source: From (a) Yamakawa, H., Abe, F., and Einaga, Y., Macromolecules, 26, 1891, 1993;\n(b) Berry, G., J. Chem. Phys, 46, 1338, 1967.\n"]], ["block_9", ["338\nDynamics of Dilute Polymer Solutions\n"]], ["block_10", ["(a) M...\n[77] (mL/g)\n(b) M...\n[n] (mL/g)\n"]], ["block_11", ["For \ufb02exible coils the value of the Mark\u2014Houwink exponent tells us something about the solvent\nquality, independent of the polymer or solvent. Is there anything similarly general to be extracted\nfrom the proportionality factor k? The answer is yes. In a theta solvent, particularly, it provides\n"]], ["block_12", ["Solution\n"]], ["block_13", ["Example 9.1\n"]], ["block_14", ["log [77] \u2014l.06 + 0.49710gM\n"]], ["block_15", ["log [77] logk + alogM\n"]]], "page_351": [["block_0", [{"image_0": "351_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "351_1.png", "coords": [22, 430, 209, 471], "fig_type": "molecule"}]], ["block_2", [{"image_2": "351_2.png", "coords": [28, 38, 408, 246], "fig_type": "figure"}]], ["block_3", ["Here we have inserted CDC, in place of the numerical prefactor of Equation 9.3.17. More detailed\ntheories give (130 as 2.8 x 1023 as a universal value for all \ufb02exible chains in a theta solvent, which is\n"]], ["block_4", ["Here we have used the subscript 6 to remind us that we are dealing with a theta solvent. Equation\n9.3.17 makes explicit the part of k that depends on the particular polymer in question. For\npolystyrene, M0: 104 g/mol and b: 6.7 x 10\"8 cm (see Table 6.1), and inserting these values\ninto Equation 9.3.17 would predict that k9 $0.12 (for [n] in g/mL). From Example 9.1 above, in\nexperiments k3 0.088, which is only about 30% different from the result utilizing the naive \u201ca\ncoil is pretty much a hard sphere with R =Rg\u201d argument.\nEquation 9.3.17 can be recast in another form, whereby the polymer-specific part is presented as\n(h2)O/M (refer to Table 6.1):\n"]], ["block_5", ["or\n"]], ["block_6", ["We can replace N by M/Mo, where MO is the monomer molecular weight, to obtain\n"]], ["block_7", ["information about the coil dimensions, as the following argument shows. For a theta solvent, we\nknow that R: s/6 (recall Equation 6.5.3). Combining this with Equation 9.3.7 and Equation\n9.3.8 gives\n"]], ["block_8", ["Figure 9.5\nPlot of log [7)] versus logM for the data in Table 9.3. An analysis of the Mark\u2014Houwink\nparameters from these data is presented in Example 9.1.\n"]], ["block_9", [{"image_3": "351_3.png", "coords": [36, 379, 155, 419], "fig_type": "molecule"}]], ["block_10", ["Intrinsic Viscosity\n339\n"]], ["block_11", ["E\n9\u2019\n..\n1\n,\nE\n10\n:L\n13\u2019\n1\n8\u2019\ni\nEVE\na PE,\n3\nYamakawa et al.\n\u2018\n\ufb01x\n-\n"]], ["block_12", ["I\nI\nI IIIII\nI\nI\nI\nIIII\nI\nI\nI\nI\nI\nI IIIII\n102 E\nI-\n-1\nI-\n\u2018\n-\n-\n_\n-\n_\n_\n_\nin] \nmg\ufb01\ns\nas\n-\n"]], ["block_13", [{"image_4": "351_4.png", "coords": [42, 36, 253, 223], "fig_type": "figure"}]], ["block_14", [{"image_5": "351_5.png", "coords": [45, 569, 243, 634], "fig_type": "molecule"}]], ["block_15", ["5N...\nN \n: \u2014-\u2014-V N __ __ _ \n_\n. .15\n[7\u2019]\n2M\n\u201c\n2M\n3\n8\n6\nM63/2\n(9 \n)\n"]], ["block_16", ["b3\n(\u201912)\n3/2\nk0~4.3><10\nMan\u20144.3x\n10\n< M )\n"]], ["block_17", ["100\n1/:\n1|\u201c:a\nI\nn lItIIII\nI IllnIlI\n.\nI .1...\u201c\n102\n103\n104\n105\n106\n"]], ["block_18", ["207m,\nb3\n9 g\u2014_\u2014=4.3 x 1023 \n(9.3.17)\n65/2\n1143/2\nMg/Z\n"]], ["block_19", [".._.\n\u201cif/2\n(9.3.18)\n\u00a2,( M\n"]], ["block_20", ["2077A?av\nb3\n1 2\n"]], ["block_21", ["3/\n"]], ["block_22", ["21, \n- -EEI - -Berry\n"]], ["block_23", ["MW, g/mol\n"]]], "page_352": [["block_0", [{"image_0": "352_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 9.6\nIntrinsic viscosity versus generation number for polyether dendrimers. The dashed curved\nrepresents the simple calculation described in Example 9.2, scaled by a factor of 5. (Data from Mourey,\nT.H., Turner, S.R., Rubinstein, M., Fr\u00e9chet, J.M.J., Hawker, C.J., and Wooley, K.L., Macromolecules, 25,\n2401, 1992.)\n"]], ["block_2", ["The crucial relation we will invoke is Equation 9.3.9, that is,\n"]], ["block_3", ["and so the key step is to see how M and R grow with generation number n. The two\u2014dimensional\npictures of a dendrimer given in Figure 1.2 and Figure 4.7 suggest a roughly spherical structure,\nwith each generation adding an approximately equal increment to R. Furthermore, as the later\ngenerations have larger numbers of units, one might suspect that they become densely packed. If\nthis were the case, then R would grow as M3, and we would recover the Einstein result for hard\nspheres:\n[17] would be independent of M and therefore\n11. In fact, as the data in Figure 9.6\ndemonstrate, this is not true. Remarkably, [1;] goes through a maximum with increasing :1, near\nn r: 3; how does this arise?\nThe origin of this unusual behavior is rather easily understood. It turns out that, although R does\nincrease roughly linearly with :1, meaning that each generation adds a roughly constant increment\nto the total R, the mass added in each generation grows much more rapidly. Let us consider a\n"]], ["block_4", ["Using the arguments given above, predict the dependence of [n] on generation number for\ndendrimer molecules. Recall from Chapter\n1 and Chapter 4 that dendrimers have a tree\u2014like\nstructure, with each generation adding a new layer of material in a completely regular way.\n"]], ["block_5", ["close to the polystyrene result in Example 9.1. The difference in the numerical value of the\nprefactor arises because in our simple model we assumed that a coil behaves as a hard sphere\nwith R:Rg, whereas in fact it behaves as a hard sphere with R o: Rg and the proportionality\nconstant is a little less than 1. Thus measurements of intrinsic viscosity in a theta solvent gives\naccess to the coil dimensions (i.e., COO, b, or Ep) via Equation 9.3.18.\n"]], ["block_6", ["340\nDynamics of Dilute Polymer Solutions\n"]], ["block_7", ["Solution\n"]], ["block_8", ["Example 9.2\n"]], ["block_9", ["EH4?\n!\n\\\nj\nE\ni\n/\n~01\n_\n/\n"]], ["block_10", ["on\n\"\n\\\n\u2018l\n"]], ["block_11", [{"image_1": "352_1.png", "coords": [39, 388, 260, 623], "fig_type": "figure"}]], ["block_12", ["R3\n[n] M\n"]], ["block_13", ["%\n31\u00b0\n/\n_\n:\nf\ni\n-/\n2__\u2019\n.j\n"]], ["block_14", ["6;\n/-H\\\nL.\n"]], ["block_15", ["-\n/\n5\n1\n5;\n/.\n\\\nJ\n"]], ["block_16", ["7_[\nr\n[I\nr\nFlt\u2014TIT!\n[Ill\n[\nI-\n"]], ["block_17", ["1tl_LlJ_l|_ll_I_l_[llll\nl\n1\nIII\nI\nl\u2018\n"]], ["block_18", ["I-\n/\n\\\n-'\n'\n0\n"]], ["block_19", ["l\u2014\n/\n\\.\n_\n"]], ["block_20", ["[\nl\nI\nl\nI\nI\nI\nO\n1\n2\n3\n4\n5\nIn\n"]], ["block_21", ["0')\n"]]], "page_353": [["block_0", [{"image_0": "353_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["R\n10\n20\n30\n40\n50\n60\nM\n100\n400\n1000\n2200\n4600\n9400\nR3/M\n10\n20\n27\n29\n27\n23\n"]], ["block_2", ["We defined the equation of motion as a general expression of Newton\u2019s second law applied to a\nvolume element of \ufb02uid subject to forces arising from pressure, viscosity, and external sources.\nAlthough we shall not attempt to use this result in its most general sense, it is informative to\nconsider the equation of motion as it applies to a specific problem: the \ufb02ow of liquid through a\ncapillary. This consideration not only provides a better appreciation of the equation of motion, but\nalso serves as the basis for an important technique for measuring solution viscosity. We shall\nexamine the derivation first and then discuss its application to experiment.\nFigure 9.7a shows a portion of a cylindrical capillary of radius R and length L. We measure the\ngeneral distance from the center axis of the liquid in the capillary in terms of the variable r and\nconsider specifically the cylindrical shell of thickness dr designated by the broken line in Figure\n9.7a. In general, gravitational, pressure, and viscous forces act on such a volume element, with the\nviscous forces depending on the velocity gradient in the liquid. Our first task then is to examine\nhow the velocity of \ufb02ow in a cylindrical shell such as this varies with the radius of the shell.\nThe net viscous force acting on this volume element is given by the difference between the\nfrictional forces acting on the outer and inner surfaces of the shell:\n"]], ["block_3", ["From this simple calculation, we see that R3/M goes through a maximum for n 3, in excellent\nagreement with the experiments. The origin of the maximum lies in the fact that M grows\ngeometrically with 22 rather than as a power law. Consequently, M in the denominator eventually\nincreases more rapidly than R3 in the numerator. In other words, the dendrimer density is\nincreasing steadily, as opposed to a hard sphere, for which the density is constant with R. This\ndensity increase cannot persist indefinitely, and in fact dendrimers can typically only be grown out\nto n 6\u20148 before there is no more room on the surface to complete another generation.\n"]], ["block_4", ["In this section we consider two standard techniques for measuring viscosity. The first concerns the\nuse of capillary viscometers for low\u2014viscosity \ufb02uids, such as the dilute polymer solutions of\nrelevance to [n]. The second describes the Couette or concentric cylinder geometry, which is\nmore elaborate but is capable of covering a much wider range of n and \u201c5/.\n"]], ["block_5", ["n\n0\n1\n2\n3\n4\n5\n"]], ["block_6", ["n. : 2, M:400 + 6(100) 1000, and R =20 + 10 30. The multiplier 6 in the calculation of M\narises because the previous generation added 3 units, each of which has two functional groups left\nto react further. The results through it 5 computed in this way are listed in the table below:\n"]], ["block_7", ["simple example. Assume trifunctional reactive unit with M 100 g/mol and assume that each\ngeneration adds 10 A to the radius. We can compute a table of M, R, and R3/M as a function of n.\nFor n 0, M 100, and 10 by assumption. (Note that by convention, the initial dendrimer\n\u201ccore\u201d is labeled generation 0.) For n 1, M: 100 + 3(100) 2400, and R 10 + 10 20. For\n"]], ["block_8", ["9.4.1\nPoiseuille Equation and Capillary Viscometers\n"]], ["block_9", ["9.4\nMeasurement of Viscosity\n"]], ["block_10", ["Measurement of Viscosity\n341\n"]], ["block_11", [{"image_1": "353_1.png", "coords": [35, 149, 254, 220], "fig_type": "molecule"}]], ["block_12", [{"image_2": "353_2.png", "coords": [39, 614, 262, 678], "fig_type": "molecule"}]], ["block_13", ["Fvis,net (Fvis)0uter (FViS)inner\ndv\ndv\n=\n_\n\u2014 2\nL\n-\u2014\n9.4.1\n2770\u2018 + dr) L27 (ml-m;-\n77'!\u201d n (m),\n(\n)\n"]]], "page_354": [["block_0", [{"image_0": "354_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "354_1.png", "coords": [20, 86, 321, 227], "fig_type": "figure"}]], ["block_2", [{"image_2": "354_2.png", "coords": [30, 577, 182, 619], "fig_type": "molecule"}]], ["block_3", [{"image_3": "354_3.png", "coords": [32, 79, 170, 215], "fig_type": "figure"}]], ["block_4", ["Figure 9.7\n(a) Portion of a cylinder of radius R and length L showing (by broken lines) section of thickness\ndr. (b) Profile of \ufb02ow velocity in the cylinder. (From Hiemenz, P.C., Principles of Colloid and Surface\nChemistry, Marcel Dekker, New York, 1977.)\n"]], ["block_5", ["342\nDynamics of Dilute Polymer Solutions\n"]], ["block_6", ["Under stationary-state conditions of \ufb02ow, that is, when no further acceleration occurs, this force\nis balanced by gravitational and pressure forces. For simplicity, we assume that the capillary\nis oriented vertically so that gravity operates downward and, for generality, we assume that\nan additional mechanical pressure Ap is imposed between the two ends of the capillary.\nUnder these conditions, the net gravitational and mechanical forces acting on the volume\nelement equal\n"]], ["block_7", ["Integration in the radial direction (along r) converts Equation 9.4.5 to\n"]], ["block_8", ["where 217Lr dr is the volume of the element, 2m dr is its cross-sectional area, p is the density of\nthe \ufb02uid, and g is the acceleration due to gravity. Under the stationary-state conditions we seek to\ndescribe, Equation 9.4.4 and Equation 9.4.3 are equal, and the following relationship applies to the\nvolume element:\n"]], ["block_9", ["where the length times the circumference of the surface describes the apprOpriate area in Equation\n9.1.2. The relationship between the velocity gradient at the two locations is given by\n"]], ["block_10", ["provided that dr is small. Combining Equation 9.4.1 and Equation 9.4.2 and retaining only those\nterms that are first order in dr give\n"]], ["block_11", ["(a)\n(b)\n"]], ["block_12", [{"image_4": "354_4.png", "coords": [36, 635, 156, 675], "fig_type": "molecule"}]], ["block_13", [{"image_5": "354_5.png", "coords": [40, 327, 185, 365], "fig_type": "molecule"}]], ["block_14", ["Fgrammcch, net (2e dr)pg + (2771' dr)Ap\n(9.4.4)\n"]], ["block_15", ["d2v\ndv\nd\ndv\nF(as, net 27711L[r(d7)dr + (5) Cir] :\n217171;; \n(9.4.3)\n"]], ["block_16", ["dv\nl\nAp\n2\n\u2014 :\n\u2014\n_\u2014\n04.6\nnrdr\n2033+ L}\n(9\n)\n"]], ["block_17", ["4%\n"]], ["block_18", ["(1\ndv\nAp\n_\n_\n=\n_\n.4.5\nndr \n(pg + L )r dr\n(9\n)\n"]], ["block_19", ["dv\ndv\ndzv\n(5?)...\u2014 (aids-Jr\n\u20189'\u201c)\n"]], ["block_20", [{"image_6": "354_6.png", "coords": [68, 395, 296, 427], "fig_type": "molecule"}]], ["block_21", ["7\"\n"]], ["block_22", ["/ 11:7:2:3\"\n4\nME/\n\\i:z\u00e9\u2019i/\n"]], ["block_23", ["\\l\n.l\n4\u2014H\u2014b\u20144\u2014r-b\u2014\n\\ \\l I l\n"]], ["block_24", ["33/\nI\nL/\n"]], ["block_25", ["dr\n"]], ["block_26", [{"image_7": "354_7.png", "coords": [205, 72, 322, 232], "fig_type": "figure"}]]], "page_355": [["block_0", [{"image_0": "355_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "355_1.png", "coords": [26, 174, 158, 212], "fig_type": "molecule"}]], ["block_2", [{"image_2": "355_2.png", "coords": [28, 292, 217, 341], "fig_type": "molecule"}]], ["block_3", [{"image_3": "355_3.png", "coords": [33, 284, 221, 365], "fig_type": "figure"}]], ["block_4", ["This result is called the Poiseuille equation, after the researcher who discovered in 1844 this\nfourth\u2014power dependence of flow rate on radius [10]; the unit of viscosity, poise, is also named\nafter him. The following example illustrates the use of the Poiseuille equation in the area where it\nwas first applied.\n"]], ["block_5", ["Poiseuille was a physician\u2014physiologist interested in the \ufb02ow of blood through blood vessels in the\nbody. Estimate the viscosity of blood from the fact that blood passes through the aorta of a healthy\nadult at rest at a rate of about 84 cm3 s\u2014l, with a pressure drop of about 0.98 mmHg m\u2018l. Use 9 mm\nas the radius of the aorta for a typical human.\n"]], ["block_6", ["The units must be expressed in a common system, with the pressure gradient requiring the most\nmodification:\n"]], ["block_7", ["The pumping action of the heart rather than gravity is responsible for blood \ufb02ow; hence the term\nn can be set equal to zero in Equation 9.4.9 and the result solved for n:\n"]], ["block_8", ["where the at r is used to eliminate the integration constant. the\nvelocity gradient to the radial in the \ufb02uid: itand\n"]], ["block_9", ["This result describes a parabolic velocity profile, as sketched in Figure 9.7b.\nEquation 9.4.8 describes the velocity with which a cylindrical shell of liquid moves through a\ncapillary under stationary\u2014state conditions. This velocity times the cross-sectional area of the shell\ngives the incremental volume of liquid dV, which is delivered from the capillary in an interval of\ntime At. The total volume delivered in this interval, AV, is obtained by integrating this product over\nall values of r:\n"]], ["block_10", ["Because of the nonslip condition at the wall, v 0 when r :R, and the constant of integration can\nbe evaluated to give\nVzma)\n4\"?\n(9.4.8)\n"]], ["block_11", ["has a maximum \nEquation 9.4.6 can be integrated again to give v as a function of r:\n"]], ["block_12", ["Measurement of Viscosity\n343\n"]], ["block_13", ["Example 9.3\n"]], ["block_14", ["Solution\n"]], ["block_15", [{"image_4": "355_4.png", "coords": [38, 559, 116, 599], "fig_type": "molecule"}]], ["block_16", ["R\nV\nL\n1%! \n277(9342410/\n)\nJ (R2 _ ,2), d,\n"]], ["block_17", ["A_p _ 0.98 mmHg X 133.3Nm\u20142\nL \nm\n1 mmHg\n= 131 kg m\u20142 s\u20142\n"]], ["block_18", ["J dv wJ r dr\n(9.4.7)\n"]], ["block_19", ["AprrR4\n7\u2019 SLAV/At\n"]], ["block_20", [",__,, \n(949)\n8171.\n"]], ["block_21", ["4\n"]], ["block_22", ["0\n"]]], "page_356": [["block_0", [{"image_0": "356_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "356_1.png", "coords": [30, 262, 126, 310], "fig_type": "molecule"}]], ["block_2", ["Figure 9.8\nA typical capillary viscometer. (From Hiemenz, P.C., Principles of Colloid and Surface\nChemisty, Marcel Dekker, New York, 1977.)\n"]], ["block_3", ["where A represents a cluster of factors that are constant for a particular apparatus. The constant A\nneed not be evaluated in terms of the geometry of the apparatus, but can be eliminated from\n"]], ["block_4", ["Therefore\n"]], ["block_5", ["The Poiseuille equation provides a method for measuring 1) by observing the time required for a\nliquid to flow through a capillary. The apparatus shown in Figure 9.8 is an example of one of many\ndifferent instruments designed to use this relationship. In such an experiment, the time required for\nthe meniscus to drop the distance between the lines etched at opposite ends of the top bulb is\nmeasured. This corresponds to the drainage of a \ufb01xed volume of liquid through a capillary of\nconstant R and L. The weight of the liquid is the driving force for the flow in this case, so the Ap\nterm in Equation 9.4.9 is zero and the observed \ufb02ow time equals\n"]], ["block_6", ["344\nDynamics of Dilute Polymer Solutions\n"]], ["block_7", ["At 370C, the viscosity of pure water is about 0.69 x 10\u20143 kg m\u20141 s\u2018l; the difference between this\nfigure and the viscosity of blood is due to the dissolved solutes in the serum and the suspended red\ncells in the blood. The latter are roughly oblate ellipsoids in shape.\n"]], ["block_8", ["01'\n"]], ["block_9", [{"image_2": "356_2.png", "coords": [38, 361, 97, 648], "fig_type": "figure"}]], ["block_10", ["D\nf?\n"]], ["block_11", ["V\nA: : (:31);\n(9.4.10)\n"]], ["block_12", ["r) =Ap At\n(9.4.11)\n"]], ["block_13", ["C\n"]], ["block_14", ["_ \n_\n8(84 x10-6 m3s-1)\n: 4,0 x 10*3 kg111_15_I\n(or 0.04 P)\n"]], ["block_15", ["\\/\n"]], ["block_16", [",M\n"]]], "page_357": [["block_0", [{"image_0": "357_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where B AV/Sqr. As above, both A and B can be treated as instrument constants and evaluated by\nmeasuring two liquids, which are known with respect to \u201cn and p and then solving a pair of simultaneous\nequations for A and B. A better strategy might be to choose a capillary sufficiently narrow so that\nA: is long enough to eliminate the second term on the right\u2014hand side of Equation 9.4.13.\nOne limitation of this method is the fact that the velocity gradient is not constant in this type of\ninstrument, but varies with r, as noted in connection with Equation 9.4.6. This would be a concern\nif the viscosity were shear\u2014rate dependent over the relevant range. For dilute solutions, and the\nslow \ufb02ows appropriate to determine the intrinsic viscosity, this usually does not matter.\n"]], ["block_2", ["The second standard geometry for solution viscometry is based on concentric cylinders. The illustration\nthat enabled us to define the coefficient of viscosity also suggests a modification that would be\nexperimentally useful. Suppose the two rigid parallel plates in Figure 9.1 and the intervening layers\nof \ufb02uid were wrapped around the z\u2014axis to form two concentric cylinders, with the \ufb02uid under\nconsideration in the gap between them. The required velocity gradient is then established by causing\none of these cylinders to rotate while the other remains stationary. The velocity is now in the direction\ndescribed by the angle 6 in Figure 9.9a, and its gradient is in the radial direction r. Thus the velocity\ngradient in this arrangement may be written dvg/dr.\nSome of the reasons for our interest in this type of viscometer are the following:\n"]], ["block_3", ["Equation 9.4.11 by measuring both a known (subscript 2) and an unknown (subscript 1) liquid in\nthe same \n"]], ["block_4", ["Note that interval depends on both the density and the viscosity of the \ufb02uid, and the ratio\nn/p is sometimes referred to as the kinematic viscosity.\nIn more precise work an additional term, which corrects for effects arising at the ends of the tube, is\nadded to Equation 9.4. 1 1. This correction\u2014which is often negligible\u2014can be incorporated by writing\n"]], ["block_5", ["transmits a force, which can be measured in terms of the torque on a torsion wire to the suspended\nbob. This arrangement is sketched in Figure 9.9b. In this representation, the outer cylinder has \nradius R and the inner cylinder has a radiusfR, wheref is some fraction. The closer this fraction is\nto unity, the narrower will be the gap between the cylinders and the more closely the apparatus will\napproximate the parallel plate model in terms of which 77 is defined.\nA formal mathematical analysis of the \ufb02ow in the concentric cylinder viscometer yields the\nfollowing relationship between the experimental variables and the viscosity:\nf\ntorque force >< radius :47mLR2w 1\u2014f2\n(9.4.14)\n"]], ["block_6", ["1.\nThe basic design is a direct extension of the discussion of viscosity in Section 9.1 and Section 9.2.\n2.\nThe range of applicability is very wide, extending at least from 7; m 0.01\u2014\u20141010 P.\n3.\nThe design permits different velocity gradients to be considered, so that non-Newtonian\nbehavior (e.g., shear thinning) can be investigated, if desired.\n4.\nThe number of technically important viscosity-measuring devices may be thought of as\nvariants of this basic apparatus.\n"]], ["block_7", ["As a practical matter, the outer cylinder is part of a cup that holds the \ufb02uid, while the inner cylinder\nis a coaxial bob suspended within the outer cup. Suppose the cup is centered on a turntable that\nrotates with an angular velocity to, measured in radians per second. The viscous \ufb02uid now\n"]], ["block_8", ["9.4.2\nConcentric Cylinder Viscometers\n"]], ["block_9", ["Measurement \n345\n"]], ["block_10", ["_ \n\u201d'71 \nB; Kt:\n(9-4-12)\n"]], ["block_11", ["p\n:A\nAt\u2014B\u2014\n1)\n9\nAr\n(9.4.13)\n"]], ["block_12", [{"image_1": "357_1.png", "coords": [49, 83, 114, 117], "fig_type": "molecule"}]]], "page_358": [["block_0", [{"image_0": "358_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "358_1.png", "coords": [13, 66, 411, 237], "fig_type": "figure"}]], ["block_2", [{"image_2": "358_2.png", "coords": [28, 78, 170, 218], "fig_type": "figure"}]], ["block_3", ["asf\u2014> 1.\n4.\nThe area of contact A between the cylinders and the \ufb02uid is\n27TRL; therefore 477LR/\n2R(1\u2014f)=A/(1*f)R-\n5.\nThe product (1 f)R is the width of the gap, 5.\n"]], ["block_4", ["in the limit asf \u2014> 1. This is identical to Equation 9.1.3 and is the result we anticipated in rolling\nFigure 9.1 into a cylinder. Equation 9.4.14 is more general than Equation 9.4.16, since its\napplicability is not limited to vanishingly small gaps.\n"]], ["block_5", ["We now turn our attention to the phenomenon of diffusion. This turns out to be directly related to\nthe viscosity through the friction factor, f. Most of us have a sense of diffusion as a randomization\nprocess: place a drop of food coloring into a glass of water, and before long the entire glass will be\nuniformly colored even without stirring. The time it takes for the color to spread depends on two\n"]], ["block_6", ["This equation can be cast into a more recognizable form by assuming thatfis very close to unity. In\nthat case we have the following:\n"]], ["block_7", ["Since v9 is the difference in velocity between the inner and outer cylinders and 5 is the difference\nin the radial location of the two rigid surfaces, Equation 9.4.15 becomes\n"]], ["block_8", ["346\nDynamics of Dilute Polymer Solutions\n"]], ["block_9", ["1.\nThe radius R applies to the entire \ufb02uid sample. Since torque equals the product of force and R,\ncanceling out one power of R leaves the shearn force acting on the \ufb02uid on the left-hand side\nof Equation 9.4.14.\n2.\nThe remaining factor R times to on the right-hand side of Equation 9.4.14 can be replaced by\nthe linear velocity v9.\n3.\nThe factor 1 \u2014f2 can be replaced by 2(1 \u2014f), since 1 \u2014f2=(1 + f)(1 \u2014f) and (1 \u2014l\u2014 f)\u2014>2\n"]], ["block_10", ["Figure 9.9\nDefinition of variables for concentric cylinder viscometers; (a) the rotating cylinder and (b) the\ncoaxial cylinders.\n"]], ["block_11", ["9.5\nDiffusion Coefficient and Friction Factor\n"]], ["block_12", ["d\nUzngs\ncam)\n"]], ["block_13", ["_ =\n=\n_\n.4.15\nA\n0\nnf5\n(9\n)\n"]], ["block_14", ["Introducing these substitutions in Equation 9.4.14 gives\n"]], ["block_15", ["7Q\n_/\",.,,\u2014\\H\nJ\n"]], ["block_16", [{"image_3": "358_3.png", "coords": [138, 43, 355, 249], "fig_type": "figure"}]], ["block_17", [{"image_4": "358_4.png", "coords": [190, 82, 343, 216], "fig_type": "figure"}]]], "page_359": [["block_0", [{"image_0": "359_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "359_1.png", "coords": [29, 324, 148, 363], "fig_type": "molecule"}]], ["block_2", ["Returning to our glass of water, imagine a single dye molecule somewhere in the glass and further\nimagine that we could actually follow its motion directly in time and space. (This is not too far\nfetched, in fact, with recent developments in single-molecule spectroscopy and microscopy).\nWhat would we see? The molecule is constantly buffeted by solvent molecules, and consequently\nexecutes Brownian motion: it moves tiny distances between collisions, with the instantaneous\ndirection of motion fluctuating randomly. If we watch for a relatively long time interval, how far is\nthe dye molecule likely to move? We already have developed all the mathematics needed to\nanswer this question, when we considered random walks in Chapter 6. Namely, if the molecule\nexecutes a total of N random steps of average length b, the mean square displacement (r2) will be\ns. We can formalize this as follows:\n"]], ["block_3", [{"image_2": "359_2.png", "coords": [35, 507, 234, 541], "fig_type": "molecule"}]], ["block_4", ["Equation 6.7.12 can be recovered by substituting 6D,! =n2 into Equation 9.5.2. Also, recall that\nthe prefactor of 41a emerges because we are interested in the distance from the origin, independ-\nent of direction. Unlike the case of polymer chains, which are self\u2014avoiding walks, a Brownian\nparticle really does execute a random walk.\nThe tracer diffusion coef\ufb01cient is a property of an individual molecule or particle undergoing\nBrownian motion. Its value, however, will generally depend on the size of the molecule and on the\nmedium in which it is diffusing. Einstein showed that there is a beautifully simple relationship\nbetween the friction factor and DI:\n"]], ["block_5", ["things: how rapidly the dye molecules move, and the size of the glass. The former is quantified by a\ndz\ufb01\u2018usz\u2019on coe\ufb01\u2018i\u2019cienz, D, which depends directly on the molecular friction factor. Diffusion is an\nimportant process to understand and control in many polymer applications (e.g., how long will\nsoda in fizz? How long does it take a drug in a skin patch to pass through the\nskin?), and the diffusion coefficient itself is also a useful means of molecular characterization. In\nthis section we will emphasize the diffusion of polymers in dilute solution, but the development\nhas a much broader range of applicability. We begin the discussion by distinguishing two related,\nbut conceptually distinct, diffusion coefficients, which we shall call the tracer diffusion coefficient,\nD,, and the mutual diffusion coefficient Dm. Failure to distinguish clearly between these two\nquantities can be a major source of confusion.\n"]], ["block_6", ["The units of the diffusion coefficient are (length)2/time, or cmz/s in the cgs system. We take the\nlimit of long times just to remove all memory of past collisions and directions. We may consider\nthe factor of 6 in front of Dt to be a historical convention. The average denoted by\n(- -) is\nnecessary because by definition we cannot make any predictions about a single random walk.\nRather, if we watch a particular particle for many time intervals or watch many independent but\notherwise identical particles over a given time interval, we can generate the average mean square\ndisplacement. In fact, based on our work with random walks in Chapter 6, we even know the\ndistribution function that should describe the results of many equivalent experiments: it should be a\nGaussian. Just as we did in Chapter 6, we can ask the following question: if we define the location\nof a particle at time t: O as the origin, what is the probability of finding it a distance r away after\ntime t? This probability is known by a special name, the \u201cvan Hove space\u2014time self-correlation\nfunction,\u201d but it is mathematically equivalent to Equation 6.7.12:\n"]], ["block_7", ["9.5.1\nTracer Diffusion and Hydrodynamic Radius\n"]], ["block_8", ["Diffusion Coefficient and Friction Factor\n347\n"]], ["block_9", ["kT\nDI:\u2014\nf\n(9.5.3)\n"]], ["block_10", ["2\n1\n3/2\n1'2\nP\nz =4\n\u2014 \u2014\n.\n.\n(r\u2019 )\nw r (477 DJ)\nexP<\n4Dtt)\n(9 5 2)\n"]], ["block_11", ["2\n2\nlim (_r_) g E 6D,\n(9.5.1)\nI\u2014roo\nZ\nl\u2018\n"]]], "page_360": [["block_0", [{"image_0": "360_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "360_1.png", "coords": [34, 141, 110, 179], "fig_type": "molecule"}]], ["block_2", ["Now we turn to the mutual diffusion coefficient, Dm, which describes how a collection of\nBrownian particles will distribute themselves in space. In the context of the food coloring analogy,\nDt tells us how rapidly any individual dye molecule explores space, but Dm describes how quickly\nthe entire droplet of food coloring disperses itself. Experience tells us that after a reasonable time\ninterval, the glass of water will be uniformly colored. The underlying reason is that mutual\ndiffusion acts to eliminate any gradients in concentration. Although the individual dye molecules\nare happily diffusing about, largely oblivious of one another, collectively they tend to spread\nthemselves out evenly.\nFick first recognized that mass diffusion was analogous to thermal diffusion and he proposed an\nadaptation of Fourier\u2019s law of heat conduction for the transport of material [1 1]. Speci\ufb01cally, the\n\ufb02ux J (in units of mass per unit area per unit time) across a plane is assumed to be proportional to\nthe gradient in concentration (mass/volume) along the direction perpendicular to the plane. We\nwill restrict ourselves to one-dimensional diffusion along the x-direction, and therefore\n"]], ["block_3", ["In other words, Rh for any polymer is the radius of a hard sphere that would have the same friction\nfactor or diffusivity. In Section 9.3 we saw that \ufb02exible molecules behaved hydrodynamically\nrather like hard spheres with radius proportional to Rg. Does this hold for diffusion as well? Yes,\nindeed. Consequently Rh is directly proportional to Rg, with the proportionality factor depending\non the particular polymer shape. Thus, Rh depends on molecular weight with the same power law\nexponent as does Rg, and Dt exhibits the inverse of that dependence. For \ufb02exible chains, therefore,\nDv\u2018m~ in a theta solvent and Dv\u2018y5 in a very good solvent. Examples are shown in\nFigure 9.10 for polystyrene in cyclohexane at the theta temperature, and in toluene, a good solvent.\nThere are several experimental techniques by which diffusion may be measured, and examples will\nbe given following the next section.\n"]], ["block_4", ["In this expression, called Fick\u2019s first law, the proportionality constant is Dm and it follows that Dm\nalso has units of length2/time. The minus sign in Equation 9.5.6 recognizes the fact that the\ndirection of the \ufb02ow is that of decreasing concentration.\nWe now consider a volume element and the \ufb02ux of solute in and out of that element. Figure 9.1 1\nschematically represents three regions of an apparatus containing a concentration gradient. The end\n"]], ["block_5", ["This simple relation provides a direct connection between the tracer diffusion coefficient, the\nparticle size, and the viscosity of the solvent. We can go an important step further, however. If our\nparticle is not a hard sphere, we can use Equation 9.5.4 to define an equivalent radius in terms of a\nmeasured diffusivity, called the hydrodynamic radius, Rh:\n"]], ["block_6", ["From the definition of the friction factor, we can see that the thermal energy, H\", is playing the role\nof a generalized \u201cforce,\u201d and Dt is the resulting \u201cvelocity.\u201d As T increases, solvent molecules\nmove more rapidly and are more effective at jostling the tracer molecule, so Dt should increase. At\nthe same time, if the solvent viscosity increases, or if the particle increases in size, thenf should\nincrease and Dt will be smaller. If we now consider our tracer molecule to be a hard sphere and the\nsolvent to be a continuum, then we can incorporate Stokes\u2019 law forf (Equation 9.2.4) to arrive at\nthe Stokes\u2014Einstein relation\n"]], ["block_7", ["343\nDynamics of Dilute Polymer Solutions\n"]], ["block_8", ["9.5.2\nMutual Diffusion and Fick's Laws\n"]], ["block_9", [{"image_2": "360_2.png", "coords": [37, 229, 115, 264], "fig_type": "molecule"}]], ["block_10", ["dc\nJ \nDlm dx\n(9.5.6)\n"]], ["block_11", ["D \nkT\nt_6*n\"nSR\n(9.5.4)\n"]], ["block_12", ["6771?, Dt\n"]], ["block_13", ["kT\n"]]], "page_361": [["block_0", [{"image_0": "361_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "361_1.png", "coords": [26, 47, 265, 447], "fig_type": "figure"}]], ["block_2", [{"image_2": "361_2.png", "coords": [32, 40, 289, 262], "fig_type": "figure"}]], ["block_3", [{"image_3": "361_3.png", "coords": [32, 273, 285, 445], "fig_type": "figure"}]], ["block_4", ["Figure 9.10\nMolecular weight dependence of DI for polystyrene in (a) a theta solvent, cyclohexane at 35\u00b0C,\nand (b) a good solvent, toluene. The curve in (a) has a slope of \u2014~0.50, and the curve in (b) approaches a slope\nof \u20140.60 at large M. (Reproduced from Schaefer, D.W. and Han, C.C., Dynamic Light Scattering, R. Pecora\n(Ed), Plenum, New York, 1985. With permission.)\n"]], ["block_5", ["compartments contain the solute at two different concentrations (:1 and (:2, with C2 > CI. The center\nregion is a volume of cross\u2014sectional area A and thickness (13:, along which the gradient exists. The\narrows in Figure 9.11 represent the flux of solute from the more concentrated solutions to the less\nconcentrated one. The incremental change in the total amount of solute dQ in the center volume\nelement per time increment (it can be developed in two different ways. In terms of the \ufb02uxes,\n"]], ["block_6", ["Diffusion \n349\n"]], ["block_7", ["whereas in terms of a concentration change dc in the element of volume A dx,\n"]], ["block_8", ["~99.\n(\\I\nE\n0.\n"]], ["block_9", ["QC\nh-\n$3\n1.0\n"]], ["block_10", ["CUE\n"]], ["block_11", ["c5\u201d\nrx\n-\n._\n$2\n1.0\u2014\n__\n'_\n-\n"]], ["block_12", ["10.0I\u2014\n'\u2014\n33\n_\n_\n"]], ["block_13", ["0.\n_\n_\n"]], ["block_14", [{"image_4": "361_4.png", "coords": [41, 601, 132, 634], "fig_type": "molecule"}]], ["block_15", ["100-0\nlill|\n1\nIII]\nI IIIrI\nI\u2014II]\nI\n1H\n"]], ["block_16", ["100-0\nI\u2014IITI\nI\nI\u2014rII\nI\nil||\nI\nI\u2018ll\n"]], ["block_17", ["dQ\nE; \n-\u2014 10mm\n(9.5.7)\n"]], ["block_18", ["dQ = dc(A dx)\n(9.5.8)\n"]], ["block_19", ["10.0\n"]], ["block_20", ["0.1\nI\nIIII\n|\nIIII\nI\nI_LJ_L_L_LJ_I\n"]], ["block_21", ["L\nl\n0_1\n_I_J_1LLJ_llii\nI_IIII\nI_LJ_IJI\nI\n"]], ["block_22", ["1\n10\n100\n1000\n"]], ["block_23", ["1\n10\n100\n1000\n10,000\n"]], ["block_24", ["J\nPSintqene\n:\n"]], ["block_25", ["\u2014\n.\u2014\n"]], ["block_26", ["l\u2014\n.\u2014\n"]], ["block_27", ["PS in cyclohexane\n\u201dJ\n"]], ["block_28", ["10\u20143MW\n"]], ["block_29", ["10_3Mw\n"]]], "page_362": [["block_0", [{"image_0": "362_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "362_1.png", "coords": [25, 32, 267, 188], "fig_type": "figure"}]], ["block_2", [{"image_2": "362_2.png", "coords": [32, 519, 130, 553], "fig_type": "molecule"}]], ["block_3", [{"image_3": "362_3.png", "coords": [33, 628, 299, 673], "fig_type": "molecule"}]], ["block_4", ["This differential equation can be solved (sometimes easily, sometimes not) when one speci\ufb01es the\nappropriate initial or boundary conditions (see Example 9.4 and Example 9.5).\nFick\u2019s second law is very useful, but it leaves us with two unresolved issues. First, how is Dm\nrelated to Dt and f? Second, what is the role of thermodynamics in this process? These questions\ncan be answered together. In general, a single phase will be at equilibrium when the temperature,\npressure, and chemical potential of each species are everywhere the same. In other words, diffusion\nactually acts to remove gradients in p. rather than gradients in c. The thermodynamic \u201cdriving\nforce\u201d can be written (recall that the gradient of a potential is a force) as\n"]], ["block_5", ["The areas cancel and the term in square brackets can be recognized as\n"]], ["block_6", ["We divide Equation 9.5.12 by Avogadro\u2019s number to convert the partial molar Gibbs free energy,\nu, to a molecular quantity and the minus sign enters as in Fick\u2019s first law because the force and the\ngradient are in opposing directions. Recalling the definition of chemical potential from Equation\n7.1.13 we write #2 p3 +RTlnag p33 +RTln 720, where (.12 and y; are the activity and\nactivity coefficient, respectively, of the solute (note that when 6 is in g/mL, 'yz is defined\naccordingly). Substituting into Equation 9.5.12 we obtain\n"]], ["block_7", ["Figure 9.11\nSchematic of diffusion with respect to a volume element of thickness dx located at x =0\nbetween regions with concentration cl and c; > cl.\n"]], ["block_8", ["that leads us directly to Fick\u2019s second law (in one dimension):\n"]], ["block_9", ["350\nDynamics of Dilute Polymer Solutions\n"]], ["block_10", ["By combining these two relations, and substituting Fick\u2019s first law for J, we obtain\n"]], ["block_11", ["l</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n.</</sub>sub>\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n-r</</sub>sub>-\u2014>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n\u2014'I</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n02</</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n01</</sub>sub>\n\u2014|\u2014'|\nl</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n\u2014l</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>\n[J______3J- 14\u2014</</sub>sub> \nr</</sub>sub>\nx</</sub>sub></</sub>sub>\nx</</sub>sub></</sub>sub>\nx</</sub>sub></</sub>sub>\u201d\nz/</sub>\n\u00ab1/</sub></</sub>sub>\nz!</</sub>sub>\nx</</sub>sub></</sub>sub>/</sub>\nx</</sub>sub></</sub>sub>/</sub>\nx</</sub>sub></</sub>sub>/</sub>\nx</</sub>sub></</sub>sub>/</sub>\nz/</sub>\n/</sub>\nI</</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub></</sub>sub>f\n/</sub>\nx</</sub>sub></</sub>sub>=0\nx</</sub>sub></</sub>sub>=dx</</sub>sub></</sub>sub>\n"]], ["block_12", [{"image_4": "362_4.png", "coords": [42, 329, 186, 365], "fig_type": "molecule"}]], ["block_13", ["1\nduz\ndc\nRT\n1\nd In 72\ndc\nF i\n= _\n_\n_\n= _\n_\n_____\n\u2014\n5.13\nd H\nNa, ( dc ) (dx)\nNa, (0 \ndc\ndx\n(9\n)\n"]], ["block_14", ["Nav\ndx\n(9.5.12)\nFdiff \n"]], ["block_15", ["E (A dx) = (J(x) Jo + dx))A\ndt\ndc\ndc\n= \u201cDmKal\u2018(5)....1/4\n(9.5.9)\n"]], ["block_16", ["dc\ndzc\n"]], ["block_17", ["dt\n\u2014Dm\ndxz\n(9.5.11)\n"]], ["block_18", ["2\nIa \u00abaka\nx+dx\nx\n"]], ["block_19", ["1\nd\ufb02vz\n"]]], "page_363": [["block_0", [{"image_0": "363_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "363_1.png", "coords": [29, 370, 328, 413], "fig_type": "molecule"}]], ["block_2", [{"image_2": "363_2.png", "coords": [30, 134, 179, 170], "fig_type": "molecule"}]], ["block_3", ["Experimentally, a variety of tools can be used to determine either Dt or Dm. Probably the most\ncommonly employed technique for dilute polymer solutions is dynamic light scattering, also\nknown as quasielastic light scattering, which we will brie\ufb02y describe in Section 9.6. Another\ncommon technique for solutions, which we will not discuss, is pulsed-field gradient NMR. It\nmeasures the random motions of particular nuclear spins without regard to any gradients in\nconcentration, and thus determines Dt. A more generic approach, often applied to less \ufb02uid\nsamples, is to prepare adjacent layers of two different compositions, and watch them interdiffuse\nby some depth-pro\ufb01ling technique. This is illustrated in the following example.\n"]], ["block_4", ["In many experiments to study polymer diffusion, some \ufb01xed, small amount of the polymer is placed\nin contact with the medium into which it will diffuse. This medium, or matrix, could be a solvent,\nanother polymer, or a solution. Solve Fick\u2019s second law (Equation 9.5.11) for the one-dimensional\ncase where an extremely thin layer of polymer is placed in an \u201cin\ufb01nite\u201d beaker of solvent.\n"]], ["block_5", ["In fact, we have already almost solved this problem in Equation 9.5.2. That case concerned three-\ndimensional diffusion from the origin, whereas now we have one-dimensional diffusion from a\nplane. Nevertheless, the solution is still a Gaussian function, just as in Section 6.7, where we\nshowed that the end-to-end distance of a random walk was a Gaussian function in x, y, or z.\n"]], ["block_6", ["In the limit of low concentration of solute, c \u2014\u2014> 0, we see that Dm \u2014+ Dm, where the subscript \u201c0\u201d\nrefers to this \u201cinfinite dilution\u201d limit. This is a completely general result for any two-component\nmixture: the mutual diffusivity approaches the tracer diffusivity of the minor component as its\nconcentration tends to zero. At \ufb01nite concentrations, however, things are not so simple. Clearly we\ncan define tracer diffusion coefficients for both polymer and solvent and these may be very\ndifferent. On the other hand, there is only one mutual diffusion coefficient, but it need not bear\nany simple relation to the tracer diffusivities. In Equation 9.5.16 we should also consider the\npossibility that the solute friction factor, f, will depend on concentration. If we propose a series\nexpansion, that is,\n"]], ["block_7", ["and insert this into Equation 9.5.16, we would obtain\n"]], ["block_8", ["The product cvo de\ufb01nes the \ufb02ux J, and therefore Equation 9.5.14 becomes\n"]], ["block_9", ["Solution\n"]], ["block_10", ["Comparing Equation 9.5.6 and Equation 9.5.15 gives the desired result\n"]], ["block_11", ["Under stationary-state \ufb02ow conditions, Fdiff equals the force of viscous resistance experienced by\nthe particle. The latter, in turn, equals the friction factor times the stationary velocity v0; therefore\n"]], ["block_12", ["Example 9.4\n"]], ["block_13", ["Diffusion Coefficient and Friction Factor\n351\n"]], ["block_14", [{"image_3": "363_3.png", "coords": [41, 188, 252, 222], "fig_type": "molecule"}]], ["block_15", [{"image_4": "363_4.png", "coords": [42, 84, 193, 121], "fig_type": "molecule"}]], ["block_16", [{"image_5": "363_5.png", "coords": [44, 377, 254, 407], "fig_type": "molecule"}]], ["block_17", ["f(c)=fo{l +k+\"'}\n(9.5.17)\n"]], ["block_18", ["H\"\nd In 3/2\ndc\nJ \nl +\n\u2014\n.\n.1\nf\n<\n6\ndc\n)dx\n(9 5\n5)\n"]], ["block_19", ["kT\nd In \u2018yz\nd In \u2018yz\nDm=\u2014\u2014\n1+\n-\u2014-D\n1\n\u2014k\n9.5.18\nfo(l + k) (\nC\ndc )\n[\u2018\u00b0( C{\ndc\nf\n(\n)\n"]], ["block_20", ["Id\"\n(1 ln 3/2\n(1 ln 'yz\nDm \nl\n= D\nl\n\u2014\n.\n.\nf \ndc )\nt.o( +6\ndc\n(9516)\n"]], ["block_21", ["la?\"\n(1 ln 3/2\ndc\n_ __\n1\n._\n=\n.\n.\nc< +c\ndc \nfvo\n(9514)\n"]], ["block_22", [{"image_6": "363_6.png", "coords": [144, 377, 311, 409], "fig_type": "molecule"}]]], "page_364": [["block_0", [{"image_0": "364_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "364_1.png", "coords": [22, 386, 275, 458], "fig_type": "figure"}]], ["block_2", [{"image_2": "364_2.png", "coords": [29, 562, 254, 614], "fig_type": "molecule"}]], ["block_3", [{"image_3": "364_3.png", "coords": [32, 626, 176, 666], "fig_type": "molecule"}]], ["block_4", [{"image_4": "364_4.png", "coords": [33, 387, 266, 451], "fig_type": "molecule"}]], ["block_5", ["where we use the expression for the solution to the integral of a Gaussian given in Section 6.7.\nIn real experiments, it is more common to have a layer of finite thickness of material in contact\nwith the effectively infinite matrix. For simplicity, we center this layer at x 0 and it extends from\nx \u20141/2 to x +l/2, as illustrated in Figure 9.12b. This new initial condition changes the solution\nto Fick\u2019s law. The approach to the solution is to view the layer of thickness h as a series of\ninfinitesimal layers and then to calculate the amount of material at each point x as a sum of the\nmaterial that came from each of the infinitesimal layers. We already have the solution for one\ninfinitesimal layer, and the result for the \ufb01nite layer is\n"]], ["block_6", ["which must be equal to Dm(3zc/8x2). Now take the derivative twice with respect to x:\n"]], ["block_7", ["In order to show that this is a solution, we can follow through with the differentiation:\n"]], ["block_8", ["which is the same as the expression for 60/8: above.\nThe character of this solution is shown in Figure 9.12a. The polymer concentration begins as a\nspike at x O, and then as time progresses, it spreads out symmetrically to positive and negative\nvalues of x. The constant 00 is the total amount of polymer in the initial spike, as can be seen from\nthe following argument. If we integrate 60:, I) over all x at any time, we must get the total amount\nof polymer:\n"]], ["block_9", ["and\n"]], ["block_10", ["where erf denotes the error function\n"]], ["block_11", ["352\nDynamics of Dilute Polymer Solutions\n"]], ["block_12", ["and the solution is\n"]], ["block_13", [{"image_5": "364_5.png", "coords": [39, 223, 276, 277], "fig_type": "molecule"}]], ["block_14", [{"image_6": "364_6.png", "coords": [39, 68, 152, 104], "fig_type": "molecule"}]], ["block_15", [{"image_7": "364_7.png", "coords": [39, 120, 197, 159], "fig_type": "molecule"}]], ["block_16", [{"image_8": "364_8.png", "coords": [41, 277, 330, 320], "fig_type": "molecule"}]], ["block_17", ["(96\n1\nex\nx2\n1 [\u20141 +\nx\n[\u20142\n\u2014 : C0 -\u2014\u2014\u2014\n\u2014 \u2014\u2014\n-\u2014 -\u2014\n\u2014\u2014\na:\n#4m!\np\n40\u201c,:\n2\n40,,\n"]], ["block_18", ["D\n36\nD\n1\nex\nx2\n2x\n"]], ["block_19", ["D\n"]], ["block_20", ["erf(z) :\nJ du exp (\u2014 142)\n"]], ["block_21", ["1\nx+h\nx\u2014h\n0\u201c\" 0 \ufb02ail\u2014m) +Edi/\ufb01ll\n"]], ["block_22", ["For Fick\u2019s second law in this case\n"]], ["block_23", ["80(x,t)_\ufb01 D\n620050\nat\n_\nm\n83:2\n"]], ["block_24", ["60:, I) co\n1\nx2\n)\n,/\u2014\u20144womz\np\n40...:\n"]], ["block_25", ["00\nDO\nJ\ndx\n(\nr)\n1\nJ\ndx\nx2\nex,\n:0 \nex\n\u2014\n\u00b0\n4770,\u201c:\np\n40m:\n-00\n\u2014oo\n"]], ["block_26", ["m\nH\u2014 :\nmCO -\u2014\u2014\u2014\u2014\n_-_ \u2014\u2014\n_\n6x\n\u00ab4m:\np\n40m:\n40m:\n"]], ["block_27", ["82C\n1\nx2\n2\n2x\n2x\nm\u2014ZDmCo\u2014exp \n'_\n+\n(9x2\nx/47'rDmt\n4Dmt\n4Dmr\n40m: 4Dmr\n"]], ["block_28", [{"image_9": "364_9.png", "coords": [77, 229, 258, 262], "fig_type": "molecule"}]], ["block_29", ["\u00a7|~ 0\n"]], ["block_30", [{"image_10": "364_10.png", "coords": [83, 177, 278, 210], "fig_type": "molecule"}]], ["block_31", ["1\n:c \u2014~\u2014\u2014-\u2014\u2014\\/471'Dmt=c\n0\n47mm:\n\u00b0\n"]], ["block_32", [{"image_11": "364_11.png", "coords": [168, 394, 262, 434], "fig_type": "molecule"}]]], "page_365": [["block_0", [{"image_0": "365_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The error function is tabulated in many standard mathematical references. The solution for C(x, t)\nis also shown in Figure 9.12b and to a first approximation it looks rather like the result for a\nsingle layer.\nIn an actual diffusion experiment, the geometry is usually slightly different, namely a single thin\nlayer of material is placed on one thick layer of matrix. However, with a little thought it should be\napparent that the answer is the same as above, by symmetry. The sum of error functions is\nsymmetric with respect to x>0 and x<0 (note that erf(z)=erf(\u2014z)), and we could imagine\n"]], ["block_2", ["Figure 9.12\nTime evolution of concentration profiles for (a) an infinitesimal layer of material at x = 0,\n\u00a320, and (b) a layer of thickness 12 centered on x = 0 at t = 0. The corresponding mathematical forms for\n60:, t) are discussed in Example 9.4.\n"]], ["block_3", ["Diffusion and Friction Factor\n"]], ["block_4", ["0\n(x,\nt)\n"]], ["block_5", ["t)\n"]], ["block_6", ["C(X,\n"]], ["block_7", [{"image_1": "365_1.png", "coords": [41, 47, 271, 273], "fig_type": "figure"}]], ["block_8", ["0.2\n"]], ["block_9", ["0.1\n"]], ["block_10", ["0.3\n"]], ["block_11", ["0.4\n"]], ["block_12", ["0.2\n"]], ["block_13", ["0.6\n"]], ["block_14", [{"image_2": "365_2.png", "coords": [47, 279, 260, 504], "fig_type": "figure"}]], ["block_15", ["l. o\n"]], ["block_16", ["l'\nI\nI\nI\n"]], ["block_17", ["I\nl\n"]], ["block_18", ["I\nI\nI\nI\nI\n"]], ["block_19", ["ITTr\u2014IIII\n"]], ["block_20", ["IU'I'I\u2014lIrI\n"]], ["block_21", ["'I\nI\n"]], ["block_22", ["I\nI\nl\n"]], ["block_23", ["I\nI\nI\nl\nl\nI\n"]], ["block_24", ["*rrrllill\n\"\u20185\u2018!\n"]], ["block_25", [{"image_3": "365_3.png", "coords": [179, 339, 237, 384], "fig_type": "figure"}]], ["block_26", ["ILILLIIIIIIIIJ.\n"]], ["block_27", ["1\n"]], ["block_28", ["Ill!\n"]], ["block_29", ["353\n"]]], "page_366": [["block_0", [{"image_0": "366_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "366_1.png", "coords": [25, 325, 142, 362], "fig_type": "molecule"}]], ["block_2", [{"image_2": "366_2.png", "coords": [30, 564, 190, 598], "fig_type": "molecule"}]], ["block_3", ["This function involves taking the intensity at some time t\u2019, multiplying it by the intensity an\ninterval I later, and adding the results up over some long interval T. By then dividing by T, the\nlength of time of the experiment is taken out of the answer. To obtain the full C(I), this process\nneeds to be repeated for many different values of the interval t. In the actual experiment, the\nscattered intensity is digitized and stored as a string of numbers, which are then manipulated by a\nspecial computer called a correlator, to generate the digital version of C(t).\n"]], ["block_4", ["placing an impenetrable barrier at x=0. In this case no material starting between x: \u2014h/2 and\nx 0 would appear at x > 0, but an equal amount that started between x 0 and x 12/2 would not\nbe \u201clost\u201d to x < 0. The value of DU, would be obtained by allowing diffusion to occur for some\napprOpriate time interval, and then using some kind of depth profiling experiment capable of\nmeasuring C(x). The resulting concentration profile could be fit to the solution given above, to\nextract Dm.\n"]], ["block_5", ["From this relation it appears that, if we could measure the time decay of concentration \ufb02uctuations,\nwe could measure Dm.\nThe complication arises because there is no special time t: 0; \ufb02uctuations rise and decay all the\ntime. However, the necessary information is there in the light scattering signal, if we measure the\ntemporal \ufb02uctuations in [5, rather than the time average value as we did in Chapter 8. A schematic\nexample of [5(1\u2018) is shown in Figure 9.13a. The instantaneous value of IS bounces around the average\nvalue, as the scattering molecules move in solution and thereby alter the exact phase relations\namong the waves scattered from different polymers. Although 15(t) looks like noise, it actually has\nin it a typical time constant, which corresponds to the correlation time over which the signal loses\nmemory of whether it was, say, larger than average or smaller than average. (In a sense, this\ncorrelation time plays the same role as the persistence length of a random walk discussed in detail\nin Chapter 6.) T0 extract this time constant, the scattered intensity is analyzed via a time\nautocorrelation function, C(I), defined as follows:\n"]], ["block_6", ["for which the solution is (check it yourself):\n"]], ["block_7", ["We recall from the discussion in Section 8.2 and Section 8.4 that light scattering in a dilute\npolymer solution arises from \ufb02uctuations in concentration (e.g., see Equation 8.4.4). However,\neach \ufb02uctuation must appear and disappear over some time interval and this time interval is\ndetermined by Dm. As a consequence, the total scattered intensity also \ufb02uctuates in time and these\n\ufb02uctuations may be analyzed to extract a characteristic relaxation time, 1'. The magnitude of the\nscattering vector, q (defined in Equation 8.2.4), sets the relevant length scale in solution to be l/q\n(see Section 8.6) and the relaxation time 1' turns out to be equal to 1/(qm).\nFrom Section 8.2 and Section 8.6 we remember that the only \ufb02uctuations that will contribute to\nthe scattered intensity IS are the \ufb02uctuations that happen to have period 277/q and that are oriented\nalong the scattering vector, in other words the spontaneous \ufb02uctuations that satisfy the Bragg\ncondition. Suppose that such a \ufb02uctuation occurs at some time we designate :2 O; we could write it\nas 600\u2018: 0) 2A cos(qx), where A is its amplitude and x is the appropriate direction. This\n\ufb02uctuation would now relax according to Fick\u2019s second law (Equation 9.5.11):\n"]], ["block_8", ["354\nDynamics of Dilute Polymer Solutions\n"]], ["block_9", ["9.6\nDynamic Light Scattering\n"]], ["block_10", ["T\nC(t) E Tlim %J I,(z\u2019)1,(r\u2019+r) dt\u2019\n(9.6.3)\n20..\n0\n"]], ["block_11", ["5cm :5c(0) e\u2018qZD\u2018\u201c A cos(qx) e\u2018qm\u2019\n(9.6.2)\n"]], ["block_12", ["3(56)\n82(56)\nat \nm 7;?\u2014\n(9.6.1)\n"]], ["block_13", [{"image_3": "366_3.png", "coords": [49, 330, 128, 356], "fig_type": "molecule"}]]], "page_367": [["block_0", [{"image_0": "367_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "367_1.png", "coords": [22, 40, 347, 193], "fig_type": "figure"}]], ["block_2", ["The first term is a constant and the next two terms integrate to zero because the \ufb02uctuations are\nequally likely to be positive or negative. So, once again, it is the square of the \ufb02uctuations that\ncontain all the interesting information. Now consider the two limits, of the time interval 1: very small\nand very large, compared to the timescale for molecular motion (1/q2Dm). When 2\u2018 is very small, the\nmolecules do not move between t\u2019 and t\u2019 + r, so IS will not change. Thus 51(1\u2019) % 513(t\u2019 + t) and thus\nthe integrand is (513(t\u2019))2. This is always positive and therefore C(t) has its maximum (2 ((Is(t))2))\nwhen r 0. On the other hand, when r is very large, the molecules have completely randomized their\npositions and so 513(1\u2019) and 5130\u2019 + t) have no correlation. Consequently, their product is equally\nlikely to be positive or negative and the integral will be zero. Thus in this limit C(z\u2018) reduces to the\nconstant (If. The functional form of C(r) in the intervening region turns out to be an exponential,\njust like the solution to Fick\u2019s law above, as shown in Figure 9.13b:\n"]], ["block_3", ["The additional factor of 2 in the exponential of Equation 9.6.4 arises because the motion of the\nmolecules results in the loss of correlation of the scattered electric \ufb01eld, E, and the intensity is\nproportional to the square of the field.\nIn the actual experiment, the range of t can be adjusted to match the system under study over a\nvery wide interval (about 100 ns to 100 s) (see Example 9.5). A measurement of C(t) at any\nscattering angle is sufficient to extract a value of Drn by Equation 9.6.4, but the result will be more\nreliable if measurements are taken at several angles and the experimental time constants r are\nplotted as 1/7- versus qz. The data should follow a straight line, with zero intercept and slope equal\nto Dr\u201c. The values of D\",1 for solutions of low concentration can be extrapolated to zero concen-\ntration and thereby the tracer diffusion coefficient Dt can be obtained (Equation 9.5.16). The data\nin Figure 9.10 were obtained by this method. Finally, use of the Stokes\u2014Einstein relation (Equation\n9.5.5) gives access to Rh. This is the basis of the common use of dynamic light scattering for\nparticle sizing. It is also worth noting that this method can be used for quite small values of Rh,\napproximately 1 nm, in contrast to light scattering, which can only determine Rg when Rg is greater\nthan about 10 nm.\n"]], ["block_4", ["Figure 9.13\n(a) Scattered intensity as a function of time, with data digitized every interval t. (b) Intensity\nautocorrelation function C(t) as a function of I.\n"]], ["block_5", ["What does C(2\u2018) look like? The instantaneous 13(t) can always be written as the sum of its average\nand a \ufb02uctuation:\n13(1\u2018) (13) + 5130\u2018) (recall how we used the same strategy in treating the\npolarizability in Chapter 8). The integrand of C(t) can be written as\n"]], ["block_6", ["Dynamic Light Scattering\n355\n"]], ["block_7", ["(a)\nt'\n(b)\nt\n"]], ["block_8", [{"image_2": "367_2.png", "coords": [37, 454, 212, 485], "fig_type": "molecule"}]], ["block_9", [{"image_3": "367_3.png", "coords": [38, 48, 231, 169], "fig_type": "figure"}]], ["block_10", ["C(r) (<13) <13>2)e\u20182q20m\u2019+ <1.)2\n(9.6.4)\n"]], ["block_11", ["(1,)2 + 51,000,) + 51,(z\u2019+z)(1,) + 51,(z\u2019)51,(z\u2019 + r)\n"]], ["block_12", [{"image_4": "367_4.png", "coords": [52, 90, 227, 142], "fig_type": "molecule"}]], ["block_13", [{"image_5": "367_5.png", "coords": [226, 25, 439, 204], "fig_type": "figure"}]]], "page_368": [["block_0", [{"image_0": "368_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We conclude this section on dynamic light scattering by emphasizing one further point about\ndiffusion. Compare the three solutions to the same equation, Fick\u2019s second law, in Example 9.4\nand in Equation 9.6.2. In the former example, c(x,t) has in one instance a very complicated\ndependence on time and a Gaussian form in space, and in the other, a rather obscure answer\ninvolving error functions. In contrast, from Equation 9.6.2 c(x,t) decays exponentially in time\nand follows an oscillatory function in space. How can the same process, diffusion, and the same\nequation, Fick\u2019s second law, give such different results? The answer is that the solution depends\ncritically on the initial conditions and the boundary conditions. In Equation 9.6.2 the initial\ncondition was a cosine wave and that particular function is preserved by Fick\u2019s law; in other\nwords the x dependence of C(x, t) is unchanged and only the amplitude of the wave decays with\ntime. The initial conditions in Example 9.4 were different and so were the results. In fact, there\nare numerous practical situations with a variety of different constraints and many different\nfunctional forms for c(x,t) can result. An excellent discussion of many of these cases can be\nfound in the text by Crank [12].\n"]], ["block_2", ["These values are well within the range of commercial correlators.\n"]], ["block_3", ["We have emphasized both here and in Section 8.6 that 1/q is a length, and by taking the reciprocal\nof these numbers we can see that the typical values for light scattering fall in the range 30\u2014120 nm.\nThe values of Drn can be estimated by assuming that the latex particles are sufficiently dilute so\nthat Drn % DD and using the Stokes\u2014Einstein relation (Equation 9.5.4). For a hard sphere, remember\nthat Rh R.\n"]], ["block_4", ["The time constants are determined by the product of q2 and Dr\u201c, so we need to estimate both. For q,\nlet us assume the light source is an argon ion laser operating at 488 nm and that the instrument can\naccess scattering angles, q, from 30\u00b0 to 150\u00b0. The refractive index of water is about 1.33 (see Table\n8.1). From Equation 8.2.4\n"]], ["block_5", ["Estimate the range of time constants, 7, which would be extracted from dynamic light scattering\nmeasurements of dilute aqueous suspensions of latex particles ranging in size from 10 nm to 1 pm.\n"]], ["block_6", ["In this calculation we have used cgs units, with the viscosity of water estimated to be 0.01 P. In the\nfinal calculation of 7 we will need to convert these numbers to as, which brings in a factor of\n"]], ["block_7", ["356\nDynamics of Dilute Polymer Solutions\n"]], ["block_8", ["Solution\n"]], ["block_9", ["Example 9.5\n"]], ["block_10", ["101\u20184 (a/cmz). The results are\n"]], ["block_11", ["47m\nsin\n6\n4 x 3.14 x 1.33\nsin\n30\nto s'n\n150\n=\n\u2014\n2\n\u2014\n1\n\u2014-\u2014\nq\nA0\n2\n488\n2\n2\n= 0.0089 to 0.033 nm\u20181\n"]], ["block_12", ["D:\n"]], ["block_13", ["@6X10\u20142s\nTmax \n"]], ["block_14", ["w4><10_5 s\n7min\n"]], ["block_15", ["_\nkT\n__\n1.4 ><10\u2018l6 x 300\n\u2014\n67rnSRh \n6 x 3.14 x 0.01 x (10-6 to 10-4)\n2 2.2 X 10\"7 to 2.2 X 10'9 cmZ/s\n"]], ["block_16", ["1\n=\n2.2 x 10-7 x 1014 x (0.033)2\n"]], ["block_17", ["2.2 x 10-9 x 1014 x (0.0089)2\n"]], ["block_18", [{"image_1": "368_1.png", "coords": [90, 178, 320, 229], "fig_type": "molecule"}]], ["block_19", ["1\n"]]], "page_369": [["block_0", [{"image_0": "369_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "369_1.png", "coords": [23, 464, 222, 635], "fig_type": "figure"}]], ["block_2", [{"image_2": "369_2.png", "coords": [27, 483, 216, 580], "fig_type": "molecule"}]], ["block_3", ["In this section we address that has played an important role in interpreting intrinsic\nviscosity and the diffusivity: why does treating a \ufb02exible coil as a hard sphere with R 2&1n\nreproduce the experimental results forf (and therefore [77] and 0,)? Before providing the qualita-\ntive answer, it is worth taking a moment to emphasize why this result is surprising. A detailed\nmodel for the dynamics of a \ufb02exible chain, called the bead\u2014spring or R0use\u2014\u2014Zimm model, will be\ndiscussed in Chapter 11. Its essence is to replace the real polymer with a freely jointed sequence of\nN beads connected by elastic springs of average length b, where a particular bead\u2014spring unit can\nbe thought of as representing a persistence length or two (see Figure 9.14). The springs resist\ndeformation of the chain and act to restore random coil conformations perturbed by flow, but\n"]], ["block_4", ["This simple argument leads to the conclusion that fM, and therefore Dtrv l/M, whereas the\nexperiments clearly show fIi\"g NM\u201d, andt l/M\" (see Figure 9.10). What is wrong with this\nargument? The answer is that it neglects the phenomenon known as intramolecular hydrodynamic\ninteractions (HI for short).\nThe underlying idea is illustrated schematically in Figure 9.14. When any bead moves through\nthe solvent, it sends out a ripple, or wave, that is felt by every other bead. The amplitude of this\nwave dies off only as l/rij, the distance between beads i and j, which makes it a rather long-ranged\ninteraction. (Recall from Chapter 8 that the electric field around a point charge and the amplitude\nof the scattered electric field from a single polarizable object both fall off as l/r as well, whereas\nthe short-ranged van der Waals attractions discussed in Section 7.6 fall off as 1/r6.) Furthermore HI\nis very complicated to handle mathematically because to compute the effect on each bead at any\ninstant, we need to sum the contributions from every other bead on the chain and therefore we need\n"]], ["block_5", ["are otherwise \u201cinvisible.\u201d Each bead encounters frictional resistance as it moves relative to the\nsolvent. The bead friction factor, g, is a parameter of the model but it should correspond to the net\nfriction of a few real monomer units. In particular, the number of beads N is proportional to the\ndegree of polymerization of the real chain; for example if N:M/SMO, or five monomers per bead\u2014\nSpring unit, then 4\u201c should correspond to the friction of five monomers. This model is extremely\nsuccessful in many respects, as we shall see in the next chapter, but for now let us focus on the\nchain friction factor, f. As the chain moves through the solvent, the total friction should just be the\nsum of the friction experienced by each bead:\n"]], ["block_6", ["9.7\nHydrodynamic Interactions and Draining\n"]], ["block_7", ["Hydrodynamic Interactions and Draining\n357\n"]], ["block_8", ["Figure 9.14\nSchematic of a bead\u2014spring chain with hydrodynamic interactions.\n"]], ["block_9", ["f:N\u00a7\n(9.7.1)\n"]], ["block_10", [{"image_3": "369_3.png", "coords": [92, 580, 171, 630], "fig_type": "molecule"}]]], "page_370": [["block_0", [{"image_0": "370_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where the relation to Rg is also included. The key result is that by including HI in this way, the\nchain friction factor, f, is found to be given by 67mSRh, and not by N;.\nThe preceding paragraphs only hint at the rather elaborate mathematical machinery required,\nand the extension of the Kirkwood\u2014Riseman theory to an expression for the intrinsic viscosity is\neven more involved. Consequently, it can be helpful to view these results in a more qualitative,\nphysical way by invoking the concept of draining. This is illustrated in Figure 9.15a and Figure\n9.15b. In both panels, a \ufb02exible chain is shown, with the streamlines of the surrounding solvent\nfollowing some imposed flow. We switch back to the perspective of the \ufb02owing solvent and the\nfrictional resistance that the chain presents, rather than the friction experienced by the chain, but\nof course these must be equivalent, as in the discussion of Stokes\u2019 law in Section 9.2. In panel\n(a), the solvent streamlines pass right through the coil, a limiting behavior termed freely\ndraining. In this case the total friction offered by the chain will be NC and thus fM. In panel\n(b), the streamlines are totally diverted around the periphery of the coil, just as in the hard sphere\ncase (see Figure 9.3). This extreme of behavior is called nondraining, and the friction varies as\nfRh M\u201d. We can now see that the effect of HI is to make the coil essentially nondraining. The\nbeads on the upstream edge of the coil, which the solvent encounters first, shield the other beads\nfrom the solvent and thus the net friction is reduced. If we revert to the frame of reference of the\npolymer, when one bead moves in a certain direction, the cumulative effect of HI is to \u201cnudge\u201d\n"]], ["block_2", ["where again we omit the details of the evaluation of the double sum. The Kirkwood\u2014Riseman\nprediction for the hydrodynamic radius of a Gaussian chain (i.e., in a theta solvent) is finally\nobtained by inverting Equation 9.7.3:\n"]], ["block_3", ["where we omit the details of how to evaluate this integral. The hydrodynamic radius, defined in\nEquation 9.5.5, is obtained from the average of Equation 9.7.2 over all pairs of beads on the chain,\nand thus\n"]], ["block_4", ["to know the distance and direction between every pair of beads. Clearly substantial simplifications\nare required.\nThe first thorough treatment of H1 in polymers was developed by Kirkwood and Riseman in\n1948 [13]. The crucial step in the development is to \u201cpreaverage\u201d the HI, which means to replace\nthe instantaneous positions of the beads with their average positions, in particular as given by the\nGaussian distribution. This average can be evaluated for the separation of any pair of beads using\nthe Gaussian distribution (Equation 6.7.12) because we recognize r5] as the end-to-end distance of a\nchain from bead z' to bead j, that is, with li\u2014jl steps of length 19:\n"]], ["block_5", ["358\nDynamics of Dilute Polymer Solutions\n"]], ["block_6", [{"image_1": "370_1.png", "coords": [37, 158, 317, 217], "fig_type": "molecule"}]], ["block_7", [{"image_2": "370_2.png", "coords": [38, 272, 274, 347], "fig_type": "molecule"}]], ["block_8", ["Rh 0.271N1/2b 0.66Rg\n(9.7.4)\n"]], ["block_9", ["1\nN\nN\n1\nN\nN\nTrii;=*r;\u00a7\nlz\u2014\u2018l\n"]], ["block_10", ["1\n1\n3\n3/2\n3\nr3;-\n_\n=\n_4\n.2._____\ndry\n<33)\n3r..- Withdraw)\nexP< ZIP-31152\n"]], ["block_11", ["6\n1/2\n2 (\u201dIt _jlb2)\n(9.7.2)\n"]], ["block_12", ["12812/\n1\n3.69\n"]], ["block_13", ["377\nNI/Zb :NW;\n(973)\n"]]], "page_371": [["block_0", [{"image_0": "371_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["all the other beads in the same direction so that there is a concerted element to the chain motion.\nThis cooperation among segments contributes to a reduction in f, relative to Ng.\nAlthough the limits of freely draining and nondraining coils are easy to visualize, it may not be\nobvious why a random coil should tend to approximate the nondraining limit. A rather straight-\nforward argument shows why. The cartoon in Figure 9.15c depicts a layer of pure solvent \ufb02owing\nwith velocity v past an infinite field of beads; the beads have a number density of n/V. Although the\nflow may penetrate into the field of beads by some finite distance, if we are far enough from the\nsurface the \ufb02ow velocity must tend to zero. Without worrying about the exact functional form of\nthe decay, we will just assert that there is a characteristic penetration distance L, over which\nthe velocity vanishes. Now we will estimate the force F required to maintain the solvent \ufb02ow over\nan area of surface A by two separate routes; numerical prefactors will be omitted. In the first\nroute, we add up the friction of each bead (i.e., as if the beads were freely draining within the\nlayer of thickness L). The number of beads is given by (n/V) times the volume LA and thus the\nforce would be\n"]], ["block_2", ["Figure 9.15\nIllustration of (a) a freely draining coil, (b) a nondraining coil, and (c) the argument for the\nimportance of (b) given in the text.\n"]], ["block_3", ["(a)\n(b)\n"]], ["block_4", ["Hydrodynamic Interactions and Draining\n359\n"]], ["block_5", ["In the second route, we recognize that the force will be proportional to 775, A, and the velocity\ngradient (recall the discussion accompanying Figure 9.1). In this case the velocity gradient should\nbe proportional to v/L, that is, the velocity falls from v to 0 over the distance L. Therefore the force\ncan be written as\n"]], ["block_6", [">\n\u2019I\u2019\n\u2019 ------\nxxx\n..\n-\u201d,\"\u201d\nI,\u2019\nxxx.-\n._____~>\n--------------i\n_____\u2018,\u2019\n~~..____y\n----------------------------- p\n"]], ["block_7", ["--------------------------------- >\n(\u201d\ufb02y\n-_---~\"-V\n"]], ["block_8", ["\\\\u2018k\u2018.\n-._ _____\n\u201d/\n--------------------------------- >\n\u2018x\u2018\ufb01\n\u2018-__.__-\u201d\n"]], ["block_9", ["_________________________________ >\n\u2018H\u2018N\u2019n\ufb01\n\\\\~_\n\ufb02,\u201d\n-\"'\"'->\n"]], ["block_10", [{"image_1": "371_1.png", "coords": [40, 543, 118, 578], "fig_type": "molecule"}]], ["block_11", [{"image_2": "371_2.png", "coords": [42, 625, 112, 660], "fig_type": "molecule"}]], ["block_12", ["rI % (%)LA\u00a7V\n(9.7.5)\n"]], ["block_13", ["F, (am.\n(9.7.6)\n"]], ["block_14", ["________________ >\n---------- >\n-\u2018~\\\\\n\u2019r\u201d\ufb02--->\n"]], ["block_15", ["(C)\n"]], ["block_16", ["A *\nO\nO\nO\nO\nO\nO\nO\nO\nO\n"]], ["block_17", ["9.6_Q__o_,o\no\no\no\no\no\no\no\n0\n0\nO\n0\n9-49\n0\nO\nO\no\no\nO\nV\no \nO\no\no\no\no\no\no\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\nO\n"]], ["block_18", ["............................ y\n............................ *\n"]], ["block_19", [{"image_3": "371_3.png", "coords": [208, 48, 396, 137], "fig_type": "molecule"}]]], "page_372": [["block_0", [{"image_0": "372_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["SEC is one of several modes of liquid chromatography in which a mixture of solutes is separated\nby passing a solution through an appropriate column. As the solution (mobile phase) passes\nthrough the column, different solutes are retained to various degrees according to their interaction\nwith the column packing (stationary phase). Surface adsorption and ion exchange are examples\nof interactions that serve as the basis for other types of liquid chromatography. In SEC, the\ncolumns are \ufb01lled with porous particles and the separation occurs because molecules of different\nsizes penetrate the pores of the stationary phase to varying degrees. The method is akin to a\n\u201creverse sieving\u201d at the molecular level. The largest molecules are excluded from the pores to\nthe greatest extent and, hence, are the first to emerge (elute) from the column. Progressively\nsmaller molecules permeate the porous stationary phase to increasing extents and are eluted\nsequentially. The eluting liquid is monitored for the presence of solute by a suitable detector and\nan instrumental trace of the detector output (chromatogram) provides distinct peaks for well-\nresolved mixtures and broad peaks for a continuous distribution of molecular sizes. With suitable\ncalibration or multiple detection schemes, this information can be translated into a quantitative\ncharacterization of the sample in terms of molecular weight, molecular weight distribution,\nchemical composition, and even architecture. Owing to this wealth of potential information,\nthe relatively rapid sample throughput (typically ~30 min per solution), and the ease of automa\u2014\ntion, SEC is currently the single most important characterization tool in the polymer industry. In\nour discussion, we will \ufb01rst describe the basic separation process and identify associated\nstrengths and weaknesses. Then we will explore two general strategies for column calibration,\nthat is, how the measured quantity (the elution time or elution volume) can be related to the\nsolute molecular weight. We conclude with a description of various detection schemes that are\ncun\u2018ently employed, and in particular how they can be used to overcome some of the difficulties\nin calibration.\nTo avoid confusion, it is helpful to realize that what is essentially the same method is known by\nseveral different names\u2014and their acronyms\ufb02\u2014by workers in different fields. Some other terminO-\nlogies are noted below:\n"]], ["block_2", ["The key question is how does this penetration distance compare to the typical coil size? If L >> Rg,\nthen we expect freely draining behavior, but if L << Rg, nondraining will be a better description.\nNote that for a coil, the number of beads per unit volume n/V will be proportional to N/R3 or\nN14\", so from Equation 9.7.7 we can see that L depends on N raised to the power of (312\u2014 1)/2, or\n1/4 for a Gaussian coil. But, we know that Rg grows as N\u201d2, so that for big polymers Rg > L and\ntherefore the coil will be nondraining.\nIn real polymer solutions, the situation is more complicated and tends to fall between these\ntwo extremes, albeit closer to the nondraining limit. For example, as the solvent quality is\nimproved beyond a theta solvent, the chain expands and thus the average distance between\nbeads increases. Hydrodynamic interactions are therefore diminished. For chains in good solv-\nents, part of the variation in the Mark\ufb02Houwink exponent, a, can be attributed to variable degrees\nof draining. The effect also depends on molecular weight. For shorter chains the degree of\ndraining increases simply because there are not enough monomers to shield the inside of the coil\nfrom the solvent \ufb02ow. In summary, the main qualitative effect of H1 is to make the chain friction\nfactor close to that of a hard sphere with radius Rh, but a full mathematical treatment is extremely\ncomplicated.\n"]], ["block_3", ["9.8\nSize Exclusion Chromatography (SEC)\n"]], ["block_4", ["360\nDynamics of Dilute Polymer Solutions\n"]], ["block_5", ["We now equate these two forces, and solve for L:\n"]], ["block_6", ["(9.7.7)\n"]]], "page_373": [["block_0", [{"image_0": "373_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "373_1.png", "coords": [23, 401, 222, 633], "fig_type": "figure"}]], ["block_2", ["A cartoon of an SEC experiment is shown in Figure 9.16. The column is packed with porous\nparticles with some characteristic pore size, which should correspond roughly to the typical sizes\nof the polymers to be analyzed. The first rule of SEC is that the separation is based on size and not\non molecular weight; as we shall see the relevant size parameter is the hydrodynamic volume, Vh,\nwhich1s roughly proportional to R3 (recall Equation 9.3.8) From the radius of gyration data for\npolystyrenes in Figure 7.17, we can see that a typical range of pore sizes might be 10\u2014104 A. In\nhigh--resolution applications, it is often desirable to have a sequence of two to four columns with\ndifferent average porosities, so that a broader range of sizes can be resolved; the price to be paid is\nthat more time is required for each polymer to elute. The packing material can be made from a\nvariety of substances. but the chosen substance must be compatible with the solvent. polymer, and\ntemperature to be employed. For example, Styragel columns are made from styrene/divinylbenzene\ncopolymers, in the form of cross\u2014linked beads. Such columns are suitable for relatively nonpolar\npolymers that dissolve in good solvents for polystyrene such as toluene, THF, or chloroform.\n"]], ["block_3", ["2.\nGel filtration chromatography (GFC) is a name often used in the biochemical literature to\ndescribe this method of separation. Under this heading, the method is primarily applied to\naqueous solutions of solutes of biological origin.\n3.\nAll three of these names (SEC, GPC. and GFC) are also modified by the term high perform-\nance. or its acronym, to give HPSEC, HPGPC, and HPGFC. The additional feature implied by\nthis terminology is the increased speed of efficient separations due to rapid \ufb02ow through the\ncolumn under the in\ufb02uence of relatively high applied pressures. At the time of writing. this\nprefix is generally omitted because \u201clow performance\u201d instruments are little used.\n"]], ["block_4", ["Size Exclusion \n351\n"]], ["block_5", ["9.8.1\nBasic Separation Process\n"]], ["block_6", ["Figure 9.16\nSchematic illustration of an SEC column, packed with spherical porous particles, and indi-\nvidual molecules either inside or outside the pores.\n"]], ["block_7", ["1.\nGel permeation chromatography (GPC) is a very common name for this method of separation;\nit is a less desirable term than SEC in that it emphasizes the column packing material (the gel)\nrather than the supposed mechanism for the separation (size exclusion).\n"]], ["block_8", [{"image_2": "373_2.png", "coords": [75, 455, 210, 583], "fig_type": "figure"}]]], "page_374": [["block_0", [{"image_0": "374_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "374_1.png", "coords": [27, 26, 84, 291], "fig_type": "figure"}]], ["block_2", ["(Consider what would happen to a Styragel column if the mobile phase were a nonsolvent for\npolystyrene.) A great deal of clever engineering has gone into the development of packing materials\nof controlled pore size that can withstand large pressure drops and remain stable for weeks of constant\nuse. The solvent is pumped through the column at a slow but steady rate, typically on the order of 1\nmL/min. The pump itself plays an important role in the experiment, in maintaining a constant\npressure drop across the column; a good one is expensive. The solvent must also be good for the\npolymer to be analyzed so that the individual molecules are swollen and have no tendency to either\nprecipitate or adsorb on the column. A small volume of dilute polymer solution is injected through a\nspecial port upstream of the column. The detector (or series of detectors) responds to various\npr0perties of the eluting solution. Concentration detectors, such as those based on refractive index\nor uv\u2014vis absorption, have been the most commonly used, but these are now often supplemented by\nlight scattering and/or viscometric detectors. The resulting chromatogram is a plot of detector signal\nversus time; the time axis is typically converted to volume by multiplying by the known \ufb02ow rate.\nTo use SEC for molecular weight determination, we must relate the volume of solvent that\npasses through the column before a polymer of a particular M is eluted, to M. This quantity is\ncalled the retention volume VR. Figure 9.17 shows schematically the relationship between M and\nVR; it is an experimental fact that such calibration curves are approximately linear over about two\norders of magnitude in M when plotted as log M versus VR. In practice, the column is calibrated by\nconstructing such a curve with standards of known molecular weight, as we will discuss later.\nHowever, for the present purposes, we will assume that this curve is known. There are three\nregimes of response. For all M above a certain value, the curve is vertical. This means that all these\npolymers elute together; there is no separation. This happens for all polymers whose size is larger\nthan the largest pore; they are excluded from all the pores and thus travel entirely with the imposed\nflow through the interstices or voids between packing particles. Similarly, at low M there is a\n"]], ["block_3", ["362\nDynamics of Dilute Polymer Solutions\n"]], ["block_4", ["Figure 9.17\nCalibration curve for SEC as log M versus the retention volume VR, showing how the location\nof the detector signal can be used to evaluate M. Also shown are the void volume Vv and the internal volume\n"]], ["block_5", ["Detector\noutput\n"]], ["block_6", ["V, in relation to VR and KVi as a fraction of V,.\n"]], ["block_7", ["Log M\n"]], ["block_8", [{"image_2": "374_2.png", "coords": [43, 49, 240, 292], "fig_type": "figure"}]], ["block_9", ["/\nCalibration curve:\n"]], ["block_10", ["VF:\n"]], ["block_11", ["Idealized (solid)\nActual (broken)\n"]]], "page_375": [["block_0", [{"image_0": "375_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "375_1.png", "coords": [14, 388, 336, 632], "fig_type": "figure"}]], ["block_2", ["region with no separation, when all polymers that are small enough to enter all the pores will elute\nwith the solvent. The intermediate range of M is the useful regime: a peak at a particular VR can be\nrelated to a particular M, as shown in Figure 9.17. Note the curious fact that it is a common practice\nto plot the measured quantity, or dependent variable, VR along the horizontal axis, and the\n"]], ["block_3", ["Size Exclusion Chromatography (SEC)\n363\n"]], ["block_4", ["A broad chromatogram, shown in Figure 9.18, is subdivided into 20 slices, each 1 mL wide, and\nthese are indexed from 1': 1\u201420. The height h of the curve above horizontal base line is carefully\nmeasured for each slice. The molecular weight of the ith slice is assigned from independent\ncalibration via the retention volume. Columns 2\u20144 in Table 9.4 list hi, VR,,-, and M,- values,\nrespectively, for a particular chromatogram. Explain the significance use of the remaining\ncolumns in Table 9.4 for the determination of a molecular weight distribution from these data.\n"]], ["block_5", ["The basic premise of this method is that the magnitude of the detector output, as measured by for\n"]], ["block_6", ["independent variable, log M, along the vertical axis.\nPolydisperse polymers do not yield sharp peaks in the detector output, in contrast to the one\nillustrated in Figure 9.17. Instead, broad bands are produced that re\ufb02ect the polydispersity of\nsynthetic polymers. Assuming that suitable calibration data are available, we can \napproximate molecular weight distributions from this kind of experimental data. An \nof how this is done is provided in the following example.\n"]], ["block_7", ["a particular fraction, is pr0portiona1 to the weight fraction of that component in the sample. (This is\n"]], ["block_8", ["a reasonable premise for a detector, such as an R1 detector, that has a response proportional to\n"]], ["block_9", ["Solution\n"]], ["block_10", ["Example 9.6\n"]], ["block_11", ["Figure 9.18\nDiscrete version of the SEC chromatogram for the data in Example 9.6. The inset shows the\nweight fraction versus molecular weight.\n"]], ["block_12", ["0.2\nI\nI\nI Illllt\nI\nI\nI IIIIIl\nI\nI\nI IOIIII\n100\nI|I|'|\u2018_r'|'l'l'l'l'l\u2018l'\n"]], ["block_13", ["40-\n"]], ["block_14", ["20-\n"]], ["block_15", ["80-\n"]], ["block_16", ["60-\nlllll\nI\nI\nIIJJIII\n.Iu-I-I\u2018IIII\n\"\n1OB\n\u2014l O\n\u00ab.4\n"]], ["block_17", [{"image_2": "375_2.png", "coords": [43, 488, 305, 617], "fig_type": "figure"}]], ["block_18", ["_\n0.15\n"]], ["block_19", ["20\n22\n24\n26\n28\n30\n32\n34\n"]], ["block_20", ["w\n'22;\n5'1.\n\ufb01g.\n.14.?\nr\n.\nrail .2\"\u2014 la\u00bb\n"]], ["block_21", ["VR, \u201d1\u2018.\n"]], ["block_22", [{"image_3": "375_3.png", "coords": [169, 396, 322, 487], "fig_type": "figure"}]]], "page_376": [["block_0", [{"image_0": "376_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "376_1.png", "coords": [10, 105, 464, 353], "fig_type": "figure"}]], ["block_2", ["Source: From Yau, W.W., Kirkland, 1.1., and Bly, D.D., in Modern Size Exclusion Chromatography, Wiley, New York, 1979.\n"]], ["block_3", ["Table 9.4\nData for the Analysis of the Size Exclusion Chromatogram of a Polydisperse Polymer\nUsed in Example 9.6\n"]], ["block_4", ["the concentration of the solute. However, the proportionality constant, related to dn/dc in this\ninstance, must itself be independent of M. Furthermore, it assumes that all of the polymer elutes\nfrom the column, or, failing that, that any adsorption or other loss of material on the column is\nindependent of M.) In this sense, the chromatogram itself presents a kind of picture of the\nmolecular weight distribution. The following column entries provide additional quantifications\nof this distribution:\n"]], ["block_5", ["364\nDynamics of Dilute Polymer Solutions\n"]], ["block_6", ["21\n0.0\n20\n4.709\n545.0\n0.0\n0.0\n545.0\n1.0\n20\n0.0\n21\n3.302\n545.0\n0.0\n0.0\n545.0\n1.0\n19\n0.8\n22\n2.327\n545.0\n0.34\n1.86\n544.6\n0.999\n18\n3.5\n23\n1.640\n544.2\n2.13\n5.74\n542.5\n0.995\n17\n16.8\n24\n1.1555\n540.7\n14.54\n19.40\n532.3\n0.977\n16\n42.4\n25\n0.8142\n523.9\n52.08\n34.52\n502.7\n0.922\n15\n67.9\n26\n0.5738\n481.5\n118.2\n38.90\n447.6\n0.821\n14\n81.5\n27\n0.4003\n413.7\n203.6\n32.62\n373.0\n0.684\n13\n81.4\n28\n0.2821\n322.2\n288.6\n22.96\n291.5\n0.535\n12\n71.0\n29\n0.1988\n250.8\n357.1\n14.12\n215.3\n0.395\n11\n57.0\n30\n0.1401\n179.8\n406.8\n7.98\n151.3\n0.278\n10\n43.0\n31\n0.09872\n122.8\n435.6\n4.24\n101.0\n0.186\n9\n30.0\n32\n0.06887\n79.8\n435.6\n2.07\n64.8\n0.1 19\n8\n19.0\n33\n0.04853\n49.8\n391.5\n0.92\n40.3\n0.074\n7\n12.1\n34\n0.03420\n30.8\n356.7\n0.42\n24.7\n0.045\n6\n9.0\n35\n0.02410\n18.6\n373.4\n0.22\n14.1\n0.026\n5\n4.0\n36\n0.01698\n9.6\n235.6\n0.07\n7.6\n0.014\n4\n2.6\n37\n0.01197\n5.6\n217.2\n0.03\n4.3\n0.008\n3\n2.0\n38\n0.00843\n3.0\n237.1\n0.02\n2.0\n0.004\n2\n1.0\n39\n0.00588\n1.0\n170.0\n0.01\n0.5\n0.001\n"]], ["block_7", ["(1)\n(2)\n(3)\n(4)\n(5)\n(6)\n(7)\n(8)\n(9)\ni\nh,\nv,\u201d mL\nM, ><10_6 g/mol\n2.11,\n12,7114, x 106\n11,114, x 10\u20146\nA,-\n4,714,,t\n"]], ["block_8", [{"image_2": "376_2.png", "coords": [39, 511, 322, 560], "fig_type": "molecule"}]], ["block_9", ["1\n0.0\n40\n0.00414\n0.0\n0.0\n0.0\n0.0\n0.0\n"]], ["block_10", ["Column 5. 2h, is proportional to the cumulative weight of all polymers in all categories up to\nthe nth.\nColumn 6. The ratio hg/Mg is proportional to the weight of materials in the ith slice, 142,-, divided by\nM,-, that is, to the number of moles in that class 12,. Therefore Mn can be evaluated as follows:\n"]], ["block_11", ["Column 7. The product hiM, is proportional to ni, and Mw is evaluated as follows:\n"]], ["block_12", ["Column 8. A,- 2:1 [hi + 1/2(l7,-+1\u2014 h,-)]. Adding (1/2)(h,+1\u2014h,-) gives h, the height of the\nmidpoint of each slice, and since each slice is 1 unit wide, the summation gives the area under\nthe curve up to the nth class.\n"]], ["block_13", ["MW\n"]], ["block_14", ["Mn\n"]], ["block_15", ["_ ZiniMi __ Z; _\n2,11:\n545\n_\nDm- \nZ,(hi/M1)\n\" Z, (h./M,) = 4.29 x 10-3\n= 127,000 g mol\u20141\n"]]], "page_377": [["block_0", [{"image_0": "377_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "377_1.png", "coords": [30, 459, 310, 612], "fig_type": "figure"}]], ["block_2", ["Figure 9.19\nSchematic illustration of size exclusion in a cylindrical pore: (a) for spherical particles of\nradius R and (b) for a \ufb02exible chain, showing allowed (solid) and forbidden (dotted) conformations.\n"]], ["block_3", ["where K is called the distribution coefficient. K is a function of both the pore size and the\nmolecular size and indicates what fraction of the internal volume is accessible to the particular\nsolute. The relationships among VR, Vv, Vi, and KV, are also indicated in Figure 9.17. When K 0,\nthe solute is totally excluded from the pores; when K 1, it totally penetrates the pores.\nIt is instructive to consider a simple model for the significance of the constant K in Equation\n9.8.2. For simplicity, assume a spherical solute molecule of radius R and a cylindrical pore of\nradius a and length L. As seen in Figure 9.19a, an excluded volume effect prevents the center of the\n"]], ["block_4", ["The entire interstitial volume moves through the column at the imposed \ufb02ow rate and must pass\nthrough the column before any polymer emerges. Then the first polymer that does appear is the one\nwith the highest molecular weight. This solute has spent all its time in the voids\u2014not the pores\u2014of\nthe packing and passes through the column with the velocity of the solvent. Progressively smaller\nmolecules have access to successively larger fractions of the internal volume. Therefore, as V,\nemerges, consecutive fractions of the polymer come with it. Thus we can write the retention\nvolume for a particular molecular weight fraction as\n"]], ["block_5", ["(a)\n"]], ["block_6", ["A more thorough examination of the correlation between VR and M can be found in [14]. We shall\nonly outline the problem, with particular emphasis on those aspects that overlap other topics in this\nbook. To consider the origin of VR(M), begin by picturing a narrow band of polymer solution being\nintroduced at the top of a solvent\u2014filled column. The volume of this solvent can be subdivided into\ntwo categories: the stagnant solvent in the pores (subscript i for internal) and the interstitial liquid\nin the voids (subscript v) between the packing particles:\n"]], ["block_7", ["__._\u2014\n"]], ["block_8", ["9.8.2\nSeparation Mechanism\n"]], ["block_9", ["A plot of the last entry versus M gives the integrated form of the distribution function. The more\nfamiliar distribution function in terms of weight fraction versus M is given as the inset to Figure 9.18.\n"]], ["block_10", ["Size Exclusion Chromatography (SEC)\n365\n"]], ["block_11", ["Column 9. Aj/Atot gives that fraction of the area under the entire curve that has accumulated up to\nthe nth class. Since the curve is a weight distribution, this is equal to the weight fraction of\nmaterial in the sample having M 2\n2\n[hm\u2014TEL: (1\u20145)\n(9.8.3)\n(1\n"]]], "page_379": [["block_0", [{"image_0": "379_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "379_1.png", "coords": [31, 27, 209, 300], "fig_type": "figure"}]], ["block_2", ["other details of the distribution for each sample along the lines illustrated in Example 9.6 above.\nThis approach is in common practice and it is appealingly simple. The calibration should be\nrepeated at regular intervals because column performance will drift with time, but as only\nminuscule amounts of the standards are required for each run, this is not particularly expensive.\nWhat are the limitations of this approach? There are several to consider, any of which may or may\nnot be important for a particular application.\n"]], ["block_3", ["the packing material can drastically increase VR. Furthermore, this contribution causes VR to\nincrease with M, in opposition to size exclusion, because the net attraction will increase with the\nnumber of monomers. In short, a quantitative theory would be very complicated. The second\nreason is more practical. Calibration techniques, to be described in the next section, are perfectly\nadequate for many applications, and modern detectors, which can even circumvent the need for\ncolumn calibration, satisfy most requirements. Consequently, the need for a full theory is reduced.\n"]], ["block_4", ["The simplest calibration strategy was alluded to above: take a series of polymers of known M,\nreferred to as calibration standards, run them through the column, and generate an empirical plot of\nlog M versus VR. Some kind of polynomial function (it need not be linear) can be used to obtain a\nsmooth function, which can be stored in a computer. That function can be used to assign a molecular\nweight to each value of VR and the computer can easily calculate the molecular weight averages and\n-\n"]], ["block_5", ["Figure 9.20\nComparison of theory with experiment for Rg/(a) versus K. The solid line is drawn according\nto the theory for \ufb02exible chains in a cylindrical pore. Experimental points show some data, with pore\ndimensions determined by mercury penetration (circles, (a) =21 nm) and gas adsorption (squares, (a) 41\nnm). (From Yau, W.W. and Malone, C.P., Polymer Preprints (Am. Chem. 506., Div. Polym. Chem), 12, 797,\n1971. With permission.)\n"]], ["block_6", ["9.8.3\nTwo Calibration Strategies\n"]], ["block_7", ["9.8.3.1\nLimitations of Calibration by Standards\n"]], ["block_8", ["Size Exclusion Chromatography (SEC)\n367\n"]], ["block_9", ["1.\nSeparation is based on the hydrodynamic volume, V1,, and not on M directly. Consequently,\naccuracy in M requires the use of standards of the same polymer as the analyte. Although\n"]], ["block_10", ["0.04\n|\nI\nI\nI\nI\nI\nI\nI\nI\nI\n|\n0\n0.2\n0.4\n0.6\n0.8\n1 .0\nK\n"]], ["block_11", ["0.1\n0.08\n"]], ["block_12", ["0.06\nIIIII\n"]], ["block_13", ["2.0\n"]], ["block_14", ["1.0\n0.8\n"]], ["block_15", ["0.6\nI_Ill|l\n"]], ["block_16", ["0.4\n"]], ["block_17", ["0.2 l-\n"]], ["block_18", ["_|__|'\n"]]], "page_380": [["block_0", [{"image_0": "380_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The second mode of calibration in routine use is known by the optimistic name of universal\ncalibration. The hypothesis is that in SEC, VR depends solely on the hydrodynamic volume, Vh,\nwhich itself will be proportional to R2. Now recall the discussion of intrinsic viscosity leading up to\n"]], ["block_2", ["9.8.3.2\nUniversal Calibration\n"]], ["block_3", ["controlled polymerization, especially anionic polymerization as discussed in Chapter 4, is used\nto produce standards of a variety of polymers (polystyrene, poly(methyl methacrylate), poly-\nbutadiene, polyisoprene, poly(ethylene oxide). . . ), there are, of course, a much larger number\nof polymer structures for which standards are simply not available. One expedient often\nencountered in the literature is to quote \u201can apparent M, based on polystyrene standards,\u201d\nwhich gives some very approximate idea of the real M.\n2.\nThe absolute M of a given stand may not be known to an accuracy better than 5%\u2014\u201410%,\nwhich ultimately limits the accuracy achievable with this mode of SEC.\n3.\nIf one injected an absolutely monodisperse standard, an ideal instrument would give a very\nsharp spike at the particular value of VR. In reality, for such a sample the instrument would\nshow a narrow peak with a finite width. This width characterizes the instrument response\nfunction, which quantifies the deviation of the instrument from ideality. In typical SEC, a\nmonodisperse sample would show a peak width corresponding to a polydispersity of about\n1.01\u20141.05. The reason for this is the phenomenon of band broadening. One significant and\neasily visualized contribution to band broadening is termed axial di\ufb01cusion. As molecules of\nidentical M pass through the column, even if they were injected at precisely the same time,\nthey will all diffuse randomly. Some will show a net displacement relative to the average that\nplaces them further down the column and others will lag behind. Indeed, based on the\narguments in Section 9.5, we might anticipate a Gaussian distribution of concentration versus\nVR. This has important consequences. For example, it means that no matter how narrow a\n\u201cslice\u201d of the chromatogram we select, the material eluting at that particular VR is never a\nsingle M; each slice i has its own MW and MN. Furthermore, for a given column, a particular\nVR does not always correspond to one particular MW},- or Mm; the average M of the material at a\nparticular VR will depend on the sample.\n4.\nIn most cases calibration standards have very narrow molecular weight distributions (Mw/\nMn 3 1.1). However, the true polydispersity of such materials is often not known. For example,\nif MW is obtained by light scattering and MH by osmotic pressure, the combined uncertainties\nof, say, 5% in each number means that one cannot distinguish among polydispersities of 1.03,\n1.01 , or 1.001. Recall from Chapter 4 that standards prepared by living anionic polymerization\nshould ideally follow the Poison distribution; a polystyrene with Mn :100,000 would have a\ntheoretical ideal polydispersity of 1.001. It is therefore quite probable that most calibration\nstandards are narrower than can be determined by SEC. Recent progress in mass spectrometric\nmethods (see Chapter 1) offers the possibility that such materials may be characterized more\nprecisely.\n5.\nMany polymers have nonlinear or branched architectures (recall the examples in Chapter 1).\nTwo polymers of identical M but different amounts of branching will have different Rg and\ntherefore different VR. For example, in the free radical polymerization of ethylene, the product\ncalled \u201clow density\u201d polyethylene has a substantial degree of long\u2014chain branching. It is a\nnotoriously difficult problem to characterize the resulting molecular structures. In particular, it\nis impossible to learn anything about the degree of branching from SEC when only this simple\nmethod of calibration is employed.\n6.\nIt is often the case that polymer samples are heterogeneous not just in molecular weight or\narchitecture, but also in composition (e.g., copolymers), microstructure (e.g., polydienes), and\ntacticity. All of these factors may contribute in some way to VR and each slice of the\nchromatogram will also be heterogeneous in these variables.\n"]], ["block_4", ["368\nDynamics of Dilute Polymer Solutions\n"]]], "page_381": [["block_0", [{"image_0": "381_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This is the most commonly employed. In Section 8.7 we discussed the central role that the\nrefractive index increment, arr/ac, plays in light scattering. For a dilute polymer solution (which\nis almost always the case in SEC), the refractive index of the eluent may be written\n"]], ["block_2", ["To conclude this discussion of SEC we offer a brief discussion on the four classes of detectors in\ncommon use: the refractive index (RI), absorption (uv\u2014vis), light scattering (LS), and viscometer\n(V) detectors. It is increasingly common practice to utilize two or more of these in series, for\nreasons that should become apparent. There are different implementations of these various\ndetectors by different companies and we shall try to present the general principles without\nreference to particular instrument con\ufb01gurations.\n"]], ["block_3", ["where c is the concentration in g/mL of the eluting polymer and n3 is the refractive index of the\nsolvent. The refractometer is typically set up such that a transmitted light beam is de\ufb02ected by an\namount proportional to n\u2014ns and thus to c. The response of the detector (e.g., in volts) can then be\nassumed proportional to c and the proportionality factor determined by injection of known\nquantities of material (assuming no adsorption on the column and that 371/60 is known). However,\nin routine cases the proportionality factor is not necessary; the value of any M depends only on VR,\nnot on the height of the peak (i.e., h,~), and in the calculation of polydispersities the proportionality\nconstant would cancel out (see Example 9.6).\n"]], ["block_4", ["In short, if we know the appropriate Mark\u2014Houwink parameters (and a great many have been\ntabulated, as noted in Section 9.3) then we can extract the absolute molecular weight of one polymer\nbased on column calibration with another. This approach is therefore often used to overcome the first\nlimitation listed above. The success of the underlying assumption behind universal calibration is\nnicely illustrated in Figure 9.21. In the first panel, plots of log M versus VR are shown for a variety of\ndifferent polymers, including some branched structures; clearly the different species are all over the\nmap. In the second panel, the same data are shown, but with the vertical axis being log([n]M). In this\ncase, there is a very satisfying collapse of the data onto one universal calibration curve.\n"]], ["block_5", ["Equation 9.3.8, where the essence of the argument is that, no matter the detailed molecular\nstructure or \n"]], ["block_6", ["The hydrodynamic volume, therefore, should be proportional to the product [n]M. In universal\ncalibration, we assume that the proportionality factor between the hydrodynamic volume and [n]M\nis independent of structure. Let us further suppose that [n] for our sample follows the Mark\u2014\nHouwink relation (Equation 9.3.10), with known values of k and a. We compare our sample with\nthe standard, say polystyrene, that elutes at the same time:\n"]], ["block_7", ["9.8.4. 1\nRI Detector\n"]], ["block_8", ["9.8.4\nSize Exclusion Chromatography Detectors\n"]], ["block_9", ["where the subscripts \u201cr\u201d and \u201c5\u201d refer to the reference (calibrant) and sample polymers, respect*\nively. Equation 9.8.5 can be rearranged to solve for M of the unknown at any particular VR:\n"]], ["block_10", ["Size Exclusion Chromatography (SEC)\n369\n"]], ["block_11", [{"image_1": "381_1.png", "coords": [38, 535, 155, 572], "fig_type": "molecule"}]], ["block_12", ["MS (+)\n(9.8.6)\n"]], ["block_13", ["n(c) ns \u2014|\u2014 (@) C \u2014l\u2014 \n(9.8.7)\n86\n"]], ["block_14", ["[17] \n(9.8.4)\n"]], ["block_15", ["(MIMI: krM}+\u201c\u2018 ([nlM),~\u2014\u2014 ksM\u00a7+as\n(9.8.5)\n"]], ["block_16", [{"image_2": "381_2.png", "coords": [47, 235, 153, 279], "fig_type": "molecule"}]], ["block_17", [{"image_3": "381_3.png", "coords": [49, 83, 108, 119], "fig_type": "molecule"}]], ["block_18", ["(ke+ar)\nl/l-l-as\n"]], ["block_19", ["3\n"]]], "page_382": [["block_0", [{"image_0": "382_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "382_1.png", "coords": [21, 55, 280, 258], "fig_type": "figure"}]], ["block_2", ["This approach takes advantage of Beer\u2019s law, whereby the transmittance of the solution (ratio of\ntransmitted light intensity to incident light intensity, U10) is related to the concentration of the\nabsorbing species by\nI\nsbc :A log}?\u2014\n(9.8.8)\n"]], ["block_3", ["Figure 9.21\nThe success of universal calibration. The data in (a) as molecular weight versus retention\nvolume collapse to a single curve when plotted in (b) as the product of molecular weight and intrinsic\nviscosity. (Data from Grubisic, P., Rempp, H., and Benoit, J., J. Polym. Sci, BS, 753 1967.)\n"]], ["block_4", ["9.8.4.2\nUV\u2014Vr's Detector\n"]], ["block_5", ["370\nDynamics of Dilute Polymer Solutions\n"]], ["block_6", [".0\n\u2018\nQ\nn PS\u2014PMMA grafts\nE 108 t\nE]\nPPS\nE\nE\n'21:,\n - SMMA\n"]], ["block_7", ["E\na\n[J P8 stars\n-\nz]\na\nEl\nPS\u2014PMMA grafts\n75\nI\nn\nE\nin PPS\n\u00e9\n\u201da;\nu\n - SMMA\n03 106\nE\u2014\nBE]\nZ]\nIS!\n'\n:\n\u00a7\n:\nEma\u201d\nE\n:\ni\nan\ni\nE\n\u201cb\n105 :\u2014\nE\na\n\u2014:\n"]], ["block_8", ["E\u201c\n-\nH\n3\n\u00a7\n107\n:\u201c\nHiding\n\u20185\n"]], ["block_9", [{"image_2": "382_2.png", "coords": [40, 278, 274, 509], "fig_type": "figure"}]], ["block_10", ["1010\n\u2014\nl\u2014\n1\nI\nl\n|\n|\nI\nl'\nl\n"]], ["block_11", ["108:\"'|\"'|\"'\u2014l\n:\nI:I\nPS\nI\nm PMMA\n'\n1:1 PVC\n\u2014\nPB\n107 :L\nE\nEl PS combs\n"]], ["block_12", ["104...I..Jlr..l...l..lID...\n18\n20\n22\n24\n26\n28\n30\nVR,mL\n"]], ["block_13", ["105...L_._..l...l...l...ll:l...\n18\n20\n22\n24\n26\n28\n30\n"]], ["block_14", ["106\n:\u2014\nEl\n-:\n"]], ["block_15", ["109\n:\u2014\nEl\n2] PB\n"]], ["block_16", ["E\nD\n:\n_\n-\n"]], ["block_17", ["_\n_\n"]], ["block_18", ["_\n\u2018\nQ\n"]], ["block_19", ["\u2014\nEg\na P8 stars\n"]], ["block_20", ["E\n1:! PS\n"]], ["block_21", ["E\nEl PMMA\nE\na\n[SI PVC\n"]], ["block_22", ["_\n\ufb02\n"]], ["block_23", ["E\na\na P8 combs\n"]], ["block_24", ["t\n"]], ["block_25", ["VFi! mL\n"]]], "page_383": [["block_0", [{"image_0": "383_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "383_1.png", "coords": [33, 255, 310, 373], "fig_type": "figure"}]], ["block_2", ["where A is the absorbance, b is the path length through the capillary, and 8 is the absorptivity of the\nsolute at the wavelength of interest. Thus, as with the RI detector, the signal is arranged to be\nproportional to c. The uv\u2014vis detector is less general than the RI detector because many polymers\ndo not absorb at sufficiently long wavelengths to avoid solvent absorption. On the other hand, when\ndealing with mixtures, for example copolymers, it may be possible to choose a wavelength in the uv\u2014\nvis detector that favors only one component, whereas the RI detector is likely to respond to both. In this\ncase both detectors in series can be used to establish the average composition ofthe sample at each VR.\n"]], ["block_3", ["Size Exclusion Chromatography (SEC)\n371\n"]], ["block_4", ["The crucial advantage of the LS detector is the possibility of obtaining an absolute MW, of each slice\nof the chromatogram without any column calibration. This makes use of the theoretical machinery\ndeveloped in Chapter 8. If M is sufficiently low that the form factor, P(q), is effectively 1, then a\nsingle angle detector is adequate. On the other hand, for larger polymers multiple angle detectors are\ndesirable. Some commercial models have more than photodiodes surrounding the sample\nvolume, and both RgJ- and MW can be determined for each slice. An illustration is provided in\nFigure 9.22 and discussed in the following example.\n"]], ["block_5", ["9.8.4.3\nLight Scattering Detector\n"]], ["block_6", ["(a)\n"]], ["block_7", ["Figure 9.22\nMultiangle light scattering detector signal from SEC of polystyrene \nin Example 9.7. (a) Intensity versus time traces for the individual detectors \n"]], ["block_8", ["sin3(6/2)\n(b)\n"]], ["block_9", ["sin2(6/2) for a particular slice of the chromatogram near the peak in intensity.\n"]], ["block_10", ["\"(3 3.5 \nf%\nk\n"]], ["block_11", ["0?\nt\n.\n_\n"]], ["block_12", [{"image_2": "383_2.png", "coords": [41, 378, 301, 628], "fig_type": "figure"}]], ["block_13", ["4.5 x 1c \nMW: \n_\n"]], ["block_14", ["2.5 x 10\u20146 L\n_\u2014\n"]], ["block_15", ["5X10\u20146\u2014\nl\u2014I\n1\nl\nI\nr\u2014l\u2014l\u2014l\u2014l\u2014l\u2014l\u2014l\u2014I\n1\nI\nr_'_\nI\n-\n"]], ["block_16", ["2 X 10\u20146\n"]], ["block_17", ["E\nl\n4x10-5L\n99:253 \n1|\n"]], ["block_18", ["_\n.\n.\u2014 J\u2019 \u2018 .0\nj\n3 X 106 \n.\n#0....\u201d \"'\"\no\n_\n"]], ["block_19", ["0\n0'2\n0-4\n0.6\n0.8\n1\n"]], ["block_20", ["L\nI\n"]], ["block_21", ["t\n--- (*3;\n_\n"]], ["block_22", ["_\nf r .\n"]], ["block_23", ["I.\n"]], ["block_24", ["L\u2014 i \nI\n"]], ["block_25", ["n\nI\ni\nl_l\n|_L\n|_|_\nI\nI\nI\ng\n.\n|\nI\n.\nI\n"]], ["block_26", ["2.!\n11 09\n3\n"]], ["block_27", ["'\nII\n0')\n01\nI\u201d?\n"]], ["block_28", ["\ufb02li\n1\u2018!\n"]], ["block_29", ["I.\n"]], ["block_30", ["H:\n"]], ["block_31", ["_\n"]], ["block_32", ["IIIIE\u2018H\nailing\n'umlg-\"Lin-\u2014tart...\n"]], ["block_33", ["'\n"]], ["block_34", ["\u201cll-'\u2014\n[vi-l...\u2014\n"]], ["block_35", ["\u2018\n"]], ["block_36", [":@3166\n"]], ["block_37", ["141312\n"]]], "page_384": [["block_0", [{"image_0": "384_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "384_1.png", "coords": [31, 124, 186, 159], "fig_type": "molecule"}]], ["block_2", ["In this chapter we have examined some dynamic properties of polymers in dilute solution, with a\nparticular emphasis on the viscosity. The underlying concept is that of the molecular friction\n"]], ["block_3", ["The preceding discussion gives some insight into the rather sephisticated detection schemes\nnow in use. It is perhaps appr0priate to close the discussion of SEC with the major limitation of the\ntechnique, namely that no matter how elaborate or sensitive the detector, the separation itself has\nrather low resolution. Furthermore, for samples containing distributions of composition, micro-\nstructure, and architecture, in addition to molecular weight, it is necessary to employ other\nseparation schemes in tandem that can discriminate among these various characteristics. In normal\nSEC, all of these various distributions are likely to be present to some extent in each slice, and the\ninformation the detectors can provide will be limited.\n"]], ["block_4", ["where again we assume the concentration is small enough that extrapolation to infinite dilution is\nnot necessary. Note that in Equation 9.8.9 an accurate measurement of concentration and the\ninterdetector delay time is also essential.\n"]], ["block_5", ["This detector takes advantage of the direct relation between the hydrodynamic volume, which\ndetermines VR and [n]: Vh~M[n]. Consequently by estimating [n],- at each slice, M can be\ncalculated. The detector itself is a kind of \u201cWheatstone bridge,\u201d or null detector, for viscosity.\nThe eluting solution passes across one face of a sensitive pressure transducer. Pure solvent is\ncirculated against the other face. If both liquids are \ufb02owing at the same rate, any difference in\nviscosity between them will be transformed into a differential shear force and thus a pressure on\nthe transducer. This could be measured, but a more effective approach is to increase the pure\nsolvent \ufb02ow rate to null out the pressure dr0p. The amount the \ufb02ow rate needs to be increased is\ndirectly proportional to n 775. The intrinsic viscosity is then estimated by\n"]], ["block_6", ["where we have dropped the term containing the second virial coefficient B. (This assumption is\nusually safe because the concentrations coming off the column are small, but it can always be\nchecked by reinjecting a different concentration of sample and seeing if the answer changes.) The\nvalue of q is determined by the scattering angle for each detector. The constant K contains many\nfactors (see Equation 8.4.22), including (3n/3c)2, so it must be known accurately. This can be\naccomplished by calibrating the detector with a modest M sample of known MW and 311/36. The\nintensity for each slice should be plotted against q2 (or sin2(6/2)) to obtain Rg, as illustrated in\nFigure 9.22b using the \u201cZimm format\u201d (Equation 8.5.18). Finally, to obtain M for each slice\nrequires dividing by C5, obtained from either the RI or uv\u2014vis detector. In this case, it is necessary\nthat the concentration detector be accurately calibrated as well. It should also be noted that because\nthe two detectors are arrayed in series, there is a time delay between the arrival of slice 1' at each\ndetector. This must also be determined accurately.\n"]], ["block_7", ["9.9\nChapter Summary\n"]], ["block_8", ["9.8.4.4\nViscometer\n"]], ["block_9", ["Explain how to extract MW and Rg from each slice of a chromatogram.\n"]], ["block_10", ["372\nDynamics of Dilute Polymer Solutions\n"]], ["block_11", ["Recalling Equation 8.5.19, the scattered intensity in slice 1', 15,5, can be written as\n"]], ["block_12", ["Example 9.7\n"]], ["block_13", ["Solution\n"]], ["block_14", ["I\n-\n1\n"]], ["block_15", ["77 7?\n[77L \u20145 \n(989)\n7730i\n"]], ["block_16", ["Ii:-\n= KCn,g(I EQZREJ- \n\u00b0 )\n"]]], "page_385": [["block_0", [{"image_0": "385_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["factor, f, which depends on both polymer shape (rod, coil, sphere...) and molecular weight. The\nmain points are the following:\n"]], ["block_2", ["Problems\n"]], ["block_3", ["Problems\n373\n"]], ["block_4", ["1.\nA \ufb02uid of viscosity 1; is confined within the gap between two concentric cylinders as shown in\nFigure 9.9b. Consider a cylindrical shell of radius r, length L, and thickness dr located within\nthe gap.\n"]], ["block_5", ["1.\nThe friction factor is proportional to the product of the solvent viscosity and a hydrodynamic\nradius, Rh, where Rh varies with M just as Rg does. This corresponds to an extension of Stokes\u2019\nlaw for rigid spheres to polymers of any shape. The friction factor is experimentally accessible\nthrough either the intrinsic viscosity or the tracer diffusion coefficient.\nThe hydrodynamic volume, Vh, can also be defined, and is proportional to R1: and R; The\nEinstein equation for the viscosity of a suspension of hard spheres can be extended to polymers\nof any shape, with the result that the intrinsic viscosity is proportional to the ratio of Vh/M.\nThe relation between intrinsic viscosity and molecular weight can be expressed by the Mark\u2014\nHouwink equation [1)] =kM\", where the parameters k and a have been tabulated for many\npolymer and solvent combinations. As [7)] is relatively easy to measure, this offers a simple\nroute to molecular weight characterization.\nTracer and mutual diffusion coefficients were defined and distinguished. Diffusion coeffi-\ncients offer another route to molecular characterization and play a key role in many applica-\ntions of polymers.\nThe technique of SEC is the most commonly applied technique for polymer characterization,\n"]], ["block_6", ["1\n\u2014M-/M\ni \ne\nl\nn\nf\nMn\n"]], ["block_7", ["as it can determine both average molecular weights and the molecular weight distribution. The\nseparation is based on V1, and is of relatively low resolution. Application of various detection\nschemes can obviate the need for column calibration.\nThe reason that Stokes\u2019 law and Einstein\u2019s viscosity equation can be applied to polymers of\nany shape is rather subtle. It is due to the phenomenon of hydrodynamic interactions, whereby\nthe motion of any monomer in the polymer is transmitted through the solvent to all other\nmonomers. The net effect is that monomers tend to move collectively, like a hard sphere with\nradius close to Rg, rather than as N independent objects.\n"]], ["block_8", ["A slightly different but useful way of defining the viscosity average molecular weight is the\nfollowing:\nZ\ufb01MiM?\nM:\u2014\u2014\u2014\u2014\u2019\nv\nZfiMi\n"]], ["block_9", ["a. What is the torque acting on the shell if torque is the product of force and the distance from\nthe axis and F/A nr dw/dr?\nb. Under stationary-state conditions, the torques at r and at r + dr must be equal, otherwise the\nshell would accelerate. This means that the torque must be independent of r. Show that this\nimplies the following variation of a) with r: w \u2014B/2r2 + C, where B and C are constants.\nc. Evaluate the constant B by noting that w :wex, the experimental velocity, at r=R and\nw 0 at r =fR.\nd. Combine the results of a, b, and c to obtain Equation 9.4.14.\n"]], ["block_10", ["where \ufb01M,- is the weighting factor used to average 114?. A satisfactory way of treating many\npolymer distributions is to de\ufb01ne\n"]], ["block_11", [{"image_1": "385_1.png", "coords": [57, 567, 117, 609], "fig_type": "molecule"}]]], "page_386": [["block_0", [{"image_0": "386_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["tDJ. Plazek, w. Dannhauser, and JD. Ferry, J. Colloid Sci, 16, 1010 (1961).\n1TC: Fox Jr. and PJ. Flory, J. Am. Chem. 506., 73, 1915 (1951).\n"]], ["block_2", ["For an (it-helix, the length per residue is about 1.5 131. Use this figure with the molecular weight\nto estimate the length 2a of the particle. Use the estimated 61/!) ratios to calculate the diameter\n2b of the helix, which should be approximately constant if this interpretation is correct.\nComment on the results.\n6.\nFox and Floryi used experimental molecular weights, intrinsic viscosities, and ms end-to\u2014end\ndistances from light scattering to evaluate the constant (130 in Equation 9.3.18. For polystyrene\nin the solvents and at the temperatures noted, the following results were assembled (M in kg\nmol\u2018l, [n] in dL/g):\n"]], ["block_3", ["7} experiences three kinds of forces: gravitational, buoyant, and frictional. The \ufb01rst is\ndetermined by the mass of the ball (and the acceleration due to gravity, g), the second by\nthe mass of the displaced \ufb02uid and g, and the last is given by Equation 9.2.3 and Equation\n9.2.4. During most of the fall (excluding the very beginning and the very end of the\npath), these forces balance. Use this condition to derive an equation showing how this\nstationary-state velocity of fall is related to R, p1, p2, and n. This is the basis for the so-called\nfalling-ball viscometer.\n4.\nPlazek, Dannhauser, and FerryJr measured the viscosities of a poly(dimethylsiloxane) sample\nof MW =4.1><105 over a range of temperatures using the falling-ball method. Stainless steel\n(p2 7.81 g/cm3) balls of two different diameters, 0.1590 and 0.0966 cm, were used at 25\u00b0C,\nwhere p1 0.974 g/cm3 and n: 8.64x104 P. Use the result derived in the last problem to\ncalculate the ratio of the stationary\u2014state settling velocities for the two different balls. How\nlong would it take the smaller ball to fall a distance of 15 cm under these conditions?\n5.\nThe intrinsic viscosity of poly(y-benzyl-L-glutamate) (M02210) shows such a strong mo\u2014\nlecular weight dependence in dimethyl formamide that the polymer was suspected to exist as a\nhelix, which approximates a prolate ellipsoid of revolution in its hydrodynamic behavior.\n"]], ["block_4", ["374\nDynamics of Dilute Polymer Solutions\n"]], ["block_5", ["Combine the last two expressions and integrate to express Mv in terms of Mn and a. The\nintegrals are standard forms and are listed in integral tables as gamma functions.\n3.\nA sphere of density p2 and radius R falling through a medium of density p1 and viscosity\n"]], ["block_6", [{"image_1": "386_1.png", "coords": [44, 52, 139, 135], "fig_type": "molecule"}]], ["block_7", ["Using 1.32 g/cm3 as the density of the polymer, estimate the axial ratio for these molecules,\nusing Simha\u2019s equation (for large p E a/b):\nh _\nP2\n102\np\na\n\u2014 15[1n(2p)\u20143/2]+5[1n(2p)\u20141/2]+E g 175 (5\u20140)\n"]], ["block_8", ["Mx10\u20143(g/moi)\n21.4\n66.5\n130\n208\n347\n[n](dL/g)\n0.107\n0.451\n1.32\n3.27\n7.20\n"]], ["block_9", [",lfnHa dMi\nM(1 \u2014\u2014\u2014\u2014\u2014\u2014\u2014\u00b0\n,lfiMidMi\n0\n"]], ["block_10", ["Then\n"]], ["block_11", ["'\n"]]], "page_387": [["block_0", [{"image_0": "387_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Evaluate (Do for each set of data and compare the average with the value given in the text. Does\nthe fact that these data are not from theta solvents matter? Why, or why not?\n7.\nPrecise determination of the intrinsic viscosity, [7]], and the Huggins coefficient, (CH, is not\n"]], ["block_2", ["Problems\n375\n"]], ["block_3", [{"image_1": "387_1.png", "coords": [49, 40, 278, 268], "fig_type": "figure"}]], ["block_4", ["as straightforward as one might expect, even when the instrument provides accurate and\nprecise measurements of 11,61 (211/115). The primary issue becomes how many terms to\ninclude in the concentration expansion of 11(0), and what range of c is appropriate? The\nusual interval is\n1.1 <71re1<2\u00a7 smaller values have too much uncertainty, and larger\nvalues require too many terms. The usual strategy is to perform two linear extrapolations,\nand compare the results; if they do not agree, more data or more terms are required. The\nfirst extrapolation is due to Huggins; plot nsp/c versus (3, where 715p 11,31 1, and fit to a\nstraight line:\n"]], ["block_5", ["b. One test of the validity of determination of [n] and kH is to compare the results from these\ntwo plots. The data below are for polystyrene (M :20,000) in a theta solvent. Prepare the\ntwo plots, do the linear regression, and answer these questions:\ni. How well do the two fits satisfy the criteria you derived above?\nii. What are the implied values of [n] and (CH?\niii. What would you estimate the uncertainties to be?\n"]], ["block_6", ["\ufb01zai+\ufb01lc...\n"]], ["block_7", ["where a\u201d and B\u201d are the \ufb01t parameters.\n"]], ["block_8", ["In The] 2 an +BHC'\n. .\n"]], ["block_9", ["where a\u2019 and B\u2019 are the fit parameters. The second extrapolation is due to Kraemer: plot (1n\nmeg/c versus 6, and fit to a straight line:\n"]], ["block_10", ["Methyl ethyl ketone\n22\n1760\n1.65\n1070\n22\n1 620\n1 .61\n1 01 5\n67\n1620\n1.50\n980\n22\n1320\n1.40\n900\n25\n980\n1.21\n840\n22\n940\n1.17\n750\n22\n520\n0.77\n545\n25\n318\n0.60\n475\n22\n230\n0.53\n400\nDichloroethane\n22\n1780\n2.60\n1410\n22\n1620\n2.78\n1335\n67\n1620\n2.83\n1295\n22\n562\n1.42\n760\n22\n520\n1.38\n680\nToluene\n22\n1620\n3.45\n1290\n67\n1620\n3.42\n1280\n"]], ["block_11", ["a. Derive the relationships between (i) a\u2019 and or\u201d, (ii) 8\u2019 and B\u201d, and (iii) express 03\u2019 and B\u2019 in\nterms of [n] and k\u201c. It may help to recall that ln(l +x) :3: (1/2)x2 + (1/3)x3- \n"]], ["block_12", ["Solvent\n7*\nM\n[111\nkm, (A)\n"]], ["block_13", ["C\n"]], ["block_14", [{"image_2": "387_2.png", "coords": [96, 68, 245, 248], "fig_type": "table"}]]], "page_388": [["block_0", [{"image_0": "388_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["376\n"]], ["block_2", ["10.\n"]], ["block_3", ["11.\n"]], ["block_4", [{"image_1": "388_1.png", "coords": [41, 299, 273, 420], "fig_type": "figure"}]], ["block_5", [{"image_2": "388_2.png", "coords": [44, 50, 185, 144], "fig_type": "figure"}]], ["block_6", ["6,970,000\n1 1.75\n277,000\n1.07\n4,240,000\n8.15\n63,000\n0.358\n2,530,000\n5.54\n63,100\n0.356\n838,000\n2.43\n43,200\n0.268\n784,000\n2.32\n16,050\n0.136\n676,000\n2.07\n10,430\n0.106\n335,000\n1.23\n8,370\n0.0932\n3,990\n0.0608\n"]], ["block_7", ["Estimate Rg for these polymers from these data. Use the data in Table 6.1 to compute Rgp and\nthus evaluate the coil expansion ratio as for each fraction. How does or vary with M, and how\ndoes this compare with the Flory\u2014Krigbaum prediction (Equation 7.7.10 and Equation\n7.7.12)?\nDiblock copolymers can readily form spherical micelles in a solvent that does not dissolve\none block. A typical aggregation number for such a micelle might be 100 individual\npolymers, with the inner \u201ccore\u201d formed of the 100 insoluble blocks and little or no solvent.\nImagine a solution of a block copolymer with M 100,000 and a concentration of 0.01 g/mL,\nwhich forms micelles upon cooling below some critical micelle temperature. Estimate the\nratio of viscosity of the solution before and after micellization and also the ratio of the\nhydrodynamic radius before and after micellization. Assume that both blocks are made of\npolymers with flexibilities similar to polystyrene.\nWhen the mutual diffusion coefficient is measured for dilute polymer solutions, for example\nby dynamic light scattering, it is found that the concentration dependence of Dm is linear with\nc, but the slope can be either positive or negative. Furthermore, in a good solvent the slope is\nusually positive, but in a theta solvent it is negative. This behavior is consistent with Equation\n9.5.18, but the dependence on solvent quality is not transparent. Show that, in fact,\n"]], ["block_8", ["M (g/mol)\n[\u20190] (dL/g)\nM (g/mol)\n[17] (dL/g)\n"]], ["block_9", ["Evaluate the data in the previous problem using this approach, and relate the new parameters\nto [n] and kH. What approximation is necessary to make the Huggins extrapolation and the\nSchulze\u2014Blaschke extrapolation equivalent? It is claimed in the literature that the Schulze\u2014\nBlaschke extrapolation is valid over a wider range of concentration; justify or refute this\nclaim.\nThe intrinsic viscosity of polystyrene in benzene at 25\u00b0C was measured for polymers with the\nfollowing molecular weights:\n"]], ["block_10", ["nip \ufb01ning!)\n. . .\n"]], ["block_11", ["C (g/mL)\n\u201dSp/C\n(1D nrelyc\n"]], ["block_12", ["0.005\n12.25\n11.890\n0.011\n12.7\n11.888\n0.015\n13.3\n12.127\n0.028\n14.4\n12.098\n0.034\n15.7\n12.581\n"]], ["block_13", ["Another method of extrapolation to obtain the intrinsic viscosity is due to Schulze and\nBlasehke:\n"]], ["block_14", ["dln 72\n"]], ["block_15", ["c\n2 23M :72\n"]], ["block_16", ["Dynamics of Dilute Polymer Solutions\n"]]], "page_389": [["block_0", [{"image_0": "389_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018Y. Kato, T. Kametani, K. Furukawa, and T. Hashimoto, J. Polym. Sci. Polym. Phys. Ed. 13, 1695 (1975).\n"]], ["block_2", ["Problems\n377\n"]], ["block_3", ["12.\nThe ratio of the radius of gyration, R3 (measured by light scattering as described in\nChapter 8) to the hydrodynamic radius, Rh (measured by dynamic light scattering) can be\na sensitive indicator of molecular conformation. Compare the value of this ratio for a\nhigh molecular weight linear chain in a theta solvent to that of a hard sphere. Now\nconsider a sixth-generation dendrimer and a regular four-arm star polymer with long arms;\nwhere would they rank relative to each other, and to the other two shapes? Explain your\nreasoning.\n13.\nDynamic light scattering measurements were made on a very dilute aqueous suspension\nof latex particles at a scattering angle of 45\u00b0; the measured decay rate implied a hydro-\ndynamic radius of 240 nm. When measurements were taken at a scattering angle of 90\u00b0 to\nconfirm this result, there was no significant scattering signal. Why? What can be inferred\nabout the particles based on this information? (Hint: some material in Section 8.6 may be\nhelpful.)\n14.\nThe overlap concentration (3*, which separates the dilute solution regime from the so-called\nsemidilute regime, can be estimated by space-filling arguments, as the concentration where\nthe individual coil-volumes begin to fill space. Derive the expression for c* in terms of R3\nand M and indicate how 6* scales with M in good and theta solvents. Alternatively, 6* can be\nestimated from the dilute solution viscosity expansion, such that 6* 1/[17]. Use the Flory\u2014\u2014\nFox relation to relate these two definitions, that is, find the proportionality constant between\n6* and 1/[17] that makes the two definitions equivalent.\n15.\nUse the model for the size exclusion of a spherical solute molecule in a cylindrical capillary\nto calculate KGPC for a selection of R/a values, which are compatible with Figure 9.20.\nPlot your values on a photocopy or tracing of Figure 9.20. On the basis of the comparison\nbetween these calculated points and the line in Figure 9.20 drawn on the basis of a statistical\nconsideration of chain exclusion, criticize or defend the following proposition: There is\nnot much difference between the K values calculated by the equivalent sphere and\nstatistical models. The discrepancy between various experimental methods of evaluating\n(a)\nis\nmuch\ngreater\nthan\nthe\ndifferences\narising\nfrom\ndifferent\nmodels.\nEven\nfor\nrandom coil molecules, the simple equivalent sphere model is acceptable for qualitative\ndiscussions of VR.\n16.\nSEC measurements are now often made with both an RI and an LS detector. The former\nresponds to the change in solution RI, n, as polymer elutes; It may be taken to be linear in\nconcentration at these dilute concentrations and independent of M. The latter either measures\n"]], ["block_4", ["17.\nBoth preparative and analytical GPC were employed to analyze a standard (NBS 706)\npolystyrene samplet. Fractions were collected from the preparative column, the solvent\nwas evaporated away, and the weight of each polymer fraction was obtained. The molecular\nweight of each fraction was obtained using an analytical gel permeation chromatograph. The\nfollowing data were obtained (mass in milligrams and M x 10\"4 g/mol):\n"]], ["block_5", ["IS at a series of scattering angles simultaneously, or at one very low angle. For each slice, 1', of\nthe chromatogram, one now has two pieces of data: ni and 13... Show how to compute MW and\nMn from these data (1' might easily run from 1 to 1000). What do you need to know in order to\nget absolute M averages (i.e., without calibrating the columns)? Can you suggest a way to get\n(an/ac without making any additional measurements?\n"]], ["block_6", ["where B is the second virial coefficient and 172 is the partial specific volume of the polymer\n(which we can take to be given by V2/M). To do this, start with Equation 9.5.13, and work to\nreplace (duz/dc) with arr/ac, following the discussion surrounding Equation 8.4.16 and\nEquation 8.4.17.\n"]]], "page_390": [["block_0", [{"image_0": "390_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "390_1.png", "coords": [27, 55, 350, 224], "fig_type": "figure"}]], ["block_2", ["378\nDynamics of Dilute Polymer Solutions\n"]], ["block_3", ["References\n"]], ["block_4", [".._.._..._.,_..._. PPPT\u2018QF\u2019\n"]], ["block_5", ["Calculate MW and M,1 and the ratio Mw/Mn for the original polymer. Also evaluate the ratio\nMW/Mn for the individual fractions. Comment on the significance of MW/M,1 for both the\nfractionated and unfractionated polymer.\n18.\nA polystyrene sample was prepared by living anionic polymerization (recall Chapter 4), with\nMW 34,500. The polydispersity was measured by four different techniques. Matrix\u2014assisted\nlaser desorption/ionization (MALDI) mass spectrometry (Chapter 1) gave 1.016. Tempera\u2014\nture gradient interaction chromatography (TGIC, a higher resolution technique than SEC)\ngave 1.020. Standard SEC with an R1 detector and calibration with PS standards gave 1.05.\nHowever, SEC on the same instrument but using an LS detector to obtain MW of each slice\ngave a value of 1.005. The Poisson distribution (see Chapter 4) predicts a polydispersity of\n1.003 for an ideal living polymerization, so the small value obtained by LS detection is at\nleast conceivable. However, there is good reason to believe that both MALDI and TGIC are\nmore accurate. Explain why using SEC with R1 detection gives a polydispersity that is too\nlarge, and why the LS detection gives a value that is too small.\n"]], ["block_6", ["3.\nLauffer, M.A., J. Chem. Educ, 58, 250 (1981).\n4.\nPerrin, F., J. Phys. Radium, 7, 1 (1936).\n5.\nSimha, R., J. Phys. Chem, 44, 25 (1940).\n6.\nStaudinger, H. and Heuer, W., Ber. 63, 222 (1930).\n7.\nMark, H., Der Feste Korper, Hirzel, Leipzig, 1938; R. Houwink, J. Prakt. Chem, 157, 15 (1941).\n8.\nBrandrup, J. and Immergut, E.H. (Eds), Polymer Handbook, 3rd ed., Wiley, New York, 1989.\n"]], ["block_7", ["1.\nStokes, G., Trans. Cambridge Phil. Soc., 8, 287 (1847); 9, 8 (1851).\n2.\nEinstein, A., Arm. Physik, 19, 289 (1906); 34. 591 (1911).\n"]], ["block_8", ["Kurata, M. and Stockmayer, W.H., Fortschr. Hochpolym.-Forsch., 3, 196 (1963).\nPoiseuille, J.L., Compres Rendas, 11, 961, 1041 (1840); 12, 112 (1841).\nPick, A., Ann. Physik, 170, 59 (1855).\nCrank, 1., The Mathematics of Dz\ufb01aslon, 2nd ed., Clarendon Press, Oxford, 1975.\nKirkwood, J.G. and Riseman, J., J. Chem. Phys., 16, 565 (1948).\nYau, W.W., Kirkland, J.J., and Bly, D.D., Modem Size Exclusion Chromatography, Wiley, New York,\n1979.\n"]], ["block_9", ["Fraction\nMass\nMn\nMw\nFraction\nMass\nMn\nMW\n"]], ["block_10", ["6\n2\n109\n111\n19\n42\n9.14\n9.35\n7\n8\n90.8\n92.5\n20\n30\n7.52\n7.68\n"]], ["block_11", ["8\n20\n76.7\n78.0\n21\n28\n6.16\n6.28\n"]], ["block_12", ["9\n42\n62.3\n63.5\n22\n18\n5.12\n5.22\n10\n64\n51.5\n52.5\n23\n12\n4.09\n4.18\n"]], ["block_13", ["13\n110\n28.7\n29.3\n26\n5\n2.01\n2.06\n14\n110\n23.3\n23.8\n27\n4\n1.52\n1.56\n"]], ["block_14", ["16\n86\n15.9\n16.3\n29\n2\n0.83\n0.85\n"]], ["block_15", ["17\n68\n13.0\n13.3\n30\n1\n0.59\n0.61\n18\n54\n11.0\n11.2\n"]], ["block_16", ["11\n84\n41.7\n42.5\n24\n8\n3.33\n3.40\n"]], ["block_17", ["12\n102\n34.7\n35.4\n25\n6\n2.63\n2.69\n"]], ["block_18", ["15\n96\n18.9\n19.4\n28\n3\n1.13\n1.16\n"]], ["block_19", [{"image_2": "390_2.png", "coords": [65, 67, 333, 224], "fig_type": "table"}]]], "page_391": [["block_0", [{"image_0": "391_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Berne, B]. and Pecora, Light Scattering, Wiley, New York, 1976.\nChu, B., Laser 2nd ed., Academic Press, San Diego, 1991.\nFlory, P.J., Principles of Polymer Chemistry, Cornell University Press, Ithaca, \nFujita, H., Polymer Solutions, Elsevier, Amsterdam, 1990.\nTanford, C., Physical Chemistry of Macromolecules, Wiley, New York, 1961.\nVan Holde, K.E., Physical Biochemistry, Prentice\u2014Hall, New Jersey, 1971.\nYamakawa, H., Modern Theory of Polymer Solutions, Harper & Row, New York, 1971.\nYau, W.W., Kirkland,J.J., and Bly, D.D., Modern Size Chromatography, \n"]], ["block_2", ["Further Readings\n379\n"]], ["block_3", ["Further Readings\n"]]], "page_392": [["block_0", [{"image_0": "392_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this chapter we consider one of the three general classes of polymers in the solid state: infinite\nnetworks. The other two categories, glassy polymers and semicrystalline polymers, will be taken\nup in Chapter 12 and Chapter 13, respectively. We will shortly define the term network more\nprecisely, but we have in mind a material in which covalent bonds (or other strong associations)\nlink different chain molecules together to produce a single molecule of effectively infinite\nmolecular weight. These linkages prevent flow and thus the material is a solid. There are two\nimportant subclasses of network materials: elastomers and thermosets. An etastomer is a cross-\nlinked polymer that undergoes the glass transition well below room temperature; consequently,\nthe solid is quite soft and deformable. The quintessential everyday example is a rubber band. Such\nmaterials are usually made by cross-linking after polymerization. A thermoset is a polymer in\nwhich multifunctional monomers are polymerized or copolymerized to form a relatively rigid\nsolid; an epoxy adhesive is a common example. In this chapter we will consider both elastomers\nand thermosets, but with an emphasis on the former. The reasons for this emphasis are that the\nphenomenon of rubber elasticity is unique to polymers and that it is an essential ingredient in\nunderstanding both the viscoelasticity of polymer liquids (see Chapter 11) and the swelling of\nsingle chains in a good solvent (see Chapter 7). In the first two sections we examine the two\ngeneral routes to chemical formation of networks: cross-linking of preformed chains and poly-\nmerization with multifunctional monomers. In Section 10.3 through Section 10.6 we describe\nsuccessively elastic deformations, thermodynamics of elasticity, the \u201cideal\u201d molecular description\nof rubber elasticity, and then extensions to the idealized theory. In Section 10.7, we consider the\nswelling of polymer networks with solvent.\n"]], ["block_2", ["1.\nStrand. A strand is a section of polymer chain that begins at one junction and ends at another\nwithout any intervening junctions.\n2.\nJunction. A junction is a cross-link from which three or more strands emanate. The function-\nah\u2019ty of the junction is the number of strands that are connected; in the case of the random\ncross-linking pictured in Figure 10.1 the functionality is usually four. Note that a cross-link\nmight simply connect two chains, but it would not be a junction until it becomes part of an\ninfinite network.\n"]], ["block_3", ["Figure 10.1 provides a pictorial representation of a network polymer. In panel (a), there is a\nschematic representation of a collection of polymer chains, which could be either in solution or in\nthe melt. In panel (b), a certain number of chemical linkages have been introduced between\nmonomers on different chains (or on the same chain). If enough such cross-[inks are created, it\nbecomes possible to start at one surface of the material and trace a course to the far side of the\nmaterial by passing only along the covalent bonds of chain backbones or cross-links. In such a\ncase an infinite network is formed, and we can say that the covalent structure percotates through\nthe material. The network consists of the following elements, as illustrated in Figure 10.2:\n"]], ["block_4", ["10.1.1\nDefinitions\n"]], ["block_5", ["10.1\nFormation of Networks by Random Cross-Linking\n"]], ["block_6", ["Networks, Gels, and Rubber Elasticity\n"]], ["block_7", ["10\n"]], ["block_8", ["381\n"]]], "page_393": [["block_0", [{"image_0": "393_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "393_1.png", "coords": [28, 412, 299, 651], "fig_type": "figure"}]], ["block_2", [{"image_2": "393_2.png", "coords": [28, 46, 308, 202], "fig_type": "figure"}]], ["block_3", [{"image_3": "393_3.png", "coords": [29, 195, 304, 350], "fig_type": "figure"}]], ["block_4", ["Figure 10.2\nSchematic illustration of network elements defined in the text.\n\u201834?\n"]], ["block_5", ["(b)\n"]], ["block_6", ["Figure 10.1\nSchematic illustration of (a) an uncross\u2014linked melt or concentrated solution of \ufb02exible chains\nand (b) the same material after cross-links are introduced.\n"]], ["block_7", ["382\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_8", ["\\\nR\\\nLoop\n"]], ["block_9", ["Junction\nStrand\n/\n"]], ["block_10", ["Dangling end\n"]]], "page_394": [["block_0", [{"image_0": "394_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We now consider the following question: given a collection of polymer chains, how many random\ncross\u2014links need to be introduced before a network will be formed? For simplicity, assume that all\nchains have the same degree of polymerization N, and that all monomers are equally likely to react.\nWe will give examples of cross\u2014linking chemistry in a moment, but for now we assume we can\nmeasure the extent of reaction, p, defined as the fraction of monomers that participate in cross-\nlinks. Suppose we start on a chain selected at random and find a cross\u2014link; we now use it to cross\nover to the next chain. What is the probability that, as we move along the second chain, we will\nfind a second cross\u2014link? It is simply given by (N 1)p eeNp. The probability of being able to hop\nfrom chain to chain x times in succession is therefore (Np)? (Recall that the probability of a series\nof independent events is given by the product of the individual probabilities.) For a network to be\nformed, we need this probability to be 21 as x \u2014\u2014> 00, and therefore we need Np 2 1. Conversely, if\nNp < 1, (Np)x \u2014> 0 as x \u2014> 00. Consequently, the critical extent of reaction, pc, at which an infinite\nnetwork \ufb01rst appears, the gel point, is given by\n"]], ["block_2", ["The apparently synonymous terms network, infinite network, and gel have all appeared so far and\nit is time to say how we will use these terms from now on. We have used network and infinite\nnetwork interchangeably; the modifier in\ufb01nite just serves to emphasize that the structure percolates\nthroughout a macroscopic sample and from now on we will omit it. The term gel is somewhat more\nproblematic, as it is used by different workers in rather disparate ways. We will henceforth use it to\nrefer to a material that contains a network, whereas the term network refers to the topology of the\nunderlying molecular structure. Often, an elastomeric material containing little or no sol fraction is\ncalled a rubber, whereas a material containing an equivalent network structure plus a significant\namount of solvent or low\u2014molecular\u2014weight diluent would be called a gel.\n"]], ["block_3", ["This beautifully simple result indicates how effective polymers can be at forming networks; a polymer\nwithN m 1000 only needs an average of 0.1% of the monomers to react to reach the gel point. Note that\nEquation 10.1.1 probably underestimates the true gel point because some fraction of cross-linking\nreactions will result in the formation of loops, which will not contribute to network formation.\nAny real polymer will be polydisperse, so we should consider how this affects Equation 10.1.1.\nLet us return to our first chain, find the cross\u2014link, and then ask, what is the average length of the\nnext chain? As the cross\u2014linking reaction was assumed to be random, then the chance that the next\nchain has degree of polymerization N,- is given by the weight fraction of Ni-mers, w,\u2014. In other\nwords, the probability that the neighboring monomer that forms the cross\u2014link belongs to a chain of\nlength N,- is proportional to N,\u2014. (To see this argument, consider a trivial example: the sample\n"]], ["block_4", ["3.\nDangling end. The section of the original polymer chain that begins at one chain terminus and\ncontinues to the first junction forms a dangling end. Because it is free to relax its conformation\nover time, it does not contribute to the equilibrium elasticity of the network, and as such it can\nbe viewed as a defect in the structure.\n4.\nLoop. Another defect is a loop, a section of chain that begins and ends at the same cross\u2014link,\nwith no intervening junctions. A loop might be formed by an intramolecular cross-linking\nreaction. Again, as with the dangling end, the loop can relax its conformation (at least in part)\nand is thus not fully elastically active.\n5.\nSolfraction. It is not necessary that every original polymer chain be linked into the network; a\ngiven chain may have no cross-links or it may be linked to a finite number of other chains to\nform a cluster. In either case, if the material were placed in a large reservoir of a good solvent\nthe sol fraction could dissolve, whereas the network or gel fraction could not. Thus the sol\nfraction contains all the extractable material, including any solvent present.\n"]], ["block_5", ["10.1.2\nGel Point\n"]], ["block_6", ["Formation of by Random Cross-Linking\n383\n"]], ["block_7", [{"image_1": "394_1.png", "coords": [37, 512, 127, 551], "fig_type": "molecule"}]], ["block_8", ["1\nl\npc\ufb01N\u2014INN\n(10.1.1)\n"]]], "page_395": [["block_0", [{"image_0": "395_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "395_1.png", "coords": [33, 94, 220, 150], "fig_type": "molecule"}]], ["block_2", [{"image_2": "395_2.png", "coords": [34, 586, 180, 619], "fig_type": "molecule"}]], ["block_3", ["This is a factor of 6.7 less than the stated value of p 0.003, so we may be reasonably confident\nthat the sample has passed the gel point.\nFor\nan\nindividual\nchain\nto\nbe\nuntouched,\nevery\nmonomer\nmust\nbe\nunreacted.\nThe\nprobability for each monomer to be unreacted is 1 p 0.997 and for a chain of N monomers\n"]], ["block_4", ["and thus the critical extent of reaction is determined by the weight-average degree of polymer\u2014\nization, NW.\nExamples of postpolymerization cross\u2014linking reactions are many. Free-radical initiators such\nas peroxides (see Chapter 3) can be used to cross\u2014link polymers with saturated structures (i.e., no\ncarbon\u2014carbon double bonds), such as polyethylene or poly(dimethylsiloxane). Alternatively,\nhigh-energy radiation can be utilized for the same purpose. A prime example occurs in integrated\ncircuit fabrication, where electron beam or UV radiation can be used to cross\u2014link a particular\npolymer (called a negative resist) in desired spatial patterns. The uncross-linked polymer is then\nwashed away, exposing the underlying substrate for etching or deposition. (In contrast, some\npolymers such as poly(methyl methacrylate) degrade rapidly on exposure to high-energy radiation,\nthereby forming a positive resist.) Of course, the classic example of cross\u2014linking is that of\npolydienes cross-linked in the presence of sulfur. The use of sulfur dates back to 1839 and the\nwork of Goodyear in the United States [1] and Macintosh and Hancock in the UK. The polymer of\nchoice was natural rubber, a material extracted from the sap of rubber trees; the major ingredient is\nCiS-1,4 polyisoprene. This basic process remains the primary commercial route to rubber materials,\nespecially in the production of tires, and the cross-linking of polydienes is generically referred to as\nvulcanization. Remarkably, perhaps, the detailed chemical mechanism of the process remains\nelusive. For some time a free-radical mechanism was suspected, but current thinking favors an\nionic route, as shown in Figure 10.3. The process is thought to proceed through formation of a\nsulfonium ion, whereby the naturally occurring eight\u2014membered sulfur ring, 83., becomes polarized\nor opened (Reaction A). The next stage is abstraction of an allylic hydrogen from a neighboring\nchain to generate a carbocation (Reaction B), which subsequently can react with sulfur and cross\u2014\nlink to another chain (Reaction C). A carbocation is regenerated, allowing propagation of the cross\u2014\nlinking process (Reaction D). Termination presumably involves sulfur anions. In practice, the rate\nof vulcanization is greatly enhanced by a combination of additives, called accelerators and\nactivators. Again, the mechanisms at play are far from fully understood, although the technology\nfor producing an array of rubber materials with tunable properties is highly developed.\n"]], ["block_5", ["A sample of polyisoprene with MW: 150,000 is vulcanized until 0.3% of the double bonds are\nconsumed, as determined by spectroscopy. Do you expect this sample to have formed a network,\nand what is the probability of finding a polyisoprene chain that is untouched by the reaction?\n"]], ["block_6", ["The nominal monomer molecular weight for polyisoprene is 68 g/mol, so for this sample the\ncritical extent of reaction estimated by Equation 10.1.2 is\n"]], ["block_7", ["contains 1 mole of chains of length 100 and 1 mole of chains of length 200. Any monomer selected\nat random has a probability of 2/3 to be in a chain of length 200, and 1/3 to be in chain of length\n100; 2/3 and 1/3 correspond to the weight fractions.) The critical probability therefore becomes\n"]], ["block_8", ["Solution\n"]], ["block_9", ["Example 10.1\n"]], ["block_10", ["384\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_11", ["1\n68\n\u20180c \nN,\n150,000\n"]], ["block_12", ["P\n"]], ["block_13", ["1\nC:\n00\ng 001\n:_1_\n(10.1.2)\nXmas-\ufb02 1)\nZw.-N.-\n"]], ["block_14", ["\u00a321\n121\n"]], ["block_15", ["NW\n"]]], "page_396": [["block_0", [{"image_0": "396_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "396_1.png", "coords": [33, 531, 110, 589], "fig_type": "molecule"}]], ["block_2", ["we must raise 0.997 to the Nth power. For simplicity, we assume all chains to have the same\nN=150,000/68:2200; then (0.997)2200e0.0013 or there is about 0.1% chance that a chain\nis untouched.\n"]], ["block_3", ["Figure 10.3\nPossible mechanism for vulcanization of 1,4-polybutadiene with sulfur, following Odian.\n(From Odian, G., Principles of Polymerization, 2nd ed., Wiley, New York, 1981.)\n"]], ["block_4", ["Formation of Networks by Random Cross-Linking\n385\n"]], ["block_5", [{"image_2": "396_2.png", "coords": [42, 45, 251, 184], "fig_type": "figure"}]], ["block_6", [{"image_3": "396_3.png", "coords": [45, 46, 169, 230], "fig_type": "figure"}]], ["block_7", ["3+\n5*\ns,s, or s; + 3,:\n"]], ["block_8", [{"image_4": "396_4.png", "coords": [57, 117, 196, 592], "fig_type": "figure"}]], ["block_9", [{"image_5": "396_5.png", "coords": [69, 154, 215, 406], "fig_type": "figure"}]], ["block_10", ["38\n"]]], "page_397": [["block_0", [{"image_0": "397_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["can lead to network formation.\n"]], ["block_2", ["2. AA and Bfor BB and Ar:\n"]], ["block_3", ["3. AB with either AA and Bfor BB and Ar:\n"]], ["block_4", ["4. A, and Bf:\n"]], ["block_5", ["In this section we consider the other general approach to network formation or gelation, using\npolymerization of multifunctional monomers. Multifunctional, as noted in Chapter 2, means\nfunctionality greater than 2. We will build on the material in that chapter by considering step-\ngrowth or condensation polymerization of monomers containing A and B reactive groups. The\nresulting thermosets are widely used as engineering materials because their mechanical properties\nare largely unaffected by temperature variation.\nFor simplicity, we assume that the reaction mixture contains only A and B as reactive groups,\nbut that either one (or both) of these is present (either totally or in part) in a molecule that contains\nmore than two of the reactive groups. We use f to represent the number of reactive groups in a\nmolecule when this quantity exceeds 2 and represent a multifunctional molecule as Af or Bf. For\nexample, if A were a hydroxyl group, a triol would correspond tof3. Several reaction possibil-\nities (all written forf3) come to mind in the presence of multifunctional reactants, as shown in\nFigure 10.4. The lower case \u201ca\u201d and \u201cb\u201d refer to the corresponding groups that have reacted.\nThe third reaction is interesting inasmuch as either the AA or BB monomer must be present to\nproduce cross-linking. Polymerization of AB with only Af (or only Bf) introduces a single branch\npoint, but no more, since all chain ends are unsuited for further incorporation of branch points.\nIncluding the AA or BB molecule reverses this. The bb unit that accomplishes this is underlined.\nWhat we seek next is a quantitative relationship among the extent of the polymerization reaction,\nthe composition of the monomer mixture, and the gel point. We shall base our discussion on the\nsystem described by the first reaction in Figure 10.4; other cases are derived by similar methods (see\n"]], ["block_6", ["Figure 10.4\nPossible reaction schemes for monomer mixtures containing A and B functional groups that\n"]], ["block_7", ["386\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_8", ["1. AA and BB plus either Aror Bf:\n"]], ["block_9", ["10.2\nPolymerization with Multifunctional Monomers\n"]], ["block_10", [{"image_1": "397_1.png", "coords": [52, 513, 312, 571], "fig_type": "molecule"}]], ["block_11", [{"image_2": "397_2.png", "coords": [61, 412, 258, 493], "fig_type": "figure"}]], ["block_12", [{"image_3": "397_3.png", "coords": [61, 411, 255, 504], "fig_type": "molecule"}]], ["block_13", ["avw\nabvw\nA3 + a3\n\u2014\u2014-- A\u2014<\nba\u2014<\nbwv\nab\u2014<\nab\u2014<\nbvw\nbvw\n"]], ["block_14", ["baavw\nbaab-<\nbaavw\nAA + B3\n-\u2014\u2014*\nAab\nbaavw\nbaab\u2014<\nbaavw\nbaa\nbaavw\n"]], ["block_15", [{"image_4": "397_4.png", "coords": [71, 588, 242, 633], "fig_type": "molecule"}]], ["block_16", ["abbaabbvw\nAA + BB + As \u2014\u2014a- Aabbaabba\u2014<\nabbw.\nabbaabba\n"]], ["block_17", ["ababaww\nAB + BB + A3 \u2014I- Abababa\navw\nababa_b_9ababa \u2014<\n"]], ["block_18", [{"image_5": "397_5.png", "coords": [84, 524, 301, 555], "fig_type": "molecule"}]], ["block_19", [{"image_6": "397_6.png", "coords": [90, 350, 316, 401], "fig_type": "molecule"}]], ["block_20", [{"image_7": "397_7.png", "coords": [98, 416, 226, 484], "fig_type": "molecule"}]], ["block_21", ["abbaaww\n"]], ["block_22", ["awv\n"]]], "page_398": [["block_0", [{"image_0": "398_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["There are two additional useful parameters that characterize the reaction itself:\n"]], ["block_2", ["The methods we consider were initially developed by Stockmayer [2] and Flory [3] and have been\napplied to a wide variety of polymer systems and phenomena. Our approach proceeds through two\nstages: first we consider the probability that AA and BB polymerize until all chain segments are\ncapped by an Af monomer, and then we consider the probability that these are connected together\nto form a network. The actual molecular processes occur at random and not in this sequence, but\nmathematical analysis is more feasible if we consider the process in stages. As long as the same\nsort of structure results from both the random and the subdivided processes, this analysis is valid.\nThe arguments we employ are statistical, so we recall that the probability of a functional group\nreacting is given by the fraction of groups that have reacted at any point and that the probability of\na sequence of events is the product of their individual probabilities (as used in developing Equation\n10.1.1). As in Chapter 2 and Chapter 3, we continue to invoke the principle of equal reactivity, that\nis, that functional group activity is independent of the size of the molecule to which the group is\nattached. One additional facet of this assumption that enters when multifunctional monomers are\nconsidered is that all A groups in Af are of equal reactivity.\nNow let us consider the probability that a section of polymer chain is capped at both ends by\npotential branch points:\n"]], ["block_3", ["1.\nThe first step is the condensation of a BB monomer with one of the A groups of an Af molecule:\nSince all A groups have the same reactivity by hypothesis, the probability of this occurrence is\nsimply p.\n2.\nThe terminal B group reacts with an A group from AA rather than Af:\n"]], ["block_4", ["1.\nThe extent of reaction p is based on the group present in limiting amount. For the system under\nconsideration, p is therefore the fraction of A groups that have reacted. (Note that this p is\nslightly different from p in Section 10.1.)\n2.\nThe probability that a chain segment is capped at both ends by a branch unit is described by the\nbranching coe\ufb02icient or. The branching coefficient is central to the discussion of network\nformation, as the occurrence or nonoccurrence of gelation depends on what happens after\ncapping a section of chain with a potential branch point.\n"]], ["block_5", ["1.\nThe ratio of the initial number of A to B groups, 123/123,, defines the factor r, as in Equation\n2.7.1. The total number of A groups from both AA and Af is included in this application of r.\n2.\nThe fraction of A groups present in mulifunctional molecules is defined by the ratio\n"]], ["block_6", ["The fraction of unreacted B groups is rp, so this gives the probability of reaction for B. Since p\nis the fraction of A groups on multifunctional monomers, rp must be multiplied by\n1 p to\ngive the probability of B reacting with an AA monomer. The total probability for the chain\nshown is the product of the probabilities until now: p[rp(l p)].\n3.\nThe terminal A groups react with another BB:\n"]], ["block_7", ["Problem 3). To further specify the system, we assume that A groups limit the reaction and that B\ngroups are present in excess. Two parameters are necessary to characterize the reaction mixture:\n"]], ["block_8", ["Polymerization with Multifunctional Monomers\n387\n"]], ["block_9", ["10.2.1\nCalculation of the Branching Coefficient\n"]], ["block_10", ["The probability of this step is again p, and the total probability is p[rp(1 p)p].\n"]], ["block_11", ["Af_1abbaA + BB \u2014> Af_1abbaabB\n"]], ["block_12", ["Af_1abB + AA -> Af_1abbaA\n"]], ["block_13", ["__ \n#\nVA([O[31)\n(10.2.1)\n"]]], "page_399": [["block_0", [{"image_0": "399_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We have now completed the first (and harder) stage of the problem we set out to consider: we\nknow the probability that a chain is capped at both ends by potential branch points. The second\nstage of the derivation considers the reaction between these chain ends via the remaining f\u20141\nreactive A groups. (By hypothesis, the mixture contains an excess of B groups, so there are still\nunreacted BB monomers or other polymer chain segments with terminal B groups that can react\nwith the Afn 1 groups we have been considering.) By analogy with the discussion of the gel point in\nSection 10.1, we ask the question: if we choose an Af group at random, and follow this chain to\nanother Af group, what is the probability that we can continue in this fashion forever? If this\nprobability exceeds 1, we have a network, and the gel point corresponds to when it equals 1. The\nprobability of there being a strand, that is, a chain segment between two junctions, is a. When\nwe arrive at the next Af, there aref 1 chances to connect to a new strand and the probability of\nthere being a strand from any particular one of the f\u20141 groups is again 0:. Thus the total\nprobability of keeping going from each Af is just (f l)cr. If we want to connect x strands in\nsequence, the probability that we can is [(f 1)a]x. Just as in the argument preceding Equation\n10.1.1, therefore, the critical extent of reaction is simply given by\n"]], ["block_2", ["which can be compared directly with Equation 10.1.1. Whenever the extent of reaction, p, is such\nthat a > are, gelation is predicted to occur. Combining Equation 10.2.3 and Equation 10.2.4 and\n"]], ["block_3", ["4.\nAdditional AA and BB molecules condense into the chain to give a sequence of i bbaa units\n"]], ["block_4", ["The summation applies only to the quantity in brackets, since it alone involves i. Representing the\nbracketed quantity by Q, we note that 2:0 Q = 1 /(1 Q) (see Appendix) and therefore\n"]], ["block_5", ["As the branching coefficient gives the probability of a chain segment being capped by potential\nbranch points, the above development describes this situation:\n"]], ["block_6", ["We have just evaluated the probability of one such unit; the probability for a series of 1' units is\njust the product of the individual probabilities: p[rp(1 p)p]\u2019.\n5.\nThe terminal B groups react with an A group from a multifunctional monomer:\n"]], ["block_7", ["The probability of B reacting is rp and the fraction of these reactions that involve Af molecules\nis rpp. The probability of the entire sequence is therefore p[rp(1 p)p]\u2018rpp.\n6.\nIn the general expression above, i can have any value from 0 to 00, so the probability for all\npossibilities is the sum of the individual probabilities. Note that a different procedure is used\nfor compounding probabilities here: the sum instead of the product. This time we are interested\nin either i =0 or i 1 or i =2, and so forth, whereas previously we required the first A\u2014B\nreaction and the second A\u2014B reaction and the third A\u2014B reaction, etc.\n"]], ["block_8", ["388\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_9", ["10.2.2\nGel Point\n"]], ["block_10", [{"image_1": "399_1.png", "coords": [44, 279, 157, 314], "fig_type": "molecule"}]], ["block_11", ["a, \ufb02\n(10.2.4)\n"]], ["block_12", ["Af_1abbaabB + AA + BB \u2014>\u2014>\u2014> Af_1a(bbaa),obB\n"]], ["block_13", ["r1029\n0, 2\n1 rpm p)\n(10.2.3)\n"]], ["block_14", ["a Z rpzpppm mi\"\n(10.2.2)\ni:0\n"]], ["block_15", ["Af_]a(bbaa),-bB + Af \u2014> Afn1a(bbaa)ibbaAf_1\n"]]], "page_400": [["block_0", [{"image_0": "400_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a similar fashion (see Problem 3 for an example).\nEquation 10.2.5 is of considerable practical utility in view of the commercial importance of\nthree-dimensional polymer networks. Some reactions of this sort are carried out on a very large\nscale: imagine the consequences of having a polymer preparation solidify in a large and expensive\nreaction vessel because the polymerization reaction went a little too far. Considering this kind of\napplication, we might actually be relieved to know that Equation 10.2.5 errs in the direction of\nunderestimating the extent of reaction at gelation. This comes about because some reactions of the\nmultifunctional branch points result in intramolecular loops, which are wasted as far as network\nformation is concerned; the same comment applies to Equation 10.1.1. It is also not uncommon\nthat the reactivity of the functional groups within one multifunctional monomer decreases with\nincreasing p, which tends to favor the formation of linear structures over the branched ones.\nAs an example of the quantitative testing of Equation 10.2.5, consider the polymerization of\ndiethylene glycol (BB) with adipic acid (AA) in the presence of 1,2,3-propane tricarboxylic acid\n(A3). The critical value of the branching coefficient is 0.50 for this system by Equation 10.2.4. For\nan experiment in which r 0.800 and p 0.375, )9C 0.953 by Equation 10.2.5. The critical extent\nof reaction was found experimentally to be 0.9907, determined in the polymerizing mixture as the\npoint where bubbles fail to rise through it. Calculating back from Equation 10.2.3, the experimental\nvalue of pC is consistent with the value ac =0.578, instead of the theoretical value of 0.50.\n"]], ["block_2", ["It is apparent that numerous other special systems or effects could be considered to either broaden\nthe range or improve the applicability of the derivation presented. Our interest, however, is in\nillustrating concepts rather than exhaustively exploring all possible cases, so we shall not pursue\nthe matter of the gel point further here. Instead, we conclude this section with a brief examination\nof the molecular-weight averages in the system generated from AA, BB, and Af. For simplicity, we\nrestrict our attention to the case of 12% :12%. It is useful to define the average functionality of a\nmonomer (f) as\n"]], ["block_3", ["where n, and f,- are the number of molecules and the functionality of the ith component in the\nreaction mixture, respectively. The summations are over all monomers. If n is the total number of\nmolecules present at the extent of reaction )9 and no is the total number of molecules present\ninitially, then 2(n0 n) is the number of functional groups that have reacted and (f)no is the total\nnumber of groups initially present. Two conclusions immediately follow from these concepts:\n"]], ["block_4", ["rearranging gives the critical extent of reaction for gelation, pC as a function of the properties of the\nmonomer mixture r, p, and f:\n"]], ["block_5", ["Corresponding equations for any of the reaction schemes depicted in Figure 10.4 can be derived in\n"]], ["block_6", ["Polymerization with Multifunctional Monomers\n339\n"]], ["block_7", ["where ND is the number-average degree of polymerization, and\n"]], ["block_8", ["10.2.3\nMolecular-Weight Averages\n"]], ["block_9", [{"image_1": "400_1.png", "coords": [37, 449, 110, 524], "fig_type": "molecule"}]], ["block_10", ["2\n_\np \n(\u201d0\n\u20191)\n(10.2.8)\n(f>110\n"]], ["block_11", ["1\nPo \nx/r+rp(f\u20142)\n"]], ["block_12", ["Nn = @\n(10.2.7)\n11\n"]], ["block_13", ["Znsfs\n(f) \nZn.-\n"]], ["block_14", [{"image_2": "400_2.png", "coords": [52, 461, 107, 509], "fig_type": "molecule"}]], ["block_15", ["(10.2.5)\n"]], ["block_16", ["(10.2.6)\n"]]], "page_401": [["block_0", [{"image_0": "401_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This result is known as the Carothers equation [4]. It is apparent that this expression reduces to\nEquation 2.2.4 for the case of (f) 2, that is, the result for the most probable distribution in\npolycondensation reactions considered in Chapter 2. Furthermore, when (f) exceeds 2, as in the\nAA/BB/Af mixture under consideration, then Nn is increased over the value obtained at the same\np for (f) :2 2. A numerical example will help clarify these relationships.\n"]], ["block_2", ["The Carothers approach, as described above, is limited to the number-average degree of\npolymerization and gives no information concerning the breadth of the distribution. A statistical\napproach to the degree of polymerization yields expressions for both NW and N\u201c. Ref. [4] contains a\n"]], ["block_3", ["These results demonstrate how for a fixed extent of reaction, the presence of multifunctional\nmonomers in an equimolar mixture of reactive groups increases the degree of polymerization.\nConversely, for the same mixture a lesser extent of reaction is needed to reach a specific Nlrl with\nmultifunctional reactants than without them. Remember that this entire approach is developed for\nthe case of stoichiometric balance. If the numbers of functional groups are unequal, this effect\nworks in opposition to the multifunctional groups.\n"]], ["block_4", ["An AA, BB, and A3 polymerization mixture is prepared in which v51 :12% z: 3.00 mol, with 10%\nof the A groups contributed by A3. Use Equation 10.2.9 to calculate Nn for p =0.970 and p for\nNn=200. In each case, compare the results with what would be obtained if no multifunctional\nA were present.\n"]], ["block_5", ["Solve Equation 10.2.9 with ND 200 and (f) = 2.034:\n"]], ["block_6", ["Solve Equation 10.2.9 with Nn 200 and (f) 2:\n"]], ["block_7", ["For comparison, solve Equation 10.2.9 with p 0.970 and (f) 2:\n"]], ["block_8", ["Determine the average functionality of the mixture. The total number of functional groups is 6.00\nmol,\nbut\nthe\ntotal\nnumber\nof\nmolecules\ninitially\npresent\nmust\nbe\ndetermined.\nUsing\n"]], ["block_9", ["Solving Equation 10.2.9 with p = 0.970 and (f) 2.034:\n"]], ["block_10", ["3nAAA +2nAA 23.00 and 3nAAA/3 : 0.100, We ld that \u201dAA=1\u00b0350 and \u201dAAA=0\u00b01000\u00b0 Since\nnBB 1.500 the total number of moles initially present is no 1.350+ 0.100+ 1.500 = 2.950:\n"]], ["block_11", ["Elimination of n between these expressions gives\n"]], ["block_12", ["Solution\n"]], ["block_13", ["390\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_14", ["Example 10.2\n"]], ["block_15", [{"image_1": "401_1.png", "coords": [42, 517, 211, 550], "fig_type": "molecule"}]], ["block_16", ["1\n1\nP (\nN.) (\nzoo)\n"]], ["block_17", ["Nn \u20181\u2014p\"1\u20140.97\"\n33.3\n"]], ["block_18", ["Nn\n2\n73.8\n: 2 \u2014 0.970.034) :\n"]], ["block_19", ["2\nN\u201d\n2 -\u2014p(f)\n(10.2.9)\n"]], ["block_20", ["(f) \n2.950\n=2.034\n"]], ["block_21", ["_ _ \n\"\n(f)\n\u201c\u2018\n2.034\n= 0.978\n"]]], "page_402": [["block_0", [{"image_0": "402_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The chemistry underlying an epoxy adhesive is illustrated in Figure 10.5. An excess of epichloro\u2014\nhydrin is reacted with a diol to form a linear prepolymer, terminated at each end with epoxide\n"]], ["block_2", ["The value of a to be used in these expressions is given by Equation 10.2.3 for the specific mixture\nunder consideration. At the gel point ac=1/(f 1) according to Equation 10.2.4, and thus\nEquation 10.2.11 predicts that NW becomes infinite, whereas N\u201d remains finite. This is a very\nimportant point. It emphasizes that in addition to the network molecule, or gel fraction, of\nessentially infinite molecular weight, there are still many other molecules present at the gel\npoint, the sol fraction. The ratio Nw/Nn also indicates a divergence of the polydispersity as\na \u2014>ac. Expressions have also been developed to describe the distribution of molecules in the\nsol fraction beyond the gel point. We conclude this discussion with an example that illustrates\napplication of some of these concepts to a common household product.\n"]], ["block_3", ["from which it follows that\n"]], ["block_4", ["derivation of these quantities for the self\u2014polymerization of Af monomers. Although this speci\ufb01c\nsystem might appear to be very different from the one we have considered, the essential aspects of\nthe two different averaging procedures are applicable to the system we have considered as well.\nThe results obtained for the Af case are\n"]], ["block_5", ["OH\no\nO\nO\n,\n__\nn k0! + (n \u2014 1)HO/R\u2018OH\n\u2014O-H-+ MO\u2018R\u2019DNEHH\u2019IOf\n+ n HCl\n"]], ["block_6", ["Figure 10.5\nIllustration of an epoxy formulation. A prepolymer, formed by base-catalyzed Condensation\nof an excess of epichlorohydrin with bisphenol A, is cured by cross-linking with 4,4\u2019-methylene dianiline.\n"]], ["block_7", ["and\n"]], ["block_8", ["Example 10.3\n"]], ["block_9", ["Polymerization with Multifunctional Monomers\n391\n"]], ["block_10", [{"image_1": "402_1.png", "coords": [40, 205, 148, 243], "fig_type": "molecule"}]], ["block_11", ["1+ (.1:\nNW \n.\n.\nl-\u2014~a(f\u2014l)\n(10211)\n"]], ["block_12", ["2\nNn \n2 af\n(10.2.10)\n"]], ["block_13", ["NW\n1 +\n1 \n2\n\u2014- :(\na)(\naf/ )\n(10.2.12)\nNn\n1\u2014a(f\u2014\u20141)\n"]], ["block_14", [{"image_2": "402_2.png", "coords": [93, 473, 334, 553], "fig_type": "figure"}]], ["block_15", [{"image_3": "402_3.png", "coords": [94, 548, 287, 643], "fig_type": "figure"}]], ["block_16", ["H\u201d\nOH\nOH\nHEN/\n\\NHZ\n"]], ["block_17", [{"image_4": "402_4.png", "coords": [104, 549, 246, 602], "fig_type": "molecule"}]], ["block_18", ["Prepolymer\nm\nN\n+\nOH\nOH\n___)...\nN_Rn_N\n"]], ["block_19", ["R\u201d:\n4,4'-Methylene dianiline\n"]], ["block_20", ["Fl\u2019 =\nO\nO\nBisphenol A\n"]], ["block_21", [{"image_5": "402_5.png", "coords": [126, 425, 400, 474], "fig_type": "molecule"}]], ["block_22", [{"image_6": "402_6.png", "coords": [178, 486, 322, 547], "fig_type": "figure"}]], ["block_23", ["Prepolymer\n"]]], "page_403": [["block_0", [{"image_0": "403_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["For the remainder of this chapter we will emphasize elastomers rather than thermosets, and our\nprimary focus will be the elasticity of such network materials. The various elastic phenomena we\ndiscuss in this chapter will be developed in stages. We begin with the simplest case: a sample that\ndisplays a purely elastic response when deformed by simple elongation. On the basis of Hooke\u2019s\nlaw, we expect that the force of defonnation\u2014related to the stress\u2014and the distortion that\nresults\u2014related to the strain\u2014will be directly proportional, at least for small deformations. In\naddition, the energy spent to produce the deformation is recoverable: the material snaps back when\nthe force is released. We are interested in the molecular origin of this property for polymeric\nmaterials but, before we can get to that, we need to define the variables more precisely. One\ncautionary note is appropriate here. A full description of the elastic response of materials requires\ntensors, but we will avoid this complication by emphasizing one kind of deformation\u2014uniaxial\nextension\u2014and touching on another, shear.\nA quantitative formulation of Hooke\u2019s law is facilitated by considering the rectangular sample\nshown in Figure 10.6a. If a forcef is applied to the face of area A, the original length of the block\nL0 will be increased by AL. Now consider the following variations:\n"]], ["block_2", ["We can interpret the time for the bond to set as a time when the gel point is consistently exceeded,\nperhaps p m 0.7, so that the adhesive has solidified. The time to develop full mechanical strength\nre\ufb02ects the time required for p to approach 1.\n"]], ["block_3", ["Following the reaction scheme in Figure 10.5, the prepolymer has functionality 2 whereas the\ndiamine has functionality f: 4, so we will call the epoxide group \u201cB\u201d and the diamine A4. We\nnow need to find out which group is in excess, that is, to calculate the ratio r. The molecular weight\nof the diamine is 198 g/mol and that of the prepolymer is 914 g/mol. If we mix 1 g of the diamine\nwith 10 g of the prepolymer we have a molar ratio of (1/198):(10/914) or 000505200109. As there\nare four A groups per diamine and two B groups per prepolymer, the final ratio of A:B groups is\n0.0101:0.0109 or 0.93:1. Thus the A group is limiting the reaction, albeit only just.\nFrom Equation 10.2.1 we can see that p l, as all the A group are in A4 units. This also makes\nthe development of the branching coefficient quite simple, as every chain between two A4 groups\ncontains one and only one prepolymer (BB) unit. The addition of the first BB to an A4 group takes\nplace with probability p, and the addition of the subsequent A4 has probability rp. Thus a rpz,\nwhich we could also obtain from Equation 10.2.3 after inserting p: l. The critical extent of\nreaction corresponds to are :1/3 from Equation 10.2.4, and from Equation 10.2.5 we have\n"]], ["block_4", ["rings. For the example in Figure 10.5, the diol is based on bisphenol A. The prepolymer is then\nreacted (cured) with a multifunctional anhydride or amine (methyl dianiline in the figure) to form a\nhighly cross-linked material. Adapt the analysis in the preceding section to find the gel point for\nthis system, assuming that the two compounds are mixed in the weight ratio 1:10 diamine to\nprepolymer and that the prepolymer has n 4 (see Figure 10.5). Then interpret the statement found\nin the instructions for a typical \u201ctwo\u2014part\u201d epoxy that \u201cthe bond will set in 5 minutes, but that full\nstrength will not be achieved until 6 hours.\u201d\n"]], ["block_5", ["392\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_6", ["Solution\n"]], ["block_7", ["10.3\nElastic Deformation\n"]], ["block_8", ["1.\nImagine subdividing the block into two portions perpendicular to the direction of the force, as\nshown in Figure 10.6b. Each slice experiences the same force as before, and the same net\ndeformation results. A deformation AL/2 is associated with a slice of length L0/2. The same\n"]], ["block_9", [{"image_1": "403_1.png", "coords": [40, 326, 122, 363], "fig_type": "molecule"}]], ["block_10", ["\u2014\n1\nN 0 6\nPC\nx/3\u2014r\nN\n-\n"]]], "page_404": [["block_0", [{"image_0": "404_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "404_1.png", "coords": [29, 59, 272, 310], "fig_type": "figure"}]], ["block_2", ["With these considerations in mind, we write\n"]], ["block_3", ["(b)\n"]], ["block_4", ["Figure 10.6\n(a) A forcef applied to area A extends the length of the sample from L0 by an amount AL. Parts\n(b) and (0) illustrate the argument that f/A AL/LO.\n"]], ["block_5", ["(C)\n"]], ["block_6", ["Elastic \n393\n"]], ["block_7", [{"image_2": "404_2.png", "coords": [41, 183, 266, 421], "fig_type": "figure"}]], ["block_8", [{"image_3": "404_3.png", "coords": [42, 187, 269, 306], "fig_type": "figure"}]], ["block_9", [{"image_4": "404_4.png", "coords": [43, 52, 262, 184], "fig_type": "figure"}]], ["block_10", [{"image_5": "404_5.png", "coords": [43, 303, 273, 423], "fig_type": "figure"}]], ["block_11", ["argument could be applied for any number of slices; hence it is the quantity AL/Lo that is\nproportional to the force.\nImagine subdividing the face of the block into two portions of area A/2. A force only half as\nlarge would be required for each face to produce the same net distortion. The same argument\ncould be applied for any degree of subdivision; hence it is the quantity f/A that is proportional\nto AL/LO.\nThe force per unit area along the axis of the deformation is called the uniaxial tension or stress.\nWe shall use the symbol 0\u2018 as a shorthand replacement for f/A and attach the subscript t to\nsignify tension; we will use 0' for the shear stress, as in Chapter 9 and Chapter 11. The\nelongation expressed as a fraction of the original length, AL/LO, is called the strain. We shall\nuse a as the symbol for the resulting extensional strain to distinguish it from the shear strain (7)\nalso discussed in Chapter 9 and Chapter 11.\n"]], ["block_12", ["AL\n0, 2E8 25(3)\n(10.3.1)\n"]], ["block_13", ["Lo\nAL\n/\n-\"7\"\"'31\nI\nI\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nl\nI\nI\nl\n------\nI\ni\nr\n\u2014I\u2014\u2014+\nI\n"]], ["block_14", ["-------------\n"]], ["block_15", [":\nl\nI\nI\nA\n)\nI\nI\nI\n,\u2019\nI\nI I,I\nI;\n\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014 J\n"]], ["block_16", ["m-\u2018\u2014---------\n"]], ["block_17", ["r--\u2014--\u2014-\u2014--\u00a2\n"]], ["block_18", ["III I I I\n"]], ["block_19", ["Il\n"]], ["block_20", ["T\nl\nf\n"]], ["block_21", ["I\n\u2014-|\u2014\u2014>\nI\n"]], ["block_22", ["\\\n"]], ["block_23", ["\\x\n"]], ["block_24", ["I\n"]], ["block_25", ["L-\n\\._...\n"]], ["block_26", ["I\n"]], ["block_27", ["a\n"]], ["block_28", ["~\\\n"]], ["block_29", ["I\n\\I\n"]], ["block_30", ["-\u2014---\u2019\n"]], ["block_31", ["-\n"]]], "page_405": [["block_0", [{"image_0": "405_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "405_1.png", "coords": [24, 165, 169, 201], "fig_type": "molecule"}]], ["block_2", ["where the minus signs indicate that Ad and Ah are negative when AL is positive. The constant p is\nan important property of a material called Poisson\u2019s ratio; it may also be written as\n"]], ["block_3", ["For isotrOpic materials such as those we are considering in this chapter, the small strain elastic\nresponse can therefore be described by any two of the parameters of E, G, K, and v. For elastomers,\nwhere the volume change on deformation tends to be very small, v z 0.5 and E m 30. For example,\npolyisoprene has v 0.4999, so this approximation is excellent; in contrast, for metals, 1) typically\nlies between 0.25 and 0.35.\n"]], ["block_4", ["It is not particularly difficult to introduce thermodynamic concepts into a discussion of elasticity.\nWe shall not explore all of the implications of this development, but shall proceed only to the point\nof establishing the connection between elasticity and entr0py. Then in the next section we shall go\nfrom macrosc0pic thermodynamics to statistical thermodynamics, in pursuit of a molecular model\nto describe the elastic response of cross\u2014linked networks.\n"]], ["block_5", ["We begin by remembering the mechanical definition of work and apply that definition to the\nstretching process of Figure 10.6. Using the notation of Figure 10.6, we can write the increment of\nelastic work associated with an increment in elongation dL as\n"]], ["block_6", ["where V is the volume of the sample (see Problem 9). Thus, if the volume does not change on\nelongation, the factional contraction in each of the perpendicular directions is half the fractional\nincrease in length and V: 0.5. In general two parameters, for example E and v, are required to\ndescribe the response of a sample to tensile force. Poisson\u2019s ratio also provides a means to relate E\nto the shear modulus, G, and the compressional modulus, K:\n"]], ["block_7", ["It is necessary to establish some conventions conceming signs before proceeding further. When the\napplied force is a tensile force and the distortion is one of stretching,f, dL and dw are all defined to\nbe positive quantities. Thus dw is positive when elastic work is done on the system. The work done\nby the sample when the elastomer snaps back to its original size is a negative quantity.\n"]], ["block_8", ["where the proportionally constant E is called the tensile modulus or Young\u2019s modulus. Remember,\nit will be different for different substances and for a given substance at different temperatures,\nSince 8 is dimensionless, E has the same units as f/A, namely, force/lengthz, or N/m2(Pa) in the SI\nsystem. Note that for Equation 103.1 to be useful as a definition of E, the strain must be\nsufficiently small so that the stress remains proportional to the strain.\nThere is another aspect of tensile deformation to be considered. The application of a distorting\nforce not only stretches a sample, but also causes the sample to contract at right angles to the\nstretch. If d and I: represent the width and height of area A in Figure 10.6, both contract by the same\nfraction, a fraction that is related to the strain in the following way:\n"]], ["block_9", ["394\nNetworks. Gels, and Rubber Elasticity\n"]], ["block_10", ["10.4.1\nEquation of State\n"]], ["block_11", ["10.4\nThermodynamics of Elasticity\n"]], ["block_12", [{"image_2": "405_2.png", "coords": [41, 227, 133, 270], "fig_type": "molecule"}]], ["block_13", ["02(1 + v) = E\n(10.3.4a)\n"]], ["block_14", ["K3(1 2v) E\n(10.3.4b)\n"]], ["block_15", ["dw =d\n(10.4.1)\n"]], ["block_16", ["__=__..=,,_=pg\n(10.3.2)\n"]], ["block_17", ["l\n1 dV\n"]]], "page_406": [["block_0", [{"image_0": "406_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Note this is the same derivation that yields the important results V (80/6187 and S (80/87)},\nwhen no elastic work is considered; these will arise in the discussion of the glass transition in\nChapter 12.\n"]], ["block_2", ["thus establishing the desired connection between the stretching experiment and thermodynamics.\nSince G is a state variable and forms exact differentials, an alternative expression for dG is\n"]], ["block_3", ["Comparing Equation 10.4.10 and Equation 10.4.9 enables us to write\n"]], ["block_4", ["This relationship can be used to replace dq by TdS in Equation 10.4.3, since the infinitesimal\nincrements implied by the differentials mean that the system is only slightly disturbed from\nequilibrium and the process is therefore reversible:\ndUzT\ufb01\u2014p\ufb02Wf\ufb02.\n(was\n"]], ["block_5", ["We now turn to the Gibbs free energy G (recall the treatment of mixtures in Chapter 7)\ndefined as\n"]], ["block_6", ["where the enthalpy\n"]], ["block_7", ["Comparing Equation 10.4.8 with Equation 10.4.5 enables us to replace several of these terms\nbyf dL\ndG=V\u00ae\u2014S\ufb02%f\ufb02;\n(mam\n"]], ["block_8", ["A consistent sign convention has been applied to the pressurewvolume work term: a positive dV\ncorresponds to an expanded system, and work is done by the system to push back the surrounding\natmosphere.\nThe second law of thermodynamics gives the change in entrOpy associated with the isothermal,\nreversible absorption of an element of heat dq as\n"]], ["block_9", ["Combining the last two results and taking the derivative gives\n"]], ["block_10", ["The element of work is generally written \u2014p dV, where p is the external pressure, but with the\npossibility of an elastic contribution, it is\n\u2014p dV \u2014l\u2014 fdL. With this substitution, Equation 10.4.2\nbecomes\n\ufb027=dqmpdV+d\n0043\n"]], ["block_11", ["Thermodynamics \n395\n"]], ["block_12", ["The first law of thermodynamics defines the change dU in the internal energy of a system as the\nsum of the heat absorbed by the system, dq, plus the work done on the system, dw:\n"]], ["block_13", ["86\nf = (_.)\n(10.4.11)\n6L p,T\n"]], ["block_14", ["H U +pV\n(10.4.7)\n"]], ["block_15", ["8G\n66\n8G\nd6 (\u2014)\ndp + (\u2014\u2014)\ndT + (\u2014\u2014)\ndL\n(10.4.10)\n6p T,L\n6T p,L\n8L [3,1,\n"]], ["block_16", ["G H TS\n(10.4.6)\n"]], ["block_17", ["M\n24-\n144\ndS\nT\n(0\n)\n"]], ["block_18", ["dG=dU+pdV+Vdp\u2014TdS\u2014SdT\n(10.4.8)\n"]], ["block_19", ["dU=dq+mv\n0043\n"]], ["block_20", [{"image_1": "406_1.png", "coords": [128, 541, 259, 574], "fig_type": "molecule"}]]], "page_407": [["block_0", [{"image_0": "407_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "407_1.png", "coords": [31, 177, 172, 220], "fig_type": "molecule"}]], ["block_2", [{"image_2": "407_2.png", "coords": [34, 126, 168, 163], "fig_type": "molecule"}]], ["block_3", ["Although defined by analogy to an ideal gas, the justification for setting (are/amp; 0 cannot\nbe the same for an elastomer as for an ideal gas. All molecules attract one another and this\nattraction is not negligible in condensed phases (recall the cohesive energy density in Chapter 7).\nWhat the ideality condition requires in an elastomer is that there is no change in the enthalpy of the\nsample as a result of the stretching process. This has two implications. On the one hand, the\naverage energy of interaction between different molecules cannot change. For a given material this\nintermolecular contribution is determined primarily by the density, and therefore for a deformation\nthat does not change the volume it may be a good approximation. The intramolecular contribution\narises from the conformational energy of each chain, which is determined by the relative popula\u2014\ntion of trans and gauche conformers (recall Chapter 6). In fact, moderate changes in the end-to-end\ndistance of a chain can be accomplished with the expenditure of relatively little energy. For large\ndeformations, or for networks with strong interactions\u2014say, hydrogen bonds instead of dispersion\nforces\u2014the approximation of an ideal elastomer may be very poor. There is certainly an enthalpy\nchange associated with crystallization (see Chapter 13), so (are/amp; would not vanish if\nstretching induced crystal formation (which can occur, e.g., in natural rubber).\nWe have presented this development of the ideal elastomer in terms of the Gibbs free energy,\nwhich is generally the most appropriate for processes of importance in chemistry: p and T (and\n"]], ["block_4", ["Equation 10.4.12 shows that the force required to stretch a sample can be broken into two\ncontributions: one that measures how the enthalpy of the sample changes with elongation and\none that measures the same effect on entropy. The pressure of a system also re\ufb02ects two parallel\ncontributions, except that the coefficients are associated with volume changes. It will help to\npursue the analogy with a gas a bit further. For an ideal gas, the molecules are noninteracting and\nso it makes no difference how far apart they are. Therefore, for an ideal gas (\u00abBU/3107: 0 and the\nthermodynamic equation of state becomes\n"]], ["block_5", ["The left-hand side of this equation givesf according to Equation 10.4.11; therefore\n"]], ["block_6", ["This expression is sometimes called the equation of state for an elastomer, by analogy to\n"]], ["block_7", ["the thermodynamic equation of state for a \ufb02uid. Note the parallel roles played by length and\nvolume in these two expressions.\n"]], ["block_8", ["396\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_9", ["By analogy, an ideal elastomer is defined as one for which (\u00abWI/amp}: 0; in this case Equation\n10.4.13 reduces to\n"]], ["block_10", ["10.4.2\nIdeal Elastomers\n"]], ["block_11", [{"image_3": "407_3.png", "coords": [40, 73, 195, 110], "fig_type": "molecule"}]], ["block_12", ["(93\nf \n\u2014T(\u2014)\n(10.4.16)\n3L p,T\n"]], ["block_13", ["8H\n63\nf _.\u2014_ (_)\n\u2014- T(\u2014\u2014)\n(10.4.13)\n8L PtT\n6L p,T\n"]], ["block_14", [{"image_4": "407_4.png", "coords": [46, 357, 126, 396], "fig_type": "molecule"}]], ["block_15", ["We differentiate Equation 10.4.6 with respect to L, keeping p and T constant:\n"]], ["block_16", [{"image_5": "407_5.png", "coords": [46, 419, 117, 465], "fig_type": "molecule"}]], ["block_17", ["as\n_p (5)7\n(10.4.15)\n"]], ["block_18", ["3U\n83\n_\n= _\n_ :r __\n10.4.\n\u201d\n(at/L\n(at/l\n(\n14)\n"]], ["block_19", ["6G)\n(8H)\n(83)\n__\n:\n__\n\u2014\u2014T \n(10.4.12)\n(3L p,T\n3L p,T\n8L ptT\n"]]], "page_408": [["block_0", [{"image_0": "408_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 10.7\nStress at a constant length for natural rubber, at the indicated elongations, as a function of\ntemperature. Thermoelastic inversion occurs below about 10% elongation. (Data from Anthony, R.L., Caston,\nR.H., and Guth, E., J. Phys. Chem, 46, 826, 1942. With permission.)\n"]], ["block_2", ["and the criterion for an ideal elastomer becomes (8U/8L)T,V=0. Because the volume changes on\nelastomer deformation are typically so small, a deformation carried out at constant T and p is very\nclose to one done at constant T and V.\n"]], ["block_3", ["Before proceeding to the statistical theory of rubber elasticity, it is instructive to examine some of\nthe classical experiments conducted on rubbers. An example is shown in Figure 10.7, where the\ntensile stress (proportional tof) was measured as a function of temperature at the indicated constant\n"]], ["block_4", ["Thermodynamics of Elasticity\n397\n"]], ["block_5", ["number of moles) are the natural independent variables. However, in the majority of texts the\nHelmholtz free energy, A: U TS is employed, so it is worthwhile to take a moment and\ncompare the answers. For an experiment at constant temperature, we can write\n"]], ["block_6", ["which may then be compared to Equation 10.4.5 to yield\n"]], ["block_7", ["L\nAt both constant temperature and constant volume, therefore,\n"]], ["block_8", ["10.4.3\nSome Experiments on Real Rubbers\n"]], ["block_9", ["(\ufb01g\n220/0\n0B) N\nas 2.0 \n"]], ["block_10", ["U)\n(De\na)\n13%\n"]], ["block_11", [{"image_1": "408_1.png", "coords": [44, 174, 216, 205], "fig_type": "molecule"}]], ["block_12", ["3A\n6U\nBS\nf \n(Elm/E (\u201d\u00e9\u2014Elm\n"]], ["block_13", ["dA =d\u2014pdV\n(10.4.18)\n"]], ["block_14", ["d4 dU TdS\n(10.4.17)\n"]], ["block_15", [{"image_2": "408_2.png", "coords": [50, 328, 278, 448], "fig_type": "figure"}]], ["block_16", ["\u201840[\u2014\n"]], ["block_17", ["00\nI\n|\n|\n|\n0\n20\n40\n60\n80\n"]], ["block_18", ["3.0 \n"]], ["block_19", ["10 \n696\nW\n"]], ["block_20", ["Temperature (\u00b0C)\n"]], ["block_21", ["\u2014\nT(i)r,v\n(104.19)\n"]], ["block_22", ["38%\n"]]], "page_409": [["block_0", [{"image_0": "409_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "409_1.png", "coords": [32, 224, 144, 273], "fig_type": "molecule"}]], ["block_2", [{"image_2": "409_2.png", "coords": [32, 178, 126, 227], "fig_type": "molecule"}]], ["block_3", ["These expressions are useful because they permit extraction of information about S and U from\nthe measured behavior off. Figure 10.8a shows data forf versus elongation and the decomposition\ninto an entropic and an internal energy contribution, following Equation 10.4.20 and Equation\n10.4.21. Clearly at large elongation, the entropic part of the force dominates, but at low elongations\nthe internal energy contribution is larger. Again, however, this effect is largely eliminated by\nplotting the data at constant strain, as shown in Figure 10.8b. These results and many others\nconfirm, to a good approximation, that there is only a modest internal energy contribution to the\nforce for a deformation at constant volume.\nOne further example of a \u201cmodel\u2014free\u201d thermodynamic interpretation of rubber elasticity is\ngiven by the temperature increase observed in adiabatic extension of a rubber band. This underlies\nthe standard classroom demonstration of the entropic origin of rubber elasticity, whereby a rubber\nband is rapidly extended and placed in contact with a (highly temperature-sensitive) upper lip. This\nkind of experiment goes back at least as far as Gough [5] and Joule [6], and some of Joule\u2019s data\nare shown in Figure\n10.9 along with some from James and Guth [7]. At low extensions,\nthe temperature actually decreases slightly, but then increases steadily. The interpretation of the\nexperiment is as follows. In the adiabatic extension of an ideal elastomer, the work done on\nthe sample is retained entirely as heat; there is a loss of entropy but no change of internal energy\nand dc]: ~dw. The work is given by Equation 10.4.1 and the heat by Equation 10.4.4; therefore\nthe temperature change is\n"]], ["block_4", ["We now proceed to use a molecular model to derive predictions for the stress\u2014strain behavior of an\nideal elastomer. In the subsequent section, we will consider various nonidealities that could occur\nin a real material, but even granted the existence of some or all of these nonidealities, the\n"]], ["block_5", ["length. These data show an interesting feature, known as thermoelastic inversion, whereby at\nelongations below about 10%, the stress decreases with temperature, in contrast to the larger strain\nbehavior. As we are anticipating that the elasticity is primarily due to entropy, we expect the force\nto increase with temperature. The reason for the behavior at small elongation is actually quite\nsimple; it is due to thermal expansion. The unstrained length increases with temperature due to\nexpansion and thus the actual strain at fixed length decreases with increased temperature and\nconsequently the force decreases. Thus the thermoelastic inversion can be eliminated by comparing\nthe data at constant strain.\nThis kind of thermoelastic data can be further analyzed in terms of the thermodynamic\ncontributions. From Equation 10.4.19 we can write\n"]], ["block_6", ["where CL is the appropriate heat capacity at constant length. As in the previous examples, the\nnegative change in temperature at small extensions is due to the positive entrOpy of deformation,\nthat is, it corresponds to the thermoelastic inversion.\n"]], ["block_7", ["398\nNetworks, Gels. and Rubber Elasticity\n"]], ["block_8", ["and\n"]], ["block_9", ["10.5\nStatistical Mechanical Theory of Rubber Elasticity: Ideal Case\n"]], ["block_10", [{"image_3": "409_3.png", "coords": [41, 503, 203, 549], "fig_type": "molecule"}]], ["block_11", ["L\n1\nT\nas\nAT 2 a: dL Fir\u2014J (a) dL\n(10.4.22)\n"]], ["block_12", ["3\n8f _\nas\n(0\u20142\");\n_\n(0\u20141.),\n(10.4.20)\n"]], ["block_13", ["6U\n.2\n6f\n(67);\n4671\n(10.4.21)\n"]]], "page_410": [["block_0", [{"image_0": "410_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "410_1.png", "coords": [32, 299, 229, 598], "fig_type": "figure"}]], ["block_2", [{"image_2": "410_2.png", "coords": [32, 38, 225, 319], "fig_type": "figure"}]], ["block_3", ["Figure 10.8\nStress versus elongation for natural rubber, resolved into internal energy and entropic contri\u2014\nbutions, at (a) constant temperature and (b) constant strain. (Data from Anthony, R.L., Caston, RH, and\nGuth, E., J. Phys. Chem, 46, 826, 1942. With permission.)\n"]], ["block_4", ["0?\n_2\nI\nl\nI\n___J\n100\n200\n300\n400\n(a)\nElongation (%J\n"]], ["block_5", ["_2\n\\\\\\\\\\\\J\nI\nI\nI\n0\n100\n200\n300\n400\n(b)\nElongation (%)\n"]], ["block_6", ["Statistical of Rubber Elasticity: Ideal Case\n399\n"]], ["block_7", ["N\nE\nO\n\"6)\nx\n8 \n"]], ["block_8", ["(T12 \n"]], ["block_9", ["U)\n0')\n93\nX\n55\n3 \n"]], ["block_10", ["E\n-Q\n3\u2019\n\u2014T(aS/aI)T\n"]], ["block_11", ["20-\n"]], ["block_12", ["16-\n"]], ["block_13", ["12 \n"]], ["block_14", ["4 _\n"]], ["block_15", ["O\n"]], ["block_16", ["20\n"]], ["block_17", ["16-\n"]], ["block_18", ["4 \n"]], ["block_19", ["+\n"]], ["block_20", ["6f\nH?\n"]], ["block_21", ["f\n"]], ["block_22", ["(aU/aI)T\n"]], ["block_23", ["f\n6f\n71?);\n"]], ["block_24", [")1\n"]]], "page_411": [["block_0", [{"image_0": "411_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Since entr0py plays the determining role in the elasticity of an ideal elastomer, let us review some\nideas about this important thermodynamic variable. We used a probabilistic interpretation of\nentr0py extensively in Chapter 7 to formulate the entropy of mixing. The starting point was the\nBoltzmann relation:\n"]], ["block_2", ["where k is Boltzmann\u2019s constant and 0. is the number of possible states. As then, the difference in\nentropy between two states of different thermodynamic probability is\n"]], ["block_3", ["so that AS is positive when 02 > 01 and negative when 02 < 0.1.\n"]], ["block_4", ["qualitative success of the ideal model is really a remarkable triumph of statistical mechanics. We\nhave already considered the most famous equation of state, that of the ideal gas. That simple result\nis illuminating, but only describes the behavior of very dilute gases with any reliability and dilute\ngases are of limited significance from the point of view of materials science. In contrast, the ideal\nelastomer equations will provide a reasonable description of a practically important, but extremely\ncomplex, amorphous condensed phase, even though the derivation is not appreciably more\nelaborate than that for the ideal gas. We will begin by considering the force required to extend a\nsingle Gaussian chain, an example that already arose in the context of chain swelling in Section 7.7\nand that will resurface in the bead\u2014spring model of viscoelasticity in Section 11.4. Then we will\napply this result to an entire ensemble of cross-linked chains.\n"]], ["block_5", ["Figure 10.9\nTemperature change during adiabatic extension of natural rubber. (Data from Joule, J.P., Phil.\nTrans. R. 506., 149, 91, 1859; James HM. and Guth, E., J. Chem Phys, 11, 455, 1943; 15, 669, 1947.) (From\nTreloar, L.R.G., The Physics ofRubber Elasticity, 3rd ed., Clarendon Press, Oxford, 1975. With permission.)\n"]], ["block_6", ["10.5.1\nForce to Extend a Gaussian Chain\n"]], ["block_7", ["400\nNetworks, Gels. and Rubber Elasticity\n"]], ["block_8", ["S\n"]], ["block_9", ["_E'_,,\u2019 0.12\n5\nS\n"]], ["block_10", ["CD\nI\u2014\n"]], ["block_11", ["\u201c\u201832\n8\nE \n"]], ["block_12", [{"image_1": "411_1.png", "coords": [39, 48, 280, 302], "fig_type": "figure"}]], ["block_13", ["0.00\n"]], ["block_14", ["0.04\n"]], ["block_15", ["S=kln0\n(10.5.1)\n"]], ["block_16", ["0.20 \n"]], ["block_17", ["0 James and Guth (1943)\n0.16\n"]], ["block_18", ["as s; s, kln (93)\n(10.5.2)\n01\n"]], ["block_19", ["[\nl\nr\nl_\nJ\n0\n20\n40\n60\n80\n100\nExtension (%)\n"]], ["block_20", ["_\nA Joule (1859)\n"]]], "page_412": [["block_0", [{"image_0": "412_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "412_1.png", "coords": [31, 432, 242, 670], "fig_type": "figure"}]], ["block_2", [{"image_2": "412_2.png", "coords": [34, 178, 231, 242], "fig_type": "figure"}]], ["block_3", ["\u20186' \u20199\nand the subscriptr\non P denotes the \u201cinitial\u201d state. We then extend the chain to a new end-to\u2014\nend distance, 11, with coordinates between (x, y, z) and (x + dx, y + dy, z + dz). The corre3ponding\n\u201cfinal\u201d state distribution function Pf is\n"]], ["block_4", ["Figure 10.10\nExtension of a single Gaussian chain from initial end\u2014to-end distance ho to final \ndistance 11.\n"]], ["block_5", ["We now associate the number of possible conformations with the entropy defined by Equation\n10.5.1, that is, we take .QzAP, with A as some unspecified proportionality constant. Then we\n"]], ["block_6", ["can say\n"]], ["block_7", ["and (x0+dx0, y0+ 20+ dzo), as shown in Figure 10.10. The imposed end\u2014to-end distance\nis ho: (x0\u2014I- yo + 20)] (2, which may be compared to the equilibrium mean square end-toend\ndistance (112): N192. The number of ways that this chain can satisfy the imposed constraint is given\nby the Gaussian distribution (recall Equation 6.7.1):\n"]], ["block_8", ["In the previous section, we identified the force of extension with the associated change in free\nenergy (Equation 10.4.11 or Equation 10.4.19). Then, if the change in free energy is entirely due to\nthe entropy, the material is an ideal elastomer. Figure 10.8 provides an example of how reasonable\nthis assumption is for a material; now we apply it to one chain. Consider extending a single\nGaussian chain of N units, with statistical segment length b (recall Section 6.3). The chain has one\nend fixed at the (0,0,0) and the other is held in the infinitesimal cube between (x0, yo, 20)\n"]], ["block_9", ["where we de\ufb01ne the normalization factor, 6 as\n"]], ["block_10", ["Statistical Mechanical Theory of Rubber Elasticity: Ideal Case\n401\n"]], ["block_11", [{"image_3": "412_3.png", "coords": [37, 559, 269, 638], "fig_type": "figure"}]], ["block_12", [{"image_4": "412_4.png", "coords": [39, 461, 222, 545], "fig_type": "molecule"}]], ["block_13", [{"image_5": "412_5.png", "coords": [40, 391, 224, 450], "fig_type": "molecule"}]], ["block_14", [{"image_6": "412_6.png", "coords": [41, 179, 221, 236], "fig_type": "molecule"}]], ["block_15", [{"image_7": "412_7.png", "coords": [43, 452, 215, 589], "fig_type": "figure"}]], ["block_16", ["Pfav, Z) [33/2 exp [43112]\n(10.55)\n"]], ["block_17", ["B E 277012) 271's\n(10.5.4)\n"]], ["block_18", ["i\nP\nAscham :klnAPf\u2014 klnAP\u2014kln (Pf)\n: \u2014kw[3(h2 \u2014 213)\n(10.5.6)\n"]], ["block_19", ["(0,0,0)\n(x, y, 2)\n"]], ["block_20", ["(0 o 0)\nah\\>\n(X0 Yaszo)\n"]], ["block_21", [{"image_8": "412_8.png", "coords": [53, 563, 259, 638], "fig_type": "molecule"}]], ["block_22", [".523}UF<\n"]], ["block_23", ["3\n_\n3\n"]], ["block_24", ["X\n"]], ["block_25", ["3\n3/2\n3\u201913\n<>war\u20141\n= B3/ZeXP[-w{3h8]\n(10.5.3)\n"]]], "page_413": [["block_0", [{"image_0": "413_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "413_1.png", "coords": [27, 554, 120, 638], "fig_type": "figure"}]], ["block_2", [{"image_2": "413_2.png", "coords": [32, 91, 174, 136], "fig_type": "molecule"}]], ["block_3", ["end\u2014to\u2014end vector of each strand is deformed so that the coordinates of one end transform\nx0 \u2014>x=)txx0, yo \u2014> y :Ayyo, 20 \u2014:- z A220, when we take the other end as the origin. We already\nknow the entropy change per strand associated with this process: it is simply the result for a single\nchain, see Equation 10.5.6, applied to a single strand. Writing it out in more detail, we have\n"]], ["block_4", ["This is a fundamental result, and one we will use extensively in modeling the viscoelastic\nprOperties of polymer liquids in Chapter 11. Equation 10.5.7 indicates that a single Gaussian\nchain behaves like a Hooke\u2019s law spring, with force constant 3kT/(112) and zero rest-length. Note\nthe interesting result that this spring will stiffen as T increases, in contrast to intuitive expectation\nfor a metal spring; this is a direct result of its entropic basis. Equation 10.5.7 contains most of the\nphysical concepts that are required to describe rubber elasticity from a molecular viewpoint.\n"]], ["block_5", ["402\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_6", ["We now consider an ideal network made up of Gaussian strands. If the cross-links were introduced to\na melt of Gaussian chains, for example, by vulcanization, it is plausible that the strands will be more\nor less Gaussian as well. For simplicity, we will assume that all strands contain an identical number\nof statistical segment lengths, NI; this simpli\ufb01cation will subsequently be removed. We now impose\n"]], ["block_7", ["where we use the subscript \u201cchain\u201d to emphasize that this is a single chain calculation. The unknown\nconstant A cancels out when we calculate the change in entropy. The force to extend the chain to h is\ngiven by\n"]], ["block_8", ["This is a reasonable approximation for bulk elastomers, where Poisson\u2019s ratio is nearly 0.5, but is\nnot appropriate, for example, when the network is swollen with solvent. The removal of this\nassumption will be discussed in Section 10.7.\nWe now make a final, very important assumption, the so\u2014called a\ufb02ine junction assumption:\neach junction point moves in proportion to the macroscopic deformation. Consequently, the\n"]], ["block_9", ["a macroscopic deformation on the network; for example, we might stretch it in the x direction.\nHowever, to be more general, we describe the deformation by three extension ratios Ax, )1y, and A,,\ngiven by [ox/LO, Ly/LO, and Lz/LO, respectively. If we begin with a cube of material of length L0 on each\nside, that cube will be deformed to a three-dimensional volume element with sides LI, Ly, and L2, as\nshown in Figure 10.11. We assume that there is no volume change on deformation, and thus\n"]], ["block_10", ["Figure 10.11\nDeformation of a cube of material subjected to uniaxial elongation along x.\n"]], ["block_11", ["10.5.2\nNetwork of Gaussian Strands\n"]], ["block_12", [{"image_3": "413_3.png", "coords": [40, 557, 331, 635], "fig_type": "figure"}]], ["block_13", ["_\naASchain\n_E\nf__ (T) s\n(\u201912)):\n(10.5.7)\n"]], ["block_14", ["P\nASslrand \n(Ff)\n: 477508 + y2 + 22) _ (4773018 + yci + 28))\n"]], ["block_15", ["V LxLyLz V0 L8;\nAxAyAz 1\n(10\"58)\n"]], ["block_16", [{"image_4": "413_4.png", "coords": [61, 482, 289, 554], "fig_type": "molecule"}]], ["block_17", ["L0\nLX\nL\nL0\n"]], ["block_18", ["wk 2N b2 (n1} 1) + )3013 1) + 23,013 1))\n(10.5.9)\n"]], ["block_19", ["A\n,1\n"]]], "page_414": [["block_0", [{"image_0": "414_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "414_1.png", "coords": [30, 190, 182, 225], "fig_type": "molecule"}]], ["block_2", ["This equation represents the principal result of this molecular network theory. We will now\nconsider a specific deformation to obtain expressions for the modulus, but the necessary manipu-\nlations are all results of continuum elasticity theory and require no further assumptions about what\nthe molecules are doing.\n"]], ["block_3", ["We begin with a uniaxial extension, say along x, by a stretch ratio )1. Thus Ax =21, and by volume\nconservation (see Equation 10.5.8) A), )1, 1/ x/X. Furthermore, 8 A 1. In this case, then\n"]], ["block_4", ["The stress given by Equation 10.5.14 is sometimes called the true stress to distinguish it from\nthe quantity given Equation 10.5.15, which is known as the engineering stress or the nominal\n"]], ["block_5", ["(121 d8):\n"]], ["block_6", ["Alternatively, it is often experimentally more convenient to divide by the initial cross-sectional\n"]], ["block_7", ["Note that the same result is obtained if we use either the true stress or the engineering stress\nbecause they coincide in the small strain limit.\n"]], ["block_8", ["In this rather simple N,does not appear, so the assumption of constant N,C was actually\nnot necessary. To obtain the total entrOpy change for the material, we simply need the number\nof strands per unit volume. For our ideal network this is given by pNav M,, where M, is the\n(number average) molecular weight between cross-links, but in anticipation of defects such as\ndangling ends and loops in real networks, we will just define the total number of elastically\ne\ufb01fective strands, ve. The number of strands per unit volume is thus ve/V and the total entropy\nchange becomes\nk\nas: \u2014%(A\u00a7+A\u00a7+A\u00a7~3)\n(10.5.11)\n"]], ["block_9", ["and the force is given by\n"]], ["block_10", ["Note that the force changes sign, as it should, when )1 1. If we now divide both sides by the cross\u2014\nsection area normal to the stretching direction, LyLz lag/A, we obtain the tensile stress:\n"]], ["block_11", ["area, Lg, which leads to the following result:\n"]], ["block_12", ["stress.\nWe can now obtain an expression for Young\u2019s modulus, E, recalling Equation 10.3.1 (and that\n"]], ["block_13", ["We now note that on average x8 Nxb2/3, and the same for yo and 20, so that\n"]], ["block_14", ["10.5.3\nModulus of the Gaussian Network\n"]], ["block_15", ["Statistical Mechanical Theory of Rubber Elasticity: Ideal Case\n403\n"]], ["block_16", [{"image_2": "414_2.png", "coords": [37, 73, 205, 101], "fig_type": "molecule"}]], ["block_17", [{"image_3": "414_3.png", "coords": [37, 379, 259, 414], "fig_type": "molecule"}]], ["block_18", ["6A5\nT\n6A5\nelcT\n1\nfzm\nzit.\n_ 2\u201d\n,1__2\n(10.5.13)\n3L\nL0\n8A\nL0\n)1\n"]], ["block_19", [",\nact\nV6\nE :l\n\u2014\u2014\u2014\u2014 =\nkT\u2014\n10.5.16\nAl\u2014I\u2014Ti 6A\n3\nV\n(\n)\n"]], ["block_20", ["vk\n2\nas: \u2014i\n212\n\u2014-\u2014-3\n105.12\n2 \n+)t\n)\n(\n)\n"]], ["block_21", ["1\na, kTFVE (A P)\n(105.15)\n"]], ["block_22", ["at \n(12 -9\n(105.14)\n\u201c\u2014\narea\nL5\n._\nV\n"]], ["block_23", ["k\n(Mama) 5(213 + + \n(105.10)\n"]], ["block_24", [{"image_4": "414_4.png", "coords": [50, 449, 195, 484], "fig_type": "molecule"}]], ["block_25", [{"image_5": "414_5.png", "coords": [56, 331, 158, 366], "fig_type": "molecule"}]]], "page_415": [["block_0", [{"image_0": "415_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "415_1.png", "coords": [31, 400, 193, 605], "fig_type": "figure"}]], ["block_2", ["where in the last step we have substituted the ideal value for ve/V in terms of the molecular weight\nbetween cross-links, Mx. From these equations (Equation 10.5.14 through Equation 10.5.17) we\ncan extract some important conclusions:\n"]], ["block_3", ["1.\nThe modulus increases with temperature, just as with the spring constant of a single chain, due\nto its entropic origin.\n2.\nThe modulus increases as a function of cross-link density, because M, decreases; a \u201ctighter\u201d\nnetwork is \u201cstiffer.\u201d\nThe modulus is independent of the functionality of the cross-links.\n4.\nThe extensional stress is not a linear function of the strain, even though the individual network\nstrands are supposed to be Hookean. (In contrast, the shear stress turns out to be linear in the\nstrain, but we will not take the time to derive this relation.)\n5.\nAssuming a density of 1 g/cm3 at room temperature, and M, = 10,000 g/mol, Equation 10.5.17\ngives a modulus of 2.5 x 106 dyn/cmz, or 0.25 MPa. Typical values for elastomers fall within\nan order of magnitude of this number.\n"]], ["block_4", ["404\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_5", ["Figure 10.12\nStress for cross-linked natural rubber in compression and extension. (Data from Treloar,\nL.R.G., Trans. Faraday 506., 40, 59, 1944. With permission.)\n"]], ["block_6", ["We finally obtain an expression for the shear modulus, G, using the approximate relation G E/3\n(Equation 10.3.4a):\n"]], ["block_7", ["U.)\n"]], ["block_8", [{"image_2": "415_2.png", "coords": [35, 309, 310, 610], "fig_type": "figure"}]], ["block_9", [{"image_3": "415_3.png", "coords": [39, 300, 91, 625], "fig_type": "figure"}]], ["block_10", ["(N/mm2)\n"]], ["block_11", ["force\n"]], ["block_12", ["Tensile\n"]], ["block_13", ["or\n"]], ["block_14", ["compressive\n"]], ["block_15", [{"image_4": "415_4.png", "coords": [43, 81, 131, 115], "fig_type": "molecule"}]], ["block_16", ["p RT\nG \u201d$3\n___ M\n(10.5.17)\n"]], ["block_17", ["00\n|\n|\nl\n|\nl\n|\nI\nI\n0.4\n0.6\n0.8\n1.0\n1.2\n1.4\n1.6\n1.8\n2.0\n,1\n| .0.p. I\n"]], ["block_18", ["| .0a: I\n"]], ["block_19", ["L\n"]], ["block_20", ["_3.2 L\n"]], ["block_21", ["__ \n0.4\n\u2014\n_..--0\"\"'\n"]], ["block_22", ["0.8\n\u2014\n"]], ["block_23", ["_'t\n"]], ["block_24", ["b)\n"]], ["block_25", ["to\n"]], ["block_26", ["l\n"]], ["block_27", ["|\n"]], ["block_28", ["Compression\nExtension\n"]]], "page_416": [["block_0", [{"image_0": "416_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "416_1.png", "coords": [30, 199, 84, 601], "fig_type": "figure"}]], ["block_2", ["An example of a test of the theory, and Equation 10.5.14 in particular, is shown in Figure 10.12.\nBoth extensional and compressive stresses were determined as a function of )1 for a piece of\nvulcanized The agreement between experiment and theory is impressive, particularly in\ncompression. The same sample was subsequently extended up to its breaking point, near )1 $7.5,\nand the results are shown in Figure 10.13. The data at low extension ratios were \ufb01t to the theory to\nobtain the MPa. theory and the data are not in perfect agreement \n"]], ["block_3", ["an appreciable fraction of the contour length, the Gaussian distribution no longer applies. This\npoint will be considered again in the next section.\n"]], ["block_4", ["the failure offor large extensions; when the end-to-end distance \n"]], ["block_5", ["the main difference in experimental stress at high )1. This is primarily \n"]], ["block_6", ["Statistical Mechanical Theory of Rubber Elasticity: Ideal Case\n405\n"]], ["block_7", ["extension ratios (A). (Data from Treloar, L.R.G., Trans. Faraday Soc, 40, 59, 1944. With permission.)\n"]], ["block_8", ["Figure 10.13\nSame sample as in Figure 10.12, but now subjected to simple extension and much larger\n"]], ["block_9", ["(a))\n"]], ["block_10", ["for\n"]], ["block_11", ["(scale\n"]], ["block_12", ["(N/mmz)\n"]], ["block_13", ["Tensile\n"]], ["block_14", ["force\n"]], ["block_15", ["curve\n"]], ["block_16", ["area\n"]], ["block_17", ["unstrained\n"]], ["block_18", ["unit\n"]], ["block_19", ["per\n"]], ["block_20", ["7.0\n-\n"]], ["block_21", ["6.0\n-\n"]], ["block_22", ["5.0\n-\n"]], ["block_23", ["3.0 \n"]], ["block_24", [{"image_2": "416_2.png", "coords": [51, 209, 253, 630], "fig_type": "figure"}]], ["block_25", ["Extension ratio\n"]], ["block_26", ["Theoretical\n"]]], "page_417": [["block_0", [{"image_0": "417_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "417_1.png", "coords": [31, 518, 246, 565], "fig_type": "molecule"}]], ["block_2", ["The approximate validity of the first two assumptions was suggested in the previous section and\nthe effects of deviations from these assumptions should be reasonably transparent. Accordingly,\nexcept for the issue of solvent swelling, which will be taken up in the following section, we will not\nconsider these further. The third assumption is more interesting from a molecular point of view.\nOne violation of this assumption can be readily imagined: upon large extensions, say beyond 100%\n(recall Figure 10.13), a strand may no longer be Gaussian. Clearly as we approach the limit of full\nextension, when the end-to-end distance becomes a significant fraction of the contour length, the\nGaussian force law (Equation 10.5.7) will not apply. This problem has been addressed theoretically\nand a reasonable solution is known. Specifically, Kuhn and Griin [8] showed that when a freely\njointed chain of N, links of length b is extended to an end-to-end distance 11, the distribution\nfunction is not the familiar Gaussian, but rather\n"]], ["block_3", ["where the quantity B is the so-called inverse Langevin function, L _1(x):\n"]], ["block_4", ["This formulation is not particularly transparent, but it turns out that the exponential in Equation\n10.6.2 can be expanded as a power series in (h/Nxb) as follows:\n"]], ["block_5", ["It turns out that refinement or relaxation of almost any of these assumptions has generated\nsignificant amounts of controversy over the years, and to address these issues thoroughly would\nrequire an entire book. Accordingly, we will have to be content with a few examples.\n"]], ["block_6", ["406\nNetworks, Gels. and Rubber Elasticity\n"]], ["block_7", ["3.\nThe number of conformations available to the strands both before and after deformation is\ngiven by a Gaussian distribution.\n4.\nThe number of conformations available to the strands before deformation is the same as for\nequivalent chains in the uncross\u2014linked state.\n5.\nThe junction points deform affinely with the macroscopic deformation.\n6.\nThe number of elastically effective strands per unit volume, ve/V, is given by pNav/Mx for a\nperfect network.\n"]], ["block_8", ["The deve10pment of a statistical thermodynamic approach to rubber elasticity in the previous\nsection involved a series of assumptions that could be questioned. In this section we touch brie\ufb02y\non some of these and give an indication of how they might be addressed. To begin with, we recall\nthe central result of the theory (see Equation 10.5.11):\n"]], ["block_9", ["from which the stress and modulus can be computed for any deformation. The main assumptions\ninvoked in the development and application of this equation are summarized below:\n"]], ["block_10", ["10.6.1\nNon-Gaussian Force Law\n"]], ["block_11", ["10.6\nFurther Developments in Rubber Elasticity\n"]], ["block_12", ["1.\nThere is no change in internal energy, U, upon deformation at constant T and p.\n"]], ["block_13", [{"image_2": "417_2.png", "coords": [37, 627, 332, 678], "fig_type": "molecule"}]], ["block_14", [".\nThere is no change in volume, V, upon deformation at constant T and p.\n"]], ["block_15", [{"image_3": "417_3.png", "coords": [43, 130, 186, 161], "fig_type": "molecule"}]], ["block_16", ["B L_1(h/Nxb)\n(10.6.3)\n"]], ["block_17", ["h\nB\nP Nx, 12\n\u2014\u2014NJ.\n1\n,\n10.6.2\n(\n)\u201cexp(\n[Wm \u201c(smh\ufb01m\n(\n)\n"]], ["block_18", ["3\nh\n2\n9\nh\n4\n99\nh\n6\n\u2014 \u2014\n\u2014 \u2014\n\u2014\n10.6.4\n2 (Mb) +20 (Mb) +350 (N, >\nD\n(\n)\nP(Nx, h) oc exp (\u2014N,C\n"]], ["block_19", ["AS=\u2014%\u2019E()t\u00a7+)tj+)tf\u20143)\n(10.6.1)\n"]], ["block_20", ["'\n"]]], "page_418": [["block_0", [{"image_0": "418_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "418_1.png", "coords": [26, 35, 293, 315], "fig_type": "figure"}]], ["block_2", ["The fourth assumption above represents a reasonable simplification, but it need not hold, even for\nrubbers lightly cross-linked in the melt state. Under some circumstances, such as cross-links\nintroduced while the material is under stress, it certainly would not apply. The first step in lifting\nthis assumption is to assume that in the undeformed state the elastomer is isotropic and the mean\n"]], ["block_3", ["from which we can see that the Gaussian result is just the first term in a series. When the argument\n(h/Nxb) is small, the Gaussian function is adequate, but at large extensions Equation 10.6.4 is more\naccurate. This result can be carried through the analysis of the previous section (see Equation\n10.5.6 and Equation 10.5.7) to obtain the corresponding force law and the result is illustrated in\nFigure 10.14. The main new feature is that for extensions such that (h/Nxb) > 0.4, the force\nincreases rather sharply. This is in excellent qualitative agreement with the data in Figure 10.13.\nIn fact, recent experiments have been able to measure the force of extension of single DNA\nmolecules and the inverse Langevin function provides a good account of the results (see Problem\n12), so although this approach is mathematically unwieldy, it is successful.\nA second difficulty with the Gaussian assumption arises from the inevitable distribution of\nstrand lengths. In the development leading up to Equation 10.5.10, we argued that because the\nquantity N, cancels out of the \ufb01nal expression for AS, the assumption of monodisperse strands was\nbenign. In reality the cross-linking process, however it is carried out, will leave some distribution\nof Nx. Consequently, the non-Gaussian character of the distribution function will become apparent\n"]], ["block_4", ["11/!) = \\/N_x, which is then stretched to Am. The smaller the Nx, the smaller the value of A\nrequired for Am to approach Nx. This problem is much more difficult to deal with, not least\nbecause of the difficulty in characterizing the distribution of Nx.\n"]], ["block_5", ["Figure 10.14\nForce versus extension for a Gaussian chain (first-term approximation) and for the full\ninverse Langevin function.\n"]], ["block_6", ["at different values of the macroscopic extension; shorter strands will sooner become more fully\nstretched than longer ones. For example,\nan average strand\nbegins with\nan unperturbed\n"]], ["block_7", ["10.6.2\nFront Factor\n"]], ["block_8", ["Further in Rubber Elasticity\n407\n"]], ["block_9", ["8\nInverse Langevin function\n8u.\n"]], ["block_10", ["20\n"]], ["block_11", ["16 L\n"]], ["block_12", ["12 \n"]], ["block_13", ["4 ._\n"]], ["block_14", ["8 _\n"]], ["block_15", ["1\nl_\nl\n_|\n0\n0 2\n0.4\n0.6\n0 8\n1 0\nh/Nxb\n"]], ["block_16", ["\u2014 --- First-term approximation\n"]], ["block_17", ["(Gaussian)\n"]]], "page_419": [["block_0", [{"image_0": "419_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Any real network will contain defects in the structure, as suggested at the beginning of the chapter.\nThese include loops, dangling ends, and the sol fraction. We have finessed this issue, in part, by\nusing the concept of elastically effective strands, 123, which suggests that the contribution from such\ndefects has been removed. However, it is not a straightforward matter to account for these in\npractice. Consider two general approaches. By NMR spectroscopy, for example, we might monitor\nthe conversion of double bonds in a polydiene. From this information we could estimate M, and\nthus 12,, by Assumption 6. But the NMR experiment could not tell us about loops or distinguish sol\nfraction from gel fraction (and, in fact, for modest degrees of cross-linking it is hard to even\nquantify that reliably). The next option is to measure the modulus itself and fit the results to the\nmodel. But, because all of these various effects contribute proportionally to the modulus, there is\nno easy way to resolve them from measurements on a single sample. In general, these three\nparticular kinds of defects are treated in the following way: (a) loops are ignored, (b) dangling ends\nare corrected for, and (c) the conversion is high enough that the sol fraction is negligible, or it is\nextracted before measurement. Under (b), the number of dangling ends can be estimated from M,,\nbecause each prenetwork chain had two ends. Therefore each prenetwork chain will contribute two\ndangling ends to the network, and the average length of these dangling ends will be M,.\nAccordingly, the fraction of these dangling ends will be 2Mx/M, where M is the molecular weight\nof the prenetwork chain and the number of effective strands may be estimated as\n"]], ["block_2", ["Another contribution to the modulus, the so-called trapped entanglement, although not strictly a\ntopological defect has been considered extensively. We will see in Section 11.6 that the visco\u2014\nelastic properties of high-molecular-weight polymers are dominated by the phenomenon of\n"]], ["block_3", ["The ratio of mean square end-to-end distances in the undeformed network relative to that in a melt\nof the same strands is known as the front factor. As it is a constant, it carries through the\nsubsequent development and therefore appears in the expressions for the stress and the modulus.\nAlternatives to Assumption 5\u2014that the junction points deform affinely with the macroscopic\nstrain\u2014have also been pr0posed. In the most common approach only the junctions on the edges of\nthe material are so constrained; those in the bulk of the network are free to \ufb02uctuate about their\nmean positions [7]. This model is sometimes referred to as the phantom network. This modification\ncannot be applied to the entropy of a single strand, as in Equation 10.6.5, but must be applied to the\nnetwork as a whole. The result is a reduction in the net stress, which is determined by the number\nof junctions, 12,; Equation 10.5.11 so modified becomes\n"]], ["block_4", ["408\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_5", ["Again, this new front factor is a constant and would carry through to the expressions for the stress\nand the modulus. For the particular case of a regular network in which f strands emerge from\neach junction, then 12,, f/Z, and (12e 12,) ve(f 2)/f. For a network made by vulcanization,\ntherefore, f=4 and this term is equal to 126/2.\n"]], ["block_6", [",2, and the expression for the strand entropy\nsquare strand length is (hi2), where the subscript\nwe would replace x3 =3% =28 with xi2 y? 2\n(Equation 10.5.10) would be replaced by\n"]], ["block_7", ["10.6.3\nNetwork Defects\n"]], ["block_8", [{"image_1": "419_1.png", "coords": [38, 263, 238, 288], "fig_type": "molecule"}]], ["block_9", [{"image_2": "419_2.png", "coords": [45, 593, 142, 633], "fig_type": "molecule"}]], ["block_10", ["Ve\npNav\n2Mx\n\u2014=\n\u2014\n.6.7\nV\nM. \nM \n(.0\n)\n"]], ["block_11", ["2\nas \u2014(u, In); \u00e9\u2014Z-gg (A3 +213, + 213- 3)\n(10.6.6)\n"]], ["block_12", ["k\nh-2\n(Asstrand>\n= \n5 \n(10.65)\n< 3)\n"]], ["block_13", [{"image_3": "419_3.png", "coords": [56, 99, 219, 132], "fig_type": "molecule"}]], ["block_14", ["(6'?!\n1\ndenotes initial. If we return to Equation 10.5.9,\n"]]], "page_420": [["block_0", [{"image_0": "420_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "420_1.png", "coords": [33, 467, 196, 503], "fig_type": "molecule"}]], ["block_2", ["where C1 and C2 are unknown parameters of the material, but not functions of the deformation.\nThe first term in this Mooney\u2014Rivlin equation, therefore, has exactly the form of the statistical\ntheory, and the second term can be viewed as a correction. The second form of Equation 10.6.11\nsuggests plotting the following quantity versus 1//\\:\n"]], ["block_3", ["The preceding discussion of modifications to the ideal elastomer theory has emphasized additional\nmolecular contributions and particularly ones that modify the prefactor. Following Equation\n10.5.15, we may represent the tensile stress from these models as\n"]], ["block_4", ["entanglement, results from the intertwining of different chains. These entanglements act like\ntemporary cross-links, imparting a rubber-like modulus to the liquid at intermediate times, before\neventual relaxation \ufb02ow. By exploiting the expression for the shear modulus of an ideal\nelastomer, Equation 10.5.17,a molecular weight between entanglements can be defined, Me. Now\nimagine we have a high-molecular-weight melt, with M >> Me, and we cross-link it enough to\nproduce a network. (Recall from Equation 10.1.1 that we only need a few cross\u2014links, eel/N, to\npass the gel point.) At this stage the modulus must be about the same, or perhaps slightly higher\nthan it before but Mx >> Me. In other words, we expect a modulus of about pRT/Me,\nbut the theory of rubber elasticity says it should be only pRT/Mx. Therefore these entanglements\ncontribute but they are trapped; full stress relaxation, or \ufb02ow, is eliminated by\nthe cross-linking. Thus at low degrees of cross-linking, trapped entanglements should make\nthe experimental modulus larger than expected by ideal elastomer theory. As the degree of\ncross-linking goes up, so that M, Gnetwork \u2018i\u2014 XGentangIements pRT (\u20181\u2014 + Xi)\n(106-8)\nMA?\nMB\n"]], ["block_17", [{"image_5": "420_5.png", "coords": [48, 531, 330, 575], "fig_type": "molecule"}]], ["block_18", [{"image_6": "420_6.png", "coords": [130, 536, 317, 567], "fig_type": "molecule"}]], ["block_19", [")1\n"]]], "page_421": [["block_0", [{"image_0": "421_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 10.15\nA \u201cMooney plot\u201d for various rubbers in simple extension. (Data from Gumbrell, S.M.,\nMullins, L., and Rivlin, R.S., Trans. Faraday 506., 49, 1495, 1953. With permission.)\n"]], ["block_2", ["which should give a straight line with intercept 2C 1 and slope 2C2. Adherence to the ideal theory\nshould give a horizontal line. Examples from a variety of rubbers are shown in Figure 10.15;\nclearly Equation 10.6.12 gives a good description of the data, with nonzero values of C2. There are\nother cases, however, such as swollen rubbers or networks prepared by cross\u2014linking in solution,\nwhere C2 :50. A convincing and generally applicable molecular explanation for C2 has proven\nelusive. Furthermore, for some materials the values of C2 are even found to depend on the kind of\nexperimental deformation employed, which suggests that Equation 10.6.11 is not the universally\ncorrect functional form.\n"]], ["block_3", ["One of the distinguishing features of a lightly cross-linked polymer material is its ability to imbibe\nand retain a large volume of solvent. Examples include:\n"]], ["block_4", ["410\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_5", ["10.7\nSwelling of Gels\n"]], ["block_6", ["1.\nHot melt adhesives. In this family of adhesives, the glue is applied as a liquid at high\ntemperature and solidifies upon cooling. A typical formulation could include about 30% of\na \u201cthermoplastic elastomer,\u201d a styrene\u2014isoprene\u2014styrene triblock copolymer. At low temper-\natures, the styrene segments segregate from the isoprene blocks to form roughly spherical\nstyrene aggregates that act as cross-links. At high temperatures, however, the segregation\nis disrupted and the polymer \ufb02ows. The remaining 70% of the material consists of low-\nmolecular-weight species, largely to dilute the isoprene segments and make the resulting gel\n"]], ["block_7", ["(N/mmz)\n"]], ["block_8", ["\u2014\n1M2)\n"]], ["block_9", ["2(2.\n"]], ["block_10", [{"image_1": "421_1.png", "coords": [59, 134, 278, 357], "fig_type": "figure"}]], ["block_11", ["0.2L\n"]], ["block_12", ["0.1\nI\nI\n1\nI\n1\n0.5\n0.6\n0.7\n0.8\n0.9\n1.0\n1H\n"]], ["block_13", ["0.3 \n"]], ["block_14", [{"image_2": "421_2.png", "coords": [68, 178, 270, 284], "fig_type": "figure"}]]], "page_422": [["block_0", [{"image_0": "422_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nDiapers. Clearly, the major function of a diaper material is to imbibe and retain a large\nquantity of aqueous solution, without dissolving. Diaper materials can absorb up to several\nhundred times their own weight. A typical ingredient is poly(acrylic acid) or (sodium poly-\nacrylate). In this instance the network strands bear charged groups, which have the effect of\ngreatly increasing the osmotic drive for water to enter the polymer and swell it.\n4.\nBiological tissue. Much of biological tissue is essentially network material, although of course\nvery complex. For example, in the \u201cextra cellular matrix\u201d collagen molecules intertwine in\ntriple helices, which in turn aggregate to form fibrils, which in turn cross\u2014link with the\nassistance of certain proteins to form three-dimensional gels.\n"]], ["block_2", ["We begin with a network formed at volume V0 and then swollen with solvent to a new volume V.\nThe volume fraction of polymer in the resulting gel is 9152\u2014\n\u2014 VO/V (assuming additivity of volumes).\nWe assume that the swelling1s isotrOpic, so that the x, y, and 2 components of the end-to-end vector\nof each strand are increased by a factor of (V/V0)l/3 =qb-l/3. We will reference the deformation of\nthe already\n2swollen\nnetwork to the isotrOpically swollen dimensions so that the reference state\nterms (to, yo, and 20) in Equation 10.5.9 should each be multiplied by a factor of qb;2/.\n/3\nThus\nEquation 10.5.11 describing the entrOpy of deformation becomes\n"]], ["block_3", ["softer, but also in part to plasticize the styrene domains and lower their glass transition\ntemperature (see Section 12.6).\n2.\nSoft contact lenses. Soft contact lenses are an example of a hydrogel\u2014a network in which the\npolymer is either water soluble or at least water compatible. The original soft contact lenses\nwere made largely from cross-linked poly(hydroxyethyl methacrylate), and were developed in\nthe late 19603.\n"]], ["block_4", ["Clearly, swollen networks are of fundamental importance in many areas of materials and\nbiological science. In this section we will brie\ufb02y address two aspects of the swelling phenomenon.\nFirst, we consider how the expression for the modulus of an ideal elastomer changes when solvent\nis incorporated. Then we consider swelling equilibrium; how much solvent can a network take up?\n"]], ["block_5", ["Swelling of Gels\n411\n"]], ["block_6", ["(The same result could be obtained by taking the front factor of Equation 10.6.5, (h2)/ (ho2), and\nrealizing that (h2')13 increased by the same factor of Q5;.) Now let us apply an elongation along\nthe x direction, so )1 =)1, and )1y \u2014)1 \u2014,1/\\/X and compute the force as in Equation 10.5.13:\n"]], ["block_7", ["110'; V0.\n_f_\n\u20142/3\n1\n"]], ["block_8", ["where now LS refers to the swollen but unstrained network. To compute the stress, again referenced\nto the swollen but otherwise undeformed network, we divide by LE, recognizing that L3 V,\n"]], ["block_9", ["10.7.1\nModulus of a Swollen Rubber\n"]], ["block_10", ["Comparing this result with Equation 10.5.\n115,\nthe main result of swelling is that the stress in the\nswollen network18 reduced by a factor of d12/300mpared to the original network, and the modulus\nis reduced by the same factor when computed for constant cross-sectional area.\n"]], ["block_11", [{"image_1": "422_1.png", "coords": [35, 481, 291, 527], "fig_type": "molecule"}]], ["block_12", [{"image_2": "422_2.png", "coords": [41, 558, 185, 631], "fig_type": "molecule"}]], ["block_13", ["3A5\nT\n3A5\nVekT\n_2/3\n1\n= \u2014T\n\u2014 \u2014\n=\n)1 \u2014 \u2014\n10.7.2\nf\n(\n)3L:L, ( a1 )\nL,\n2\n)1?\n(\n)\n"]], ["block_14", ["AS\nIce\u2014\u201d3()12 +21; +213 3)\n(10.7.1)\n"]], ["block_15", ["_\u2014kT\u2014\u201d291.1:(1 \u2014/\u00a7)\n(10.7.3)I\n"]], ["block_16", [{"image_3": "422_3.png", "coords": [107, 485, 281, 518], "fig_type": "molecule"}]]], "page_423": [["block_0", [{"image_0": "423_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "423_1.png", "coords": [29, 634, 171, 677], "fig_type": "molecule"}]], ["block_2", [{"image_2": "423_2.png", "coords": [33, 371, 192, 401], "fig_type": "molecule"}]], ["block_3", ["A simple and very qualitative way to understand the origin of such a term is that there is an\nadditional entropy gain for the placement of the end of each strand in space, as the volume of the\nnetwork plus solvent increases. In this way the new term (which could be written as\n\u2014+ln 0.52) is\nequivalent to an ideal entropy of mixing contribution as derived in Section 7.3.1. Flory\u2019s analysis\nincluded a more explicit calculation of the entropy change associated with the cross-linking\nprocess, but in fact the appropriateness of the new logarithmic term in Equation 10.7.7 has not\nbeen without controversy [11]. Because the product of the three extension ratios is unity for a\nconstant volume deformation, this term drops out in the unswollen network, so we have not missed\nanything by omitting it in the previous sections. We retain it now to be consistent with the Flory\u2014\nRehner result and we note that the factor of 0:3 in Equation 7.7.10 from the Flory\u2014Krigbaum theory\nfor excluded volume also originates from inclusion of this term.\nAs in the consideration of the swollen network modulus above, we call the volume of the initial\npiece of polymer V0 and the swollen volume V, but now the extension ratios are\n"]], ["block_4", ["In this case X represents a combination of interaction energies between solvent and monomer and\nbetween solvent and cross\u2014linking unit, but for low cross\u2014link densities and/or systems such as\nstyrene/divinylbenzene where the monomer and the cross-linker are chemically very similar, this\ncomplication is not important. The elastic part of AG is assumed to be purely entropic\n(AGel TASBI) and we simply invoke the rubber elasticity result (Equation 10.5.11):\n"]], ["block_5", ["Now we address the question of how much solvent a network can take up. Imagine we have a piece of\nlightly cross-linked polymer and we immerse it in a beaker of solvent. We have actually considered a\nvery similar situation before. In Section 7.7 we treated a single polymer coil as an osmotic pressure\nexperiment and saw how the good solvent exponent, 12: 3/5, could be obtained from balancing the\nosmotic swelling ofthe coil (the solvent wants to dilute the monomers) and elastic resistance to swelling\n(loss ofconformational entropy as the chain stretches out). Essentially the same balance will occur here.\nThe chemical potential gradient will drive solvent into the piece of network, but the elasticity of the\nnetwork will resist unlimited deformation. A state of swelling equilibrium will be reached, from which\nit is possible to determine both X and the average molecular weight between cross\u2014links.\nThe earliest theory of swelling equilibrium was that of Flory and Rehner [10], and we will now\nrederive their main result. We begin by considering the free energy of the swollen network to be\ncomposed of two parts, one due to the mixing of solvent and polymer and the other due to\ndistortion of the network:\n"]], ["block_6", ["The former part can be represented by the Flory\u2014~Huggins theory expression (Equation 7.3.13), but\nwith the network contributing no entropy of mixing (N \u2014> oo):\n"]], ["block_7", ["where 12,, is the number of effective strands. However, it is time to address a complication. By a\ndifferent analysis than that used in this chapter, Flory [4] obtained an expression for the entropy\nthat contains an additional logarithmic term not present in Equation 10.7.6:\n"]], ["block_8", ["412\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_9", ["10.7.2\nSwelling Equilibrium\n"]], ["block_10", ["V\n1/3\nA, = A, = A, =A =\n(10.7.8)\n"]], ["block_11", ["05,1: \n{213. +213 +213 3}\n(10.7.6)\n"]], ["block_12", ["\u2014lcr/,a\nASel \n2\n{213, + 213+ 213* 3 w1n(}tx}ty}tz)}\n(10.7.7)\n"]], ["block_13", ["AG :AGm + 06,,\n(10.7.4)\n"]], ["block_14", ["AGm :RT{n11nqb1 + mag}\n(10.7.5)\n"]]], "page_424": [["block_0", [{"image_0": "424_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "424_1.png", "coords": [31, 228, 179, 269], "fig_type": "molecule"}]], ["block_2", ["Thus\n"]], ["block_3", ["The other two derivatives on the right-hand side of Equation 10.7.10 are also straightforward. The\nfirst is the solvent chemical potential of the Flory\u2014Huggins theory (Equation 7.4.14) in the infinite\nmolecular weight limit:\n"]], ["block_4", ["At the point of swelling equilibrium, the chemical potential of the solvent inside the swollen\nnetwork will equal that in the surrounding pure solvent, and so we have\n"]], ["block_5", ["If we assume no volume change on mixing as before, then the volume fraction of polymer in the\nswollen gel, (02, is simply given by\n"]], ["block_6", ["where l?) is the molar volume of the solvent. If we differentiate this expression with respect to m\nwe obtain\n"]], ["block_7", ["Swelling of Gels\n413\n"]], ["block_8", ["The second is just\n"]], ["block_9", ["where we have multiplied and divided by Avogadro\u2019s number to put this term on a molar basis. By\ncombining Equation 10.7.10 and Equation 10.7.12 through Equation 10.7.14 we obtain\n"]], ["block_10", ["where (0,, (02 is used to emphasize that we have swelling equilibrium. The prefactor on the right-\nhand side includes the number of strands per unit volume in the original network, Ve/VO, and from\nEquation 10.6.7 this can be written as\n"]], ["block_11", ["or, using Equation 10.7.8 to eliminate )t, we obtain\n"]], ["block_12", [{"image_2": "424_2.png", "coords": [45, 429, 183, 468], "fig_type": "molecule"}]], ["block_13", ["1\nV\nA3\n(V0 + n] V] )\n_. = \u2014 =\n= \n1\n-\n.\n(02\nV0\nV0\n( 0 7 11)\n"]], ["block_14", ["6AG\nm it? Am (0\u2014)\nn1\nTap\n_\n(BAGm)\n+(8AG,1)(Q_)\n_\n8m\nTm\n021\nam up\n(10.7.10)\n"]], ["block_15", ["ASel \n2\n{3212 \u2014 3 \u2014 In 213}\n(10.7.9)\n"]], ["block_16", [{"image_3": "424_3.png", "coords": [47, 636, 149, 676], "fig_type": "molecule"}]], ["block_17", ["021\n1?\n2\n---\u2014-\u2014\n321 (am) 2\u20141\nV0\n(10.7.12)\n"]], ["block_18", ["Au] 0 RT{ln(1\u2014 cbz) + a, + x0, + N... 70 (X 53)}\n(10.715)\n"]], ["block_19", ["VB\npNav\n2Mx\n _ \n1_\n"]], ["block_20", ["I\u2019e\nV1\n1n(1\u2014 <1>.)+ a, +270: \n\u2014\u2014 (\u201cbe \u2014 \u00abbi/3)\n(10.7.16)\n_\nNa, V0\n'2\u2014\n"]], ["block_21", ["673GB]\n_RTVe\n3\n(\n021 \nN...\n(321 i)\n(10.7.14)\n"]], ["block_22", ["aAGm\n"]], ["block_23", ["6\u201911 \n= RT{1n(1\u2014 (02) + 492 + X\ufb01bg}\n(10.7.13)\nTm\n"]], ["block_24", [{"image_4": "424_4.png", "coords": [67, 132, 270, 205], "fig_type": "molecule"}]], ["block_25", [{"image_5": "424_5.png", "coords": [82, 379, 234, 408], "fig_type": "molecule"}]], ["block_26", [{"image_6": "424_6.png", "coords": [103, 152, 260, 196], "fig_type": "molecule"}]], ["block_27", [{"image_7": "424_7.png", "coords": [158, 501, 316, 533], "fig_type": "molecule"}]]], "page_425": [["block_0", [{"image_0": "425_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "425_1.png", "coords": [26, 431, 201, 465], "fig_type": "molecule"}]], ["block_2", ["414\nNetworks. Gels, and Rubber Elasticity\n"]], ["block_3", ["where Mx is the average molecular weight between cross-links. Thus if X is known, Equation\n10.7.16 and Equation 10.7.17 can be solved for M, using the measured value of V (although in\npractice it is more common to actually weigh the dry and swollen polymer, and convert to volume,\nthrough the known densities of polymer and solvent). The following numerical example gives an\nidea of the magnitude of the swelling.\n"]], ["block_4", ["where 108 cm3 is the molar volume of cyclohexane.\nReturning to the Flory\u2014Rehner expression, that is, Equation 10.7.16, we have values for all the\nquantities except the desired (1),, but the equation does not reduce to a simple algebraic expression.\nTo proceed we anticipate that (1)., << 1, and then the left-hand side can be simplified to ( X 0.5) (1)5\nusing the expansion ln(l (be) 3 \u2014q_')e (122/2. Similarly, on the right-hand side we can expect\n01/3 >> 0.3/2. This gives\n"]], ["block_5", ["The last ingredient we need in order to apply the Flory\u2014Rehner expression is a value for X. Back in\nSection 7.6.2 we considered ways to estimate X using solubility parameters, and the necessary\nvalues were given in Table 7.1. Using Equation 7.6.6 we have\n"]], ["block_6", ["The polyisoprene chains had M 150,000 with the extent of vulcanization 0.003. The monomer\nmolecular weight is 68 g/mol, so the average number of cross-links per chain is (150,000 X 0.003)/\n"]], ["block_7", ["Calculate the predicted volumetric swelling for the polyisoprene network of Example 10.1, when\nexposed to a reservoir of cyclohexane at room temperature.\n"]], ["block_8", ["A test of the Flory\u2014Rehner theory (Equation 10.7.16) is provided by the data in Figure 10.16.\nA variety of butyl rubbers with differing degrees of cross-linking were swollen to equilibrium and\nthen extended to )t =4. The results show a power law dependence of force on (he\u2018s/3. This is\nconsistent with Equation 10.7.16 in the limit of (be \u2014+ 0, as can be seen by taking this limit on both\nsides of the equation as in the preceding example. (Note that force G l/Mx.)\n"]], ["block_9", ["In this chapter we have considered two routes to network formation, cross-linking of preformed\nchains and direct polymerization including multifunctional monomers. The former is the preferred\nroute to elastomers, or lightly cross-linked, low Tg materials, and the latter is commonly employed\nto produce thermosets for high-temperature applications. We then developed the theory of rubber\nelasticity in some detail, covering general thermodynamic aspects, the statistical theory in the ideal\n"]], ["block_10", ["or (126%0046. This gives a volumetric expansion of 1/0.046 =22 times. This value also justifies\nthe small (1),, approximation that we have used to simplify the algebra. However, see Problem 15 for\na numerically more realistic version of this calculation.\n"]], ["block_11", ["68 6.6, and thus M, as 150,000/(66) 23,000. Utilizing Equation 10.7.17 and density of\n0.91 g/cm3 for polyisoprene gives\n"]], ["block_12", ["Solution\n"]], ["block_13", ["Example 10.4\n"]], ["block_14", ["10.8\nChapter Summary\n"]], ["block_15", [{"image_2": "425_2.png", "coords": [42, 245, 198, 280], "fig_type": "molecule"}]], ["block_16", [{"image_3": "425_3.png", "coords": [45, 324, 257, 356], "fig_type": "molecule"}]], ["block_17", ["2 x 300\n-12\nX _RT\n(51 52? m\n(8.2 8.1)2 a: 0.002\n"]], ["block_18", ["V\n0.91\n46 000\nC\n:\n\u2014 _,___\n: \n1\n\u20145\nl\nd\n3\nNav v0\n23,000 (\n150,000)\nx\n0\nmo stran s/cm\n"]], ["block_19", ["5/3 3 \n=\n.\n(be\n0.5 \u2014 0.002\n0 0058\n"]], ["block_20", ["108\n"]]], "page_426": [["block_0", [{"image_0": "426_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 10.16\nForce at constant extension (A24) versus equilibrium swelling for butyl rubbers with\ndifferent degrees of cross-linking, showing the predicted dependence on (pg/3. (Data from Flory, P.l., Ind.\nEnghg. Chem, 38, 417, American Chemical Society, 1948. Reproduced from Treloar, L.R.G., The Physics of\nRubber Elasticity, 3rd ed., Clarendon Press, Oxford, 1975. With permission.)\n"]], ["block_2", ["case, some modifications to account for nonideal features of the response of real materials, and the\ncase of networks swollen with solvent. The main points may be summarized as follows:\n"]], ["block_3", ["Chapter Summary\n415\n"]], ["block_4", ["1.\nWhen cross\u2014linking preformed chains with weight-average degree of polymerization NW, the\n\u201cgel point\u201d is predicted to occur when the fraction of monomers participating in cross\u2014links is\nequal to l/Nw. In practice, this tends to underestimate the necessary extent of reaction.\nFor multifunctional monomers, explicit predictions for the gel point can be developed using\nthe principle of equal reactivity and probability arguments appropriate to the particular\npolymerization mechanism and combination of monomers. In practice, these expressions\nalso tend to underestimate the extent of reaction needed to reach the gel point.\nAnalysis of rubber elasticity via macroscopic thermodynamics is relatively straightforward,\nthe main new ingredient being the incorporation of the work of deformation into the free\nenergy. An ideal elastomer is defined as one for which the force resisting deformation is\nentirely entropic, which is a reasonable approximation for many rubbery materials.\nThe molecular basis of rubber elasticity rests in the reduction of conformational degrees of\nfreedom when a single Gaussian chain is extended. A single Gaussian chain acts as a Hooke\u2019s\nlaw spring, with a stiffness that is proportional to absolute temperature.\nStraightforward expressions for the force required to deform an ideal elastomer are obtained\nby modeling the network as a collection of Gaussian strands and by making an assumption as\nto how the macroscopic deformation is transmitted to each strand. The resulting shear and\nextensional moduli are proportional to the number of strands per unit volume.\nNetworks or gels are often capable of absorbing more than 100 times their own weight in\nsolvent, a phenomenon that is central to many applications, and that can be understood as a\nsimple balance between the osmotic drive to dilute the polymer and the entropic resistance to\nstrand extension.\nAlthough the statistical theory of rubber elasticity captures the main features of a wide variety\nof experimental phenomenology, attempts to bring the theory into quantitative agreement with\nexperiment have met with rather mixed success.\n"]], ["block_5", [{"image_1": "426_1.png", "coords": [37, 40, 310, 237], "fig_type": "figure"}]], ["block_6", ["(N/mm3))\n"]], ["block_7", ["Iog(force\n"]], ["block_8", ["=\n"]], ["block_9", ["at\nA.\n"]], ["block_10", ["I\nl\n_l\n[\n0.7\n0.8\n0.9\n1.0\nloan/rte)\n"]]], "page_427": [["block_0", [{"image_0": "427_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["TPJ. Flory, N. Rabjohn, and M.C. Shaffer, J. Polym. Sci., 4, 225 (1949).\n"]], ["block_2", ["(e) AB + 33\n(D A232 + A2 + 32\n6.\nSuppose you have a balloon made of an ideal elastomer that is in\ufb02ated to a reasonable size with\nan ideal gas at room temperature. If the temperature of the balloon plus gas system is then\nincreased to 100\u00b0C, will the balloon expand, contract, or stay the same size? Justify your answer.\n7.\nFind the relation between the (true) stress 0 and the strain A for a piece of ideal rubber in\nbiaxial extension. Assume the rubber has initial area A0 and thickness do, and let the final area\nbe A A2240.\n8.\nUse the result from the previous problem to calculate the relation between the pressure of an\nideal gas, p, inside a balloon made from an ideal elastomer, expanded to a radius R =AR0,\nwhere Rois the initial radius. Use a version of the Young\u2014Laplace equation to relate the excess\npressure (inside the balloon minus outside) to the stress in the rubber, p = 2do/R, where d is\nthe thickness of the balloon skin. Empirically, it often seems harder to \u201cget started\u201d blowing\nup a balloon, than to blow it up further beyond a certain point. Explain this observation based\non your result for p versus )1.\n"]], ["block_3", ["How important for this system is the end group correction introduced in Equation 10.6.7?\n3.\nDevelop the equivalent to Equation 10.2.3 and Equation 10.2.5 for the third system in Figure\n10.4, that is, AB + BB + A3.\n4.\nThe Carothers equation (Equation 10.2.9) can also be used as the basis of an estimate of the\nextent of reaction at gelation. Consider the value implied for each of the parameters in the\nCarothers equation at the threshold of gelation, and derive a relationship between pc andf on\nthe basis of this consideration. Compare the predictions of the equation you have derived with\nthose of Equation 10.2.5 for a mixture containing 2 mol A3, 7 mol AA, and 10 mol BB.\nCriticize or defend the following proposition: the Carothers equation gives higher value for pC\nthan Equation 10.2.5 because the former is based on the fraction of reactive groups that have\nreacted and hence considers wasted loops that the latter disregards.\n5.\nCategorize the following mixtures as to whether they can form linear, branched, or network\nStructures:\n"]], ["block_4", ["416\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_5", ["1.\nA constant force is applied to an ideal elastomer, assumed to be a perfect network. At an initial\ntemperature Ti the length of the sample is Li. The temperature is raised to Tf and the final\nlength is Lf. Which is larger: L, or Lf (remember F is a constant and Tf > Ti)? Suppose a wheel\nwere constructed with spokes of this same elastomer. From the viewpoint of an observer, the\nspokes are heated near the 3 o\u2019clock positionwsay, by exposure to sunlight\u2014while other\nspokes are shaded. Assuming the torque produced can overcome any friction at the axle, would\nthe observer see the wheel turn clockwise or counterclockwise? How would this experiment\ncontrast, in magnitude and direction, with an experiment using metal spokes?\n2.\nAn important application of Equation 10.5.15 is the evaluation of M,; R]. Flory, N. Rabjohn,\nand M.C. Shaffer measured the tensile force required for 100% elongation of synthetic rubber\nwith variable cross-linking at 250C.)r The molecular weight of the uncross-linked polymer was\n225,000, its density was 0.92 g/cm3, and the average molecular weight of a repeat unit was 68.\nUse Equation 10.5.15 to estimate M, for each of the following samples and compare the\ncalculated value with that obtained from the known fraction of repeat units cross-linked:\n"]], ["block_6", ["Problems\n"]], ["block_7", ["Fraction cross-linked\n0.005\n0.010\n0.015\n0.020\n0.025\nF/A(lb-force/in.2)\n61.4\n83.2\n121.8\n148.0\n160.0\n"]], ["block_8", ["(3) A2 + B2 + AB\n('3)\nA132 + A2\n(0) AB + AB2\n(d) A3 + 32 + A2\n"]], ["block_9", [{"image_1": "427_1.png", "coords": [60, 248, 328, 290], "fig_type": "molecule"}]]], "page_428": [["block_0", [{"image_0": "428_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018S.B. Smith, L. Finzi, and C. Bustamante, Science, 258, 1122 (1992).\n1C. Bustamante, J. Marko, E. Siggia, and s. Smith, Science, 265, 1599 (1994).\n"]], ["block_2", ["Problems\n417\n"]], ["block_3", ["15.\n"]], ["block_4", ["13.\n"]], ["block_5", ["14.\n"]], ["block_6", ["16.\n"]], ["block_7", ["9.\nShow that Equation 10.3.2 and Equation 10.3.3 are equivalent definitions of Poisson\u2019s ratio.\n10.\nUse Equation 10.4.22 and the data in Figure 10.8a to assess whether or not the heat capacity\n"]], ["block_8", ["11.\n"]], ["block_9", ["12.\n"]], ["block_10", [{"image_1": "428_1.png", "coords": [42, 330, 193, 365], "fig_type": "molecule"}]], ["block_11", ["f\u2014k_T(\u2014_1 _l+.,.1_)\n\u201861,\n4(1\u201421/1.)2\n4\nL\n"]], ["block_12", ["at constant length, CL, is comparable to typical values of C], for rubber of 2 J/g/K. Be careful\nwith the stated units of stress (kg/cmz) in Figure 10.8a.\nEstimate the temperature increase in a rubber band when extended to A 8 at room\ntemperature. Assume CL is 2 J/g/K and p: 1 g/cc.\nThe following data represent force ( f, in picoNewtons) versus extension ()1, in microns) for a\nsingle A\u2014DNA molecule, measured at room temperature in salt solutions.)r These data\ntherefore represent an opportunity to test basic assumptions of the theory of rubber elasticity.\nTry to fit these data in three ways. First, use the Gaussian expression; restrict the fit to the low\nextension part of the curve. Second, try the inverse Langevin function, approximated in\nEquation 10.6.4. Third, try the following formula derived for the worm-like chain (recall\nChapter 6) in C. Bustamante et al.I\n"]], ["block_13", ["Comment on the success or failure of the various expressions and provide values for the\ncontour length, L, and the persistence length.\nA rubber band, made of a styrene\u2014butadiene random copolymer, is swollen to equilibrium in\ntoluene; the volume increases by a factor of 5. Taking X z 0.4, estimate the number of strands\nper unit volume, and therefore the extent of cross-linking. Estimate Young\u2019s modulus for\nboth the dry and swollen rubber bands.\nAccording to the Flory\u2014Rehner theory, what value of X would be required to get absolutely\nno uptake of solvent for a typical rubber? What value of X would restrict the uptake to less\nthan 5% by volume?\nIn Example\n10.4, the estimated value of X was 0.002, which is not realistic. Revisit\nthe discussion in Section 7.6, and use a more realistic estimate for X. How much does the\nrubber swell at equilibrium in this case\u201c? Is it significantly different from the answer in\nExample 10.4?\nOn the following plot of a normalized equilibrium shear modulus (G/pRT) versus an inverse\n\u201ceffective\u201d molecular weight between cross-links (l/Mx) are two curves. The first is a\nstraight dashed line with unit slope to \u201cguide the eye\u201d (and the brain). The second represents\nthe modulus of a real polymer as it undergoes progressive cross-linking (the arrow represents the\ncourse of the\nmodulus with\nextent of reaction).\nWhy\ndoes the\nmodulus shoot\nup\nat l/sa3 x 10\u20185 mol/g? Why does the modulus exceed the unit slope line, just after\nl/sa 3 x 10\n\u2014 5 mol/g? Estimate Me for this polymer. (Note that the last two questions will\nbe easier to answer after reading Section 11.6.)\n"]], ["block_14", ["10.1\n0.0338\n25.0\n0.367\n30.3\n3.03\n10.8\n0.0335\n25.6\n0.364\n31.2\n5.57\n17.6\n0.102\n26.4\n0.486\n31.3\n6.82\n18.2\n0.102\n27.1\n0.604\n31.7\n11.9\n22.4\n0.218\n28.5\n1.08\n32.0\n12.2\n23.0\n0.217\n29.0\n1.53\n32.1\n18.3\n24.3\n0.382\n30.3\n3.39\n32.2\n17.0\n"]], ["block_15", ["h (um)\nf(PN)\n11mm)\nf(PN)\nh (um)\nf(PN)\n"]]], "page_429": [["block_0", [{"image_0": "429_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "429_1.png", "coords": [22, 49, 275, 302], "fig_type": "figure"}]], ["block_2", ["Flory, P.J., Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\nGraessley,\nW.W.,\nPolymeric\nLiquids\nand\nNetworks:\nStructure\nand\nProperties,\nGarland\nScience,\nNew York, 2003.\nMark, J.E. and Erman, 3., Rubber Elasticity\u2014A Molecular Primer, Wiley, New York, 1988.\nOdian, G., Principles of Polymerization, 4th ed., Wiley, New York, 2004.\nRubinstein, M. and Colby, R.H., Polymer Physics, Oxford University Press, New York, 2003.\nTreloar, L.R.G., The Physics of Rubber Elasticity, 3rd ed., Clarendon Press, Oxford, 1975.\n"]], ["block_3", ["413\nNetworks, Gels, and Rubber Elasticity\n"]], ["block_4", ["References\n"]], ["block_5", ["Further Readings\n"]], ["block_6", ["Goodyear, C., US. Patent 3, 633 (1844).\nStockmayer, W.H., J. Chem. Phys., 11, 45 (1943); 12, 125 (1944).\nFlory, P.J., Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY, 1953.\nCarothers, W.H., Trans. Faraday Soc, 32, 39 (1936).\nGough, J., Mem. Lit. Phil. Soc. Manchester, 1, 288 (1805).\nJoule, J.P., Phil. Trans. R. Soc, 149, 91 (1859).\nJames, HM. and Guth, E., J. Chem. Phys., 11, 455 (1943); 15, 669 (1947).\nKuhn, W. and Griin, F., Kolloidzschr., 101, 248 (1942).\nMooney, M., J. Appl. Phys,\n1 l, 582 (1940); Rivlin, R.S., Phil. Trans. R. Soc, A241, 379 (1948).\nFlory, R]. and Rehner, J., J. Chem. Phys, 11, 512 (1943).\nTreloar, L.R.G., The Physics ofRubber Elasticity, 3rd ed., Clarendon Press, Oxford, 1975.\nf\u2018QEOW\ufb02P\u2018E\u2018J-\u2018Wh\u2019f\ny\u2014ip\u2014a\n"]], ["block_7", ["G/t\u2018T\n"]]], "page_430": [["block_0", [{"image_0": "430_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this chapter we extend the study of the dynamic properties of polymer liquids in two new\ndirections. In Chapter 9 we considered only dilute solutions, but here we will consider also very\nconcentrated solutions and molten polymers. In Chapter 9 we also focused on the steady\u2014flow\nviscosity, or the diffusion over long time intervals; now we will examine the time\u2014 or frequency\u2014\ndependent response of polymer liquids to an imposed deformation or force. This response can be\ncharacterized by a variety of material functions, such as the viscosity, the modulus, and the\ncompliance. In general, we will find that polymer liquids are viscoelastic, i.e., their behavior is\nintermediate between (elastic) solids and (viscous) liquids. The phenomenon of viscoelasticity is\nfamiliar to anyone who has played with Silly Putty. If you roll some into a ball, and leave it for a\nfew hours, it flows to adopt the shape of its container. This behavior is that of a liquid; it just takes\na long time because the viscosity is very high. On the other hand, if you stretch a sample very\nrapidly, and immediately release one end, the sample will partially recover toward its original\ndimensions. This recovery is an elastic response, and is more typical of solids than liquids. The\nprevious chapter concerned the elasticity of polymer networks, and important results from that\ndiscussion will be directly incorporated into our treatment of viscoelasticity.\nViscoelasticity is one of the most distinctive features of polymers. As virtually all polymer\nmaterials are processed in the liquid state, viscoelasticity plays a central role in the optimization\nand control of processing. On the other hand, we will see in Chapter 12 and Chapter 13 that the\nconcepts developed here will also have broad application to the mechanical properties of solid\npolymers. Furthermore, the viscoelastic properties can be readily measured and therefore provide\nan additional route to molecular characterization, particularly for polymers that are difficult to\ndissolve in convenient solvents. Finally, measurement of the viscoelastic response of a polymeric\nmaterial provides direct and detailed information about how long it will take for that material to\nrespond to any kind of perturbation.\nTo gain a glimpse at what lies ahead, Figure 11.1 shows the steady \ufb02ow viscosity, 1), of five\nsamples of poly(0t\u2014methylstyrene) in the molten state as a function of molecular weight, M, at\n186\u00b0C. The viscosities are very large; they are about 1010\u201410l4 times larger than the viscosity of\nliquid water (about 0.01 P). Furthermore, the viscosity is a very strong function of molecular\nweight. The plot is in a double logarithmic format and the indicated straight line has a slope of\n3.24. Thus, 72 ~M3-24, which means that doubling the molecular weight is sufficient to increase 72\nby about a factor of 10. In this chapter we will provide a molecular picture for the origin of this\nresponse and see that this behavior is typical of all \ufb02exible polymers. Of course, a central objective\nof molecular models of polymer behavior is to understand the molecular weight dependence of any\nexperimental quantity.\nFigure 11.2 shows a quantity called the stress relaxation modulus (which we will define a little\nlater) as a function of time for the same five polymers. This quantity is also plotted in a double\u2014\nlogarithmic format. For now, we can just compare these modulus values with those (time\u2014\nindependent) values typical of other materials, including the cross\u2014linked rubber discussed in the\nprevious chapter. At very short times in Figure 11.2, the modulus approaches 109 Pa (1 GPa,\n"]], ["block_2", ["11.1\nBasic Concepts\n"]], ["block_3", ["Linear Viscoelasticity\n"]], ["block_4", ["11\n"]], ["block_5", ["419\n"]]], "page_431": [["block_0", [{"image_0": "431_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["420\nLinear Viscoelasticity\n"]], ["block_2", ["Figure 11.2\nStress relaxation modulus, G(t), versus time for molten poly(or-methyl styrene) at 186\u00b0C;\nsamples A\u2014l through A\u2014S correspond to the five samples in Figure 11.1. The arrow locates the longest\nrelaxation time for sample A-S, as discussed in Example 11.1. (Data taken from Fujimoto, T., Ozaki, N., and\nNagasawa, M., J. Polym. Sci. Part A-Z, 6, 129, 1968. With permission.)\n"]], ["block_3", ["Figure 11.1\nViscosity versus molecular weight for molten poly(0t-methyl styrene) at 186\u00b0C. (Data taken\nfrom Fujimoto, T., Ozaki, N., and Nagasawa, M., J. Polym. Sci. Part A-2, 6, 129, 1968.)\n"]], ["block_4", ["g\n10 L\n_\n"]], ["block_5", [{"image_1": "431_1.png", "coords": [40, 372, 260, 600], "fig_type": "figure"}]], ["block_6", [{"image_2": "431_2.png", "coords": [42, 49, 282, 284], "fig_type": "figure"}]], ["block_7", ["1011\nE\u2014\nn=24x10\u20147M3.24\nE.\n3\ni\nE\nJ\n"]], ["block_8", ["1012\n:\n|\n|\nI\n|\nI\nI\nI\n[I\nl\nl\n|\nI\n|\n|\nl\nj\n"]], ["block_9", ["109 l\u2019\nJ:\n"]], ["block_10", ["108\n|\nl\nl\nl\nl\nI\nI ll\n1\nl\nl\nL]\nl\nl\n104\n105\n106\n"]], ["block_11", ["10+\nl\n"]], ["block_12", ["E\nE\n"]], ["block_13", ["E\ni\n"]], ["block_14", ["\u20144\n\u20142\n0\n2\n4\n6\nLog time (s)\n"]], ["block_15", ["MW\n"]], ["block_16", ["A-1 A-2 A-3A-4A\u20145\n"]]], "page_432": [["block_0", [{"image_0": "432_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "432_1.png", "coords": [31, 530, 212, 632], "fig_type": "figure"}]], ["block_2", ["In a typical experiment, a sample confined in a particular geometry is subjected to a displacement,\nand the resulting force is measured. An example is shown in Figure 11.3, for the deformation\ncalled simple shear. This is exactly the geometry used to discuss viscosity in Chapter 9.1. The\nmaterial is confined between two parallel plates dy apart, and if one surface is moved a distance dx,\nwe say that the sample has been subjected to a strain, 7/, :dx/dy; \u201cy is thus dimensionless. The\nvelocity of the plate vx dx/dt, and d(dx/dy)/dt :7'\u00bb is called the strain rate or shear rate. It takes\nthe application of a force to accomplish the deformation; alternatively, we can think of the material\nexerting a force on the moving plate. The total force will depend on the area of the plate in contact\nwith the material and thus we consider the force per unit area, or stress, 0'. There are several\ndifferent kinds of deformation geometries, such as uniaxial elongation, biaxial elongation, simple\nshear, etc. In all cases it is possible to define a stress, strain, and strain rate in an analogous fashion.\nIn the most general case, the stress and the deformation in the material should be represented as\ntensor quantities. In this chapter we will restrict our attention to the simplest case of shear \ufb02ow and\nshear stress only. This avoids the need to employ tensor algebra; the mathematics will be\ninvigorating enough as it is. We will use the symbols 0', y, and \u201cj! to denote shear stress, shear\nstrain, and shear rate, respectively, just as in Chapter 9.\n"]], ["block_3", ["In general, a viscosity is defined as the ratio of a stress to a strain rate, and in shear we begin with\nNewton\u2019s relation (Equation 9.1.3)\n0':n\n(11.1.1)\n"]], ["block_4", ["or 1010 dyn/cmz); this can be compared with steel (80 GPa), silica (20 GPa), and granite (10 GPa).\nOn the other hand, at intermediate times, say l\u2014lOO s, the modulus has dropped to a value close to\nthat of a typical rubber, approximately 1 MPa. Finally, at even longer times, the modulus tends to\nzero. Another interesting observation from Figure 11.2 is that in the early time, high modulus state,\nthe modulus is independent of M, but at long times, the time decay of the modulus depends very\nmuch on M. In this chapter we will explain all of these features and even show how the viscosity\nvalues in Figure 11.1 were actually calculated from the data in Figure 11.2, rather than measured\ndirectly.\nIn the first three sections, we will define the basic terms and concepts, examine simple models\nthat reveal important features of viscoelastic response, and show how the different material\nfunctions can be related to one another in a model-independent way. The subsequent four sections\nwill examine molecular models that, collectively, are remarkably successful in capturing the\nviscoelastic response of polymers all the way from dilute solution to the melt. The final section\ncovers some aspects of experimental rheology, the science of \ufb02ow and deformation of matter.\n"]], ["block_5", ["Figure 11.3\nIllustration of simple shear \ufb02ow between two parallel plates.\n"]], ["block_6", ["11.1.2\nViscosity, Modulus, and Compliance\n"]], ["block_7", ["Basic Concepts\n421\n"]], ["block_8", ["11.1.1\nStress and Strain\n"]]], "page_433": [["block_0", [{"image_0": "433_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["422\nLinear Viscoelasticity\n"]], ["block_2", ["Two crucial limiting cases in this chapter are viscous response and elastic response. Viscosity\nre\ufb02ects the relative motion of molecules, in which energy is dissipated by friction. It is a primary\ncharacteristic of a liquid. A liquid will always \ufb02ow until the stress has gone away and it will\ndissipate energy as it does so. In contrast, elasticity re\ufb02ects the storage of energy; when a spring is\nstretched, the energy can be recovered by releasing the deformation. A solid subjected to a small\nstrain is primarily elastic, in that it will remain deformed as long as the force is still applied. In\n\ufb02exible polymers, the elasticity arises from the many conformational degrees of freedom of each\nmolecule and from the intertwining of different chains; it will turn out to be primarily entropic in\norigin, just as in Chapter\n10. When the material is subject to a deformation, the individual\nmolecules respond by adopting a nonequilibrium distribution of conformations. For example, the\nchains on average may be stretched and/or oriented in the direction of \ufb02ow; in so doing they lose\nentropy. Left to themselves, the molecules will relax back to an isotropic, equilibrium distribution\nof conformations, just like a spring. As they relax, the relative motion of the molecules through the\nsurrounding \ufb02uid dissipates the stored elastic energy. It is this interplay of viscous dissipation\nduring elastic recovery that underlies the viscoelastic prOpeIties of polymer liquids. (If you are at\nall familiar with elementary electronic circuits, these concepts of energy dissipation and storage\nare well known; a resistor (resistance R) is the dissipative element, and the capacitor (capacitance\nC) is the storage element. Voltage (V) plays the role of force, current (1) the role of velocity, and\ncharge (Q) the role of displacement. Thus Ohm\u2019s law (Vl) is the analog of Newton\u2019s law\n(Equation 11.1.1), and for a capacitor V: CQ in analogy to Equation 11.1.2. Furthermore, the\nmathematical results we will derive in the next section for simple mechanical models will be\nanalogous to the results for V and I in simple RC circuits.)\nIn experimental measurements of the viscoelastic response, several different time histories are\nroutinely employed. In a transient experiment, at some specific time a strain (or stress) is suddenly\napplied and held; the resulting stress (or strain) is then monitored as a function of time. The former\nmode is called stress relaxation and the associated modulus the stress relaxation modulus,\nG(t) = G(t)/y. The latter mode (in parentheses) is called creep and the associated compliance the\ncreep compliance, J(t)=y(t)/o-. In a steady \ufb02ow experiment, the strain rate is constant; the\nresulting steady stress gives the steady \ufb02ow viscosity. Finally, in what is arguably the most\n"]], ["block_3", ["where n is the viscosity. Recall also from Section 9.1 that if n is independent of the magnitude of\ny, the \ufb02uid is said to be Newtonian. Many polymer \ufb02uids are non-Newtonian, but in general\nNewtonian behavior is recovered in the limit that 'j/ \u2014> 0. Similarly, in the mechanics of solid\nbodies a modulus is defined as the ratio of a stress to a strain, and in shear we have\no=Gy\n (HID \n"]], ["block_4", ["where G is the (shear) modulus (recall Equation 10.5.17). Note that this relation is essentially\nHooke\u2019s law for an ideal elastic spring, F lot; the modulus is a generalized spring constant, and\nthe sign has changed because we consider the force acting on the material. When G is independent\nof the magnitude of y, we say that the response is linear. This is generally true as y 0, but in\nexperiments we always need a finite strain to generate a measurable stress, and whether the\nresponse falls in the so-called linear viscoelastic limit is something that needs to be checked.\nFor the remainder of this chapter we will assume linear response, for simplicity; treatment of the\nnonlinear response is substantially more complicated. However, we should note that in commercial\npolymer systems the nonlinear response is interesting and very important, especially for processing\noperations where strains and strain rates are typically high. One further useful concept is that of a\ncompliance, J, which can be defined as the ratio of a strain to a stress. Therefore the compliance is\nthe inverse of the modulus, but it will turn out that when the time dependence is involved, J(t) is\nnot simply equal to l/G(t).\n"]], ["block_5", ["11.1.3\nViscous and Elastic Responses\n"]]], "page_434": [["block_0", [{"image_0": "434_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["for the spring, and\n"]], ["block_2", ["for the dashpot. At time (:0 we apply an instantaneous strain of magnitude yo, and hold it\nindefinitely; if we follow the stress as a function of time, it is a stress relaxation experiment.\nQualitatively we can anticipate what the response should be. At very short times, the dashpot will\nnot want to move; that is the whole point of a dashpot (i.e., a shock absorber). The spring, on the\nother hand, only cares about how much it is stretched, not how rapidly. Thus the initial deformation\nwill be entirely taken up by the spring. However, the stretched spring will then exert a force on the\n"]], ["block_3", ["We can learn a great deal about viscoelastic response through consideration of two simplified\nmodels, the so-called Maxwell [1] and Voigt [2] elements. We will examine the stress relaxation\nmodulus, creep compliance, and dynamic moduli of the Maxwell element, which illustrates a\nviscoelastic liquid. We will also derive the creep compliance of the Voigt element, which\nexemplifies a viscoelastic solid. Each element will turn out to have a characteristic time, 1',\nwhich determines the timescale of its response.\n"]], ["block_4", ["The Maxwell element consists of an ideal, Hookean spring with spring constant C? connected in\nseries with an ideal, Newtonian dashpot with viscosity if, as shown in Figure 11.4a. Thus the stress\nin the two components is given by\n"]], ["block_5", ["important mode, the sample is subjected to a sinusoidally time-varying strain at frequency cu. The\nresulting dynamic modulus, G*(w), is resolved into two dynamic moduli: one in-phase with\nthe strain, called 6\u2019, re\ufb02ecting the elastic component of the total response and one in\u2014phase\nwith the strain rate, called G\u201d, re\ufb02ecting the viscous response.\n"]], ["block_6", ["11.2.1\nTransient Response: Stress Relaxation\n"]], ["block_7", ["Response of the Maxwell and Voigt Elements\n423\n"]], ["block_8", ["Figure 11.4\nIllustration of (a) the Maxwell element and (b) the Voigt element.\n"]], ["block_9", ["11.2\nResponse of the Maxwell and Voigt Elements\n"]], ["block_10", [{"image_1": "434_1.png", "coords": [37, 282, 97, 308], "fig_type": "molecule"}]], ["block_11", ["0\u2018 5y\n(11.2.1a)\n"]], ["block_12", ["0. :\ufb01r),\n(11.2.11))\n"]], ["block_13", ["(a)\n(b)\n"]], ["block_14", [{"image_2": "434_2.png", "coords": [55, 455, 288, 663], "fig_type": "figure"}]], ["block_15", [{"image_3": "434_3.png", "coords": [143, 468, 281, 639], "fig_type": "figure"}]], ["block_16", ["\ufb01QQ\ufb01l\ufb02l/\u2018t'\n"]], ["block_17", [":>\n"]]], "page_435": [["block_0", [{"image_0": "435_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "435_1.png", "coords": [22, 39, 237, 218], "fig_type": "figure"}]], ["block_2", [{"image_2": "435_2.png", "coords": [31, 592, 164, 630], "fig_type": "molecule"}]], ["block_3", ["where the dot denotes the time derivative. We have defined the relaxation time 7 E i776, and from\nEquation 11.2.] it is clear that this ratio has units of time (\ufb01/G (U/y)/(cr/y) (7/7) t). The\nconcept of relaxation time is central to the material in this chapter. In essence, the relaxation time is\na measure of the time required for a system to return to equilibrium after any kind of disturbance.\nThe solution to Equation 11.2.4 is an exponential decay\n"]], ["block_4", ["This is a linear, first-order, homogeneous differential equation for 0'0):\n"]], ["block_5", ["424\nLinear Viscoelasticity\n"]], ["block_6", ["dashpot, which will slowly \ufb02ow in response. Ultimately, the spring will relax back to its rest length\nand there will be no more stress; the long time-response is that of a liquid.\nTo make this argument quantitative, we see that the total applied strain is distributed between\n"]], ["block_7", ["0:\nE\n"]], ["block_8", ["as shown in Figure 11.5. The initial stress, 00, is obtainedas suggested above: at the earliest times\nthe deformation is all in the spring, and therefore 00 070. (Note that an instantaneous deform-\nation would make 7 infinite and thus the stress in the dashpot would be infinite if it moved, so it\ndoes not.) The stress relaxation modulus is obtained as\nE = 5exp(\u2014t/7)\n(11.2.6)\n70\nGO) \n"]], ["block_9", ["the elements,\n"]], ["block_10", ["and because the strain is constant for t> 0,\n"]], ["block_11", ["1 << 1', but \ufb02ows until the stress has vanished fort >> 7. Thus the magnitude of the relaxation time\n"]], ["block_12", ["as shown in Figure 11.6, in both linear and logarithmic formats. The Maxwell model captures the\nmain feature of the stress\u2014relaxation response of any liquid; the material supports the stress for\n"]], ["block_13", ["Figure 11.5\nIllustration of the stress response 00\u2018) after a step strain of magnitude yo was applied at t: 0,\nfor the Maxwell element.\n"]], ["block_14", ["_\nI\nI\nl\n'\u2014'\nr\nI\n_\n7\u20190.\"\n4.\n'\nr\n\u2018_\n"]], ["block_15", ["0\n._\n-.\n.\nI\nI\n.\n.\nI\n.\n.\nI\nI\n.-\n"]], ["block_16", [{"image_3": "435_3.png", "coords": [39, 419, 106, 467], "fig_type": "molecule"}]], ["block_17", ["-\na\nI\n"]], ["block_18", ["1\n6+;0'=0\n(11.2.4)\n"]], ["block_19", ["C170\n.\n.\nd 00\u2018)\n0(1)\nE\n2 0 = yspring + ydashpot :\n\"d: _GT ?\n(11.2.3)\n"]], ["block_20", ["0(1) 0'0 exp(\u2014t/'r)\n(11.2.5)\n"]], ["block_21", ["70 yspring + ydashpot\n(1122)\n"]], ["block_22", ["0\nTime\n"]]], "page_436": [["block_0", [{"image_0": "436_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "436_1.png", "coords": [28, 36, 232, 240], "fig_type": "figure"}]], ["block_2", ["This value is marked as the arrow in Figure 11.2. It has a very simple physical interpretation: if we\nplace a chunk of this polymer on a desktop at 186\u00b0C (do not try this at home!) it will take 2\u20143 days\nbefore we will see it flow down into a puddle of liquid.\n"]], ["block_3", ["The relaxation time is given by the ratio of a viscosity to a modulus. The highest M sample in\nFigure 11.1 (M =4.6 X 105 g/mol) has a viscosity of 5.5 x 10\u201d P(= g/cm 8). It is not so obvious\nwhat value to take for the modulus in Figure 11.2, however, because G is a function of t. In Figure\n11.6b, the Maxwell model prediction for G(t) in a log\u2014log format looks like the long time-response\nin Figure 11.2, where G(t) falls from an apparent plateau value of about 3 x 106 dyn/cm2 to zero.\nIn the Maxwell model the plateau value is G. Using this result, then, we have\n"]], ["block_4", ["In a creep experiment, a sample is subjected to a constant force and the resulting deformation is\nmonitored as a function of time. The term creep implies that the deformation will be very slow,\nwhich in turn suggests that the sample should have a rather high viscosity; think of a glacier\n"]], ["block_5", ["Use the simple Maxwell model analysis of stress relaxation to estimate the longest relaxation time\nfor the highest molecular weight sample in Figure 11.1 and Figure 11.2.\n"]], ["block_6", ["is vital in determining the properties we experience. For example, a molten high M polymer not far\nabove its glass transition temperature may not flow over a timescale of hours (see Figure 11.2 and\nExample 11.1); in contrast, the stress relaxation time for water is measured in picoseconds. We\nshall also see that polymers, with their many degrees of internal conformational freedom, show\nmultiple relaxation times spread over many orders of magnitude, as also illustrated in Figure 11.2.\nHowever, before we get to that, we will pursue these simple models further.\n"]], ["block_7", ["Figure 11.6\nNormalized stress relaxation modulus G(t)/G(O) versus reduced time t/r for the Maxwell\nelement, plotted in (a) linear and (b) double logarithmic formats.\n"]], ["block_8", ["Solution\n"]], ["block_9", ["Example 11.1\n"]], ["block_10", ["11.2.2\nTransient Response: Creep\n"]], ["block_11", ["Response of the Maxwell and Voigt Elements\n425\n"]], ["block_12", ["3\n10\u20142.\n\u00e9\n\u00e9\n\u2014 .6:\na\n:\n<5\n:\n5g\n-\n\u00e9\n-\n5 10\u20143 5\n\u20185\n(5\n_\n(5\nE\nE\n"]], ["block_13", ["I\nI\nI\nI IIIII\nI\nI\nI IIIIII\nI\nI\nI\nI \u201cIII\nI\nI\nI\nI IIIJI.\nI\nI\nL\n0\n2\n4\n6\n8\n10\n10-3\n10\u20142\n1O\"I\n100\n101\n10\u201c2\n(a)\nI]?\n(b)\nIf?\n"]], ["block_14", ["11\nrzgm%l\u2014C%\u2014m2x 105 s=2.3 days\n"]], ["block_15", [{"image_2": "436_2.png", "coords": [55, 45, 421, 240], "fig_type": "figure"}]], ["block_16", ["-\n"]], ["block_17", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n-\n"]], ["block_18", ["I\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI.\nI\n-\n100\n"]], ["block_19", ["I\n10\u20185 s\n"]], ["block_20", ["3\n10\u20144\n\u2014E\n"]], ["block_21", [{"image_3": "436_3.png", "coords": [215, 27, 404, 231], "fig_type": "figure"}]], ["block_22", ["10\u20141 :\n'3\n"]], ["block_23", ["10\u20146\n"]], ["block_24", ["-\n"]], ["block_25", ["I\nI\nI\nI \u201cIII\nI\nI\nI\nIIIIII\nI\nI\nI\nI \"III\nI\nI\nI\nI I\"\nII\nI\nI\nI\nIII:\n"]]], "page_437": [["block_0", [{"image_0": "437_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 11.7a illustrates this behavior. At long times there is a steady-state response, with the strain\nincreasing linearly in time; the slope is the reciprocal of the viscosity. At short times there is a\ntransient response, re\ufb02ecting the initial deformation of the spring; in this model, it is instantaneous.\nConsequently, if the long time linear portion is extrapolated back to t: 0, there is a finite intercept,\nJ3, called the steady-state compliance. If the stress is suddenly removed at some instant after\nsteady \ufb02ow has been achieved, then the spring will retract but the dashpot will stop moving.\nConsequently there will be an elastic recovery of the \ufb02uid; this is also indicated in Figure 11.7a.\nThe amount of this recovery is called the recoverable compliance and if the \ufb02ow achieves steady\nstate, the recoverable compliance should be equal to J2.\nThe Voigt element places the two components in parallel, thereby enforcing equality of strain\nbetween the spring and the dashpot (Figure 11.4b). This combination serves to illustrate the\nresponse of a viscoelastic solid. In a creep experiment, in which a constant stress 00 is applied\nat I: 0 and held indefinitely, the strain will grow slowly, as the dashpot resists rapid deformation.\nAs the strain continues to grow, the resistance of the spring will take over, until at long times the\nstrain will saturate. From the definitions in Equation 11.2.1a and Equation 11.2.1b, we can write\n"]], ["block_2", ["Although the step strain and step stress experiments are both useful in characterizing the visco\u2014\nelastic response of a material, the most common experimental approach is to apply a sinusoidally\n"]], ["block_3", ["The solution to this equation is also an exponential decay, but with a twist:\n"]], ["block_4", ["Now the strain starts from zero and increases exponentially to its infinite time value of 00/6, as\nshown in Figure 11.7b. The compliance, J(t), can be written as\n"]], ["block_5", ["Here the strain in the dashpot is obtained by integrating the strain rate 7, where 5/ (To/ii is a\nconstant. Thus the compliance is given by\nW)l\nJ(t)=\u2014=\n+\u00e9t=Jeo+;1\u2014J\n(11.2.8)\n00\nG\n7i\n7?\n"]], ["block_6", ["425\nLinear Viscoelasticity\n"]], ["block_7", ["where 1,, has replaced 1/6; 1,, is called the equilibrium compliance.\n"]], ["block_8", ["moving under the in\ufb02uence of gravity. Now let us subject the Maxwell element to creep: we apply\n0'0 at time \u00a320 and watch y(t) evolve (using Equation 11.2.2):\n"]], ["block_9", ["11.2.3\nDynamic Response: Loss and Storage Moduli\n"]], ["block_10", ["which we can rearrange into a linear, first-order differential equation for 7(1\u2018):\n"]], ["block_11", [{"image_1": "437_1.png", "coords": [36, 187, 191, 225], "fig_type": "molecule"}]], ["block_12", ["J(t) 3%? 2%[1\u2014 exp(\u2014t/r)] Je[1\u2014 exp(\u2014t/1')]\n(11.2.12)\n"]], ["block_13", ["00 or + 515/\n(112.9)\n"]], ["block_14", ["7(1\u2018): 5g- [1 exp(\u2014t/1')]\n(11.2.11)\n"]], ["block_15", ["1\ny+\u2014y=22\n012m)\n\u20197\n7?\n"]], ["block_16", ["1\u2018\n0'0\n1\n70) yspring + 7dashpot \nE\" g\n0\n"]], ["block_17", ["=Q+Qi\nG\n77\n(1 1.2.7)\n"]], ["block_18", ["\u201870\n"]]], "page_438": [["block_0", [{"image_0": "438_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "438_1.png", "coords": [0, 51, 437, 180], "fig_type": "figure"}]], ["block_2", [{"image_2": "438_2.png", "coords": [27, 45, 241, 182], "fig_type": "figure"}]], ["block_3", ["One of the beautiful features of a linear system is that the response to a sinusoidal input is always a\nsinusoidal output at the same frequency; the amplitude and phase will generally differ between\ninput and output, however. (We have already encountered this linearity in Chapter 8; in light\nscattering, the scattered electric field had a different amplitude and phase from the incident wave,\nbut the frequency was the same.) It is also helpful to remember that sin wt and cos wt are the same\nwave, just phase-shifted by 90\u00b0 (or 17/2 rad): cos wt sin(wt + 17/2). Thus the strain rate in this\nexperiment, 77, is 90\u00b0 out-of\u2014phase with the strain. We can say that the stress in the elastic element\nis in-phase with the strain, and the stress in the viscous element is in-phase with the strain rate and\n90\u00b0 out\u2014of\u2014phase with the strain. This is a general result: in the linear response regime, the stress\ncan always be resolved into two components, one in-phase with the strain and one 90\u00b0 out-of-phase\nwith the strain. For a purely elastic solid the latter component would vanish, whereas for a purely\nviscous liquid the former component would be zero. A viscoelastic material is one for which both\ncomponents are significant.\nReturning now to the solution to Equation 11.2.14, the answer must be a wave with frequency\nw. The most general solution is\n"]], ["block_4", ["time-varying strain (or stress), e.g., y(t)= yo sin wt and measure the sinusoidally time-varying\nstress (or strain). One advantage of this approach, as we shall see, is that both the viscous and\nelastic character of the response can be resolved concurrently. Other advantages are technical; for\nexample, small sinusoidal signals can be extracted reliably through lock-in amplifier detection\nschemes. Also, because the magnitude of the modulus can vary by several orders of magnitude (see\nFigure 11.2), in a stress relaxation experiment the signal will become smaller and smaller as time\nevolves. In contrast, in the dynamic experiment at each new driving frequency w the strain\namplitude 70 can be adjusted to bring the stress signal into the conveniently measurable range\n(of course, taking care to remain in the linear viscoelastic regime).\nTo see how the Maxwell element responds to a strain of 70 sin wt, we adapt Equation 11.2.3:\n"]], ["block_5", ["Now the first-order, linear differential equation has a driving term proportional to cos wt:\n"]], ["block_6", ["Figure 11.7\nCreep compliance J(t) for (a) a viscoelastic liquid (the Maxwell model) and (b) a viscoelastic\nsolid (the Voigt model). A stress 00 is applied at time t 0 and in (a) the stress is removed at some later time t\u2019.\n"]], ["block_7", ["1\u201d\nTime\n0\nTime\n(a)\n(b)\nM\n"]], ["block_8", ["J( 1i)\nJ\u201c)\n"]], ["block_9", ["Response of the Maxwell and Voigt Elements\n427\n"]], ["block_10", ["0(t) A sin wt + B cos wt A0 sin(wt + (p)\n(11.2.15)\n"]], ["block_11", ["1\nx.\nc'r+\u20140=G'y0wcoswt\n(11.2.14)\n7\n"]], ["block_12", ["d\nd\n.\n3:\u201c = yea; Sln wt 2 yew cos wt\n_\n, + ,\n_ 1 .+1\n\u2014' 7e]\n7vis \n60-\n\ufb01g-\n(11.2.13)\n"]], ["block_13", ["O\n"]], ["block_14", ["___.__._____._.\n"]], ["block_15", ["__.___\n"]], ["block_16", ["I\nI\nI\nI\nI\nI\nI\nI\nl\nI\nI\nI\nl\n"]], ["block_17", ["W?\n"]], ["block_18", [{"image_3": "438_3.png", "coords": [237, 49, 441, 182], "fig_type": "figure"}]], ["block_19", ["_\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014_\u2014_\u2014_.\n"]]], "page_439": [["block_0", [{"image_0": "439_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["428\nLinear Viscoelasticity\n"]], ["block_2", [{"image_1": "439_1.png", "coords": [34, 235, 138, 267], "fig_type": "molecule"}]], ["block_3", ["recalling that d(sin wt)/dt co cos (or and d(cos wt)/dt \u2014w sin cut. The sine and cosine compon-\nents are independent of each other, so we can solve for the coefficients separately. For the sine part\n(recall that this represents the stress component in-phase with the strain)\n"]], ["block_4", ["These expressions for the coefficients A and B can be substituted into the expression for the\nmodulus:\n"]], ["block_5", ["where we note that any wave with frequency to can be written as a linear combination of a sine\nwave and a cosine wave, with two adjustable amplitudes A and B, or as a single sine wave, with\nadjustable amplitude A0 and phase qb. Let us insert the former version into the differential Equation\n"]], ["block_6", ["and for the cosine part (stress 90\u00b0 out-of\u2014phase with the strain)\n"]], ["block_7", ["This last relation defines the elastic or storage modulus, G\u2019, and the viscous or loss modulus, G\u201d:\n"]], ["block_8", ["The former measures the component of the stress response that is in-phase with the strain and the\nlatter the component in-phase with the strain rate. The relationships among the applied strain, the\nstress response, and the two components are illustrated in Figure 11.8. We can say that a material is\nviscoelastic if both G\u2019 and G\u201d are significant and we can anticipate that when 6\u2019 > G\u201d, the material\nis solid like, and when G\u201d > G\u2019, the material is liquid like.\nThe normalized dynamic moduli G\u2019/G and G\u201d/G for the Maxwell element are plotted versus\nreduced frequency car in Figure 11.9a in a double logarithmic format. These functions display the\nfollowing features. At low frequencies, arr<< 1, both G\u2019 and G\u201d increase with w. The former\nincreases as (02 and the latter as co and G\u201d > G\u2019 . This scaling with frequency is characteristic of all\nliquids when the frequency of deformation is much lower than the inverse of the longest relaxation\ntime of the material. Therefore, this is what one would expect to see for all polymer liquids once to\nis low enough. At high frequencies, (or >> 1, G\u2019 > G\u201d, and G\u2019 is independent of frequency whereas\nG\u201d falls as a)\u201c. This response is characteristic of a solid: the stress is independent of frequency (or\ntime), and in-phase with the strain. The two functions are equal, G\u2019 G\", and G\u201d shows \n"]], ["block_9", ["and thus\n"]], ["block_10", ["11.2.14, and work it through:\n"]], ["block_11", [{"image_2": "439_2.png", "coords": [38, 458, 215, 492], "fig_type": "molecule"}]], ["block_12", [{"image_3": "439_3.png", "coords": [38, 322, 128, 362], "fig_type": "molecule"}]], ["block_13", ["2 2\nt\nA\n.\nA\nw = Gila\u20142 smwt+ Giza\u20143 coswt\nyo\nl+wr\nl+wr\n= 0\u2019 sin cot + G\u201d cos wt\n(11.2.19)\n"]], ["block_14", ["1\n,\n1\nA\nAw cos cot Bw sin out + \u2014A srn wt + ;B\ncos wt = 67000 cos wt\n(11.2.16)\n7\n"]], ["block_15", ["A\n(027.2\nA \n\u2014\n11.2.\n670 1 + (0272\n(\n18b)\n"]], ["block_16", [{"image_4": "439_4.png", "coords": [46, 176, 122, 220], "fig_type": "molecule"}]], ["block_17", ["22\nA\n(or\nA\nG=G\u2014\u2014\u2014A\nw=G\u2014\u2014\u2014\u2014\n1+c02r2\n1+w2r2\n"]], ["block_18", ["1\n\u2014Bw+\u2014A=0\nT\nA=ww\n(Han)\n"]], ["block_19", ["1\nA\n"]], ["block_20", [{"image_5": "439_5.png", "coords": [79, 259, 157, 317], "fig_type": "molecule"}]], ["block_21", ["=we+lB\n(11.2.18a)\n7.\nA\n(or\nB \n\u2014\n070 l + (0272\n"]], ["block_22", ["(UT\n(11.2.20)\n"]]], "page_440": [["block_0", [{"image_0": "440_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["and\n"]], ["block_2", ["maximum when (or: 1. In other words, when the frequency of deformation is exactly the\nreciprocal of the relaxation time, the material is equally liquid-like and solid\u2014like. All of these\nfeatures of G\u2019 and G\" will be evident when we consider detailed molecular models for the\nviscoelastic response of polymer liquids.\n"]], ["block_3", ["where G* is the complex dynamic modulus and 0* is the complex stress (= 0'0 exp[i(wt + 5)]). The\nstorage modulus G\u2019 is the \u201creal\u201d part of G* because it is in-phase with the applied strain and G\u201d is\nthe \u201cimaginary\u201d part because it is 90\u00b0 out-of\u2014phase with the applied strain. The complex notation\nis simply a way to keep track of relative phases; there is nothing imaginary about the viscosity or\nthe modulus. One can choose to represent the modulus via its two components, G\u2019 and G\u201d, or in\nterms of the magnitude, Gm, and the phase angle, 5. These are interrelated by\n"]], ["block_4", ["It is common to use complex notation to describe the dynamic modulus, just as we did with light\nwaves in Chapter 8; complex numbers are reviewed in the Appendix. Thus, if we apply a\nsinusodial strain, 7* yo eXp(iwr), the re8ponse can be written as\n"]], ["block_5", ["Figure 11.8\nStrain y, strain rate 7, net Stress or, and stress components in-phase with the strain 0\" and\nin-phase with the strain rate 0'\u201d for a sinusoidally time-varying strain. The phase angle between the stress\nand the strain is 5.\n"]], ["block_6", ["Response of the Maxwell and Voigt Elements\n429\n"]], ["block_7", ["11.2.4\nDynamic Response: Complex Modulus and Complex Viscosity\n"]], ["block_8", [{"image_1": "440_1.png", "coords": [38, 588, 159, 623], "fig_type": "molecule"}]], ["block_9", [{"image_2": "440_2.png", "coords": [42, 474, 220, 507], "fig_type": "molecule"}]], ["block_10", ["H\ntan5 a\n(11.222)\n"]], ["block_11", ["(:-m {(G\u2019)2+(G\u201d)2}1/2\n"]], ["block_12", ["*\nG*(w) = 31? 6\u2019 +10\u201d Gm exp(i5)\n(11.2.21)\n"]], ["block_13", [{"image_3": "440_3.png", "coords": [76, 49, 391, 295], "fig_type": "figure"}]], ["block_14", ["we.\n/<\n"]], ["block_15", ["Illlil/\nIllltllll\n"]], ["block_16", ["lllllllll\n"]], ["block_17", ["IIII\n"]], ["block_18", ["llllrilllz\n"]], ["block_19", ["lllllllll\nIllll\n"]], ["block_20", ["R\n"]], ["block_21", ["q\n"]], ["block_22", ["I\nI\nl\n\u2018\n"]], ["block_23", ["Time\n"]], ["block_24", ["__\n"]], ["block_25", ["x\nI\n"]], ["block_26", ["\u2014\u2014\n"]], ["block_27", ["Illllllll\n"]]], "page_441": [["block_0", [{"image_0": "441_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "441_1.png", "coords": [32, 33, 228, 260], "fig_type": "figure"}]], ["block_2", ["430\nLinear Viscoelasticity\n"]], ["block_3", ["Figure 11.9\n(a) Normalized dynamic moduli G76 and C375, and (b) phase angle 8 and loss tangent tan 8,\nversus reduced frequency arr for the Maxwell element.\n"]], ["block_4", ["The phase angle, 5, is often discussed in terms of the so\u2014called loss tangent, tan 5, because it is the\ndirect ratio of the viscous and elastic parts. In other words, when the material is behaving like a\nliquid, tan 5 >> 1, and when it is a solid, tan 5 << 1; the crossover occurs at tan 5 as 1. In\nFigure 11.9b both tan 5 and the phase angle 5 itself are plotted for the Maxwell element.\nSometimes the viscoelastic response is presented in terms of a dynamic viscosity, 17*, rather than\nthe dynamic modulus:\n"]], ["block_5", ["The last issue we will take up before proceeding to detailed molecular models is that of the\ninterrelationships among 6\u2019, G\u201d, G(t), J2, and the steady\u2014\ufb02ow viscosity, 1). We begin with\nthe Boltzmann superposition principle of linear viscoelasticity, which asserts that the stress in the\nmaterial is the sum of the stress contributions from all strains applied in past times [3]. The\nlinearity of this principle lies in the fact that we can simply add up all the contributions because\nthe response of the material to any particular strain is linear and independent of whatever went on\nbefore or after. From the Maxwell model, we saw that G(t) was an exponentially decaying function\nof time. The strain was applied at one instant and the effect of the strain was evident in the stress at\nall subsequent times, but with an exponentially decaying amplitude. The more detailed models we\nwill consider subsequently all predict that G(t) is sum of such exponentials, with a sequence of\ndifferent relaxation times. In fact, the modulus is really a function of the time interval, G(t t\u2019),\n"]], ["block_6", ["Note that 7* :dy*/dt iw'yi\u2018, so that the components of G* and 11* are related by a factor of a):\n"]], ["block_7", ["Note also that there is a new phase angle, lb, in Equation 11.2.23; the relation between {i and 5 is\nleft to Problem 6. However, there is no new information content in 11* relative to 6*, or even in G*\nrelative to just discussing G\u2019 and G\u201d. Consequently the choice of format (i.e., G\u2019 and G\u201d; Gm and 5;\nn\u2019 and n\u201d) is mostly a matter of convenience or convention.\n"]], ["block_8", ["(a)\n(01'\n(b)\n031'\n"]], ["block_9", ["11.3\nBoltzmann Superposition Principle\n"]], ["block_10", ["-\nI\nl\n104\n_\n'\n\"\n6 (deg)\n_i 50\nI\nl-\nHr\n-\n100 -\no\n'\n_\n"]], ["block_11", ["101\nl llllllll\nIIIIIII\u2018II\n1 FlllITl\u2018I\u2014l ll\u2019lillll\ni llll\u2018l\u201c\n200 'm\nI Ilul'l'l'r\nI IIIIITI]\nl illlli] 100\n"]], ["block_12", ["100\n-\n.\n'.\n- 80\n150 -\nI.\n-\n"]], ["block_13", ["10\"2\n-\n(\u201c___\no\n'\n- 40\n"]], ["block_14", ["10\u20143\n"]], ["block_15", ["10\u20144\n"]], ["block_16", [{"image_2": "441_2.png", "coords": [41, 374, 194, 406], "fig_type": "molecule"}]], ["block_17", ["3* _ 0* _\nI\n-\nH\n__\n.\n\u201cn\n\u2014 F\n\u2014 \u201cn -In\n\u2014 \u2018nm GXPW)\n(11.2.23)\n"]], ["block_18", ["G* iw\u2018n\u2019k,\nG\u2019 am\u201d,\nG\" (011\u2019\n(11.2.24)\n"]], ["block_19", ["10\u20143\n'10\u20142\n'10\u20141\n'100\n\u201c[01\n'102\n10\u20143\n10\u20182\n\u201810\u201c1\n100\n101\n102\n"]], ["block_20", [{"image_3": "441_3.png", "coords": [52, 43, 427, 262], "fig_type": "figure"}]], ["block_21", ["_\n-\no\n'\n\u2014. 20\n.\nI\n.-\n-\n.\nI\n"]], ["block_22", [".-\n50 [-\n.\nI\n\"\n"]], ["block_23", [",\nI\nl\nI IIIHII\nI\nI ll\u201c.\n0\ni\nf [111M\nI\nLJ.\n..\n0\n"]], ["block_24", [{"image_4": "441_4.png", "coords": [236, 51, 428, 232], "fig_type": "figure"}]], ["block_25", ["F'\n..\n"]], ["block_26", ["..\nI\n..\n"]], ["block_27", [":\ntan 6 .\nI\nZ\n"]], ["block_28", ["r-\n.\nI\n-\n"]], ["block_29", ["[\u2014\n...\n-\n"]]], "page_442": [["block_0", [{"image_0": "442_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "442_1.png", "coords": [19, 263, 159, 309], "fig_type": "molecule"}]], ["block_2", [{"image_2": "442_2.png", "coords": [24, 98, 167, 135], "fig_type": "molecule"}]], ["block_3", [{"image_3": "442_3.png", "coords": [34, 149, 195, 200], "fig_type": "molecule"}]], ["block_4", ["This equation is an example of a constitutive equation, an equation that allows calculation of the\nstress in a material based on knowledge of all past deformations and of the relevant material\nresponse function. In the general case, 77(t 5), (3(5), and 0(t) will all be tensor quantities, but as\nnoted at the outset, we are restricting ourselves to simple shear and have tacitly chosen the relevant\ntensor elements.\nFrom Equation 11.3.2 we can develop some very useful interrelations as follows:\n"]], ["block_5", ["1.\nAssume we apply an instantaneous step strain at time t\u2019 =0, so that y 7080\"). The Dirac\ndelta function 5(t\u2019) is infinite at t\u2019 0, zero everywhere else, and integrates to 1 over all time,\nso that Equation 11.3.2 becomes 0(t) =G(t)yo. In this way we recover the stress relaxation\nmodulus as defined in Equation 11.2.6.\n2.\nIn steady \ufb02ow we apply a constant shear rate, 70 s) 'j/ and by substituting in Equation\n11.3.3 we obtain:\n"]], ["block_6", ["Therefore the stress now (time t) is obtained by adding the stress increments from past strain\nhistory (\u2019j/(t\u2019), with t\u2019 ranging from\n\u2014\u2014 00 to now). The stress increments are less and less important\nfor times further and further into the past because for a liquid G(t I\") always decays; we can say\nthat the material has a memory that fades with time. Equation 1 1.3.2 can be further transformed by\n"]], ["block_7", ["Thus the steady-\ufb02ow viscosity is just the integral over the entire stress relaxation modulus.\nThe viscosity represents the superposition of all modes of relaxation in the sample (i.e., all the\nexponential terms in G(t)), but because the integral is taken over all time, it is the slowest\ndecaying modes that will dominate the long time, steady-\ufb02ow response. Equation 11.3.4 was\nused to obtain the viscosity values in Figure 11.1 from the G(t) data in Figure 11.2.\n3.\nTo obtain the dynamic moduli, we apply a strain 7\u2019: yo sin wt, and recognize that\n"]], ["block_8", ["rather than the absolute time, t; in the previous analysis we had simply chosen t\u2019 0. Now we can\nexpress the Boltzmann superposition principle as follows. We first relate the increment in stress,\ndo, to an increment in strain, (17:\n"]], ["block_9", ["a simple change of variable: 3 t t\u2019, ds \u2014dt\u2019, and therefore\n"]], ["block_10", ["and thus\n"]], ["block_11", ["Boltzmann Superposition Principle\n431\n"]], ["block_12", [{"image_4": "442_4.png", "coords": [37, 105, 178, 123], "fig_type": "molecule"}]], ["block_13", [{"image_5": "442_5.png", "coords": [42, 456, 140, 566], "fig_type": "molecule"}]], ["block_14", ["0(t) J G(s)j1(t 5) ds\n(11.3.3)\n"]], ["block_15", ["0(t) Jda J C(t t\u2019)'i/(t')dt\u2019\n(11.3.2)\n"]], ["block_16", ["dazGdyzG-(gdtsdt\n(11.3.1)\n"]], ["block_17", ["70\u2018 s) you) cos (0(t s)\n= you) cos wtcos ms + you) sin wt sin cos\n(11-3-5)\n"]], ["block_18", [{"image_6": "442_6.png", "coords": [53, 470, 135, 510], "fig_type": "molecule"}]], ["block_19", ["0\n"]], ["block_20", ["G(t) dt\n(11.3.4)\n"]], ["block_21", ["1'\n"]]], "page_443": [["block_0", [{"image_0": "443_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "443_1.png", "coords": [29, 270, 203, 342], "fig_type": "figure"}]], ["block_2", ["The BSM model represents the chain as a linear string of N Hookean springs, or harmonic\noscillators, connecting N + 1 beads or mass points, as shown in Figure 11.10. (Recall that this\nmodel was introduced brie\ufb02y in Section 9.6, in the context of hydrodynamic interactions in dilute\nsolution. The difference between the Rouse and Zimm versions of the BSM is that the latter\nincorporates the hydrodynamic interactions, whereas the former does not.) The assembly is\n"]], ["block_3", ["The Maxwell and Voigt models provide useful insight into the nature of viscoelastic response, but\nare severely lacking in terms of providing a satisfying description of real polymer liquids. First, we\nwould like to have a molecular model in which the microscopic origins of viscosity and elasticity\nare more apparent. Second, a key feature of polymer viscoelasticity, whether in dilute solution or in\nconcentrated solutions and melts, is the wide range of relaxation processes that contribute to the\nmodulus. On an isolated polymer in a solvent, for example, the individual C\u2014C bonds along\nthe backbone can reorient in fractions of a nanosecond, whereas the entire end-to-end vector might\ntake microseconds or even milliseconds to undergo substantial reorientation. In a molten polymer,\nthe difference between the timescale for monomer motion and the timescale for the entire chain\nmotion can be much greater than in dilute solution. Consequently, it is essential that a useful\nmolecular model be able to capture this effect. In the next two sections we will examine the bead\u2014\nspring model (BSM) of Rouse [4] and Zimm [5]. The BSM is over 50 years old, but it still serves\nas the starting point for describing the viscoelastic character of all \ufb02exible polymers. First we\nwill describe the BSM itself, and then make some physical arguments for the character of the\nmain predictions. A fully detailed solution to the model is beyond the scope of this book.\n"]], ["block_4", ["432\nLinear Viscoelasticity\n"]], ["block_5", ["4.\nThe steady-state recoverable compliance must also be related to G0), although it takes a few\nlines of manipulation to show that this is the case. Consequently we will just provide the result\nhere:\n"]], ["block_6", ["These various relations underscore the fact that if you have access to 60\u2018), you can calculate all\nthe other linear viscoelastic functions.\n"]], ["block_7", [{"image_2": "443_2.png", "coords": [35, 279, 194, 334], "fig_type": "molecule"}]], ["block_8", ["11.4.1\nIngredients of the Bead\u2014Spring Model\n"]], ["block_9", ["11.4\nBead\u2014Spring Model\n"]], ["block_10", ["12 \u201827\u2014\u201c? 7\n5G(5)d5\n(11.3.8)\n[I G(5) d5]\n71\n0\n0\n"]], ["block_11", ["G\"(w) a)\nG(5) cos (05 d5\n(11.3.7)\n"]], ["block_12", ["0(2\u2018) 70a) cos (at J G(5) cos 00s + you) sin cot J G(5) sin (o5d5\n(11.3.6)\n"]], ["block_13", ["G\u2019(w) = 0:)\n0(5) sin (.05 d5\n"]], ["block_14", ["Comparing with Equation 11.2.20, we can express G\u2019 and G\u201d in terms of the sine and cosine\nFourier transforms of G(t), respectively:\n"]], ["block_15", ["Inserting this result into Equation 11.3.3, the stress can be written as the sum of two terms:\n"]], ["block_16", ["?5G(5)d5\n1\n00\n"]], ["block_17", ["can\u201458\n"]], ["block_18", ["can\u201458\n"]], ["block_19", ["00\n00\n"]], ["block_20", ["0\n0\n"]]], "page_444": [["block_0", [{"image_0": "444_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "444_1.png", "coords": [26, 24, 283, 181], "fig_type": "figure"}]], ["block_2", [{"image_2": "444_2.png", "coords": [29, 57, 164, 121], "fig_type": "molecule"}]], ["block_3", [{"image_3": "444_3.png", "coords": [32, 393, 227, 469], "fig_type": "molecule"}]], ["block_4", ["suspended in a viscous continuum, or solvent, with viscosity 7),. The actual mass of the polymer is\ndistributed uniformly among the beads. Each bead has a friction coefficient, g, when it moves\nthrough its surroundings. We could imagine using Stokes\u2019 Law (Equation 9.2.4) for this friction,\ni.e., g :67rnsa, where a is the radius of the bead, but in fact it is conventional to retain g as a\nparameter of the model rather than a. The springs have an rrns end-to-end length 1) and the entire\nchain is freely jointed at each bead\u2014spring unit. A single bead\u2014spring unit is meant to represent a\nGaussian subchain of the real polymer, that is, enough real backbone bonds or monomers that the\nend-to-end length of the subchain follows the Gaussian distribution (Equation 6.7.12). Thus this\nsubchain corresponds to a small number of persistence lengths.\nWe already have enough information to make some comments about this model:\n"]], ["block_5", ["where in this expression G is the free energy, S is the entropy, and P is the Gaussian probability\ndistribution applied to one subchain. The entropy S k In P, where the proportionality factor is\nincorporated into a constant that does not matter after we take the derivative with respect to x.\nThe last term in Equation 11.4.1 can be interpreted to mean that the subchain behaves as a\nHookean spring with a spring constant of 3kT/b2.\n2.\nThe freely jointed nature of the chain tells us that the radius of gyration will be Nb2/6 (see\nEquation 6.5.3) and therefore the model is designed for a theta solvent. (In principle one could\ninsert excluded volume interactions into the detailed solution of the model, but this can only be\ndone approximately, and at the price of great numerical complexity.)\n3.\nThe details of the actual chemical structure of the chain are subsumed into b and 4:. Thus, the\nBSM should make universal predictions for the character of the chain dynamics of \ufb02exible\nmolecules, but it cannot be used to say anything about particular local motions, bond rotations,\netc. of any actual polymer, which involve sections of the polymer less than a few persistence\nlengths.\n4.\nThe model has three parameters N, b, and J. N is proportional to the chain length, and therefore\nto M. In fact, this is a minimal set of parameters; we need to know how long the chain is, we\nneed to know some length scale characteristic of the polymer, and we need to know the\n"]], ["block_6", ["Bead\u2014Spring Model\n433\n"]], ["block_7", ["Figure 11.10\nThe bead\u2014spring model of Rouse [4] and Zimm [5].\n"]], ["block_8", ["1.\nThe viscous response in the BSM arises from the relative motion of the beads and solvent,\nwhereas the elastic response will come from the tendency of the Gaussian subchains to resist\ndeformation. We already discussed the entropic resistance to stretching a Gaussian chain in the\ncontext of chain swelling (see Section 7.7), and developed it in more detail in Chapter 10 (see\nEquation 10.5.7). The force, F, to stretch the subchain will be determined (in one dimension) from\n"]], ["block_9", [{"image_4": "444_4.png", "coords": [39, 396, 235, 464], "fig_type": "figure"}]], ["block_10", ["dG\nd\nd\nF=_:n_._\n=\u2014T\u2014k1P\ndx\n (has) \ndx \n"]], ["block_11", ["_ d\nt\nt\n3\n2\n#BkT\n__\ndx_\nconsan\u2014i\u2014gix\n\u2014~b\u20142*x\n(11.4.1)\n"]], ["block_12", [{"image_5": "444_5.png", "coords": [126, 42, 267, 179], "fig_type": "figure"}]], ["block_13", [{"image_6": "444_6.png", "coords": [126, 72, 272, 168], "fig_type": "molecule"}]], ["block_14", ["N+1\n"]]], "page_445": [["block_0", [{"image_0": "445_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Now that we have defined the model, we can anticipate some of its predictions. We have a system\nof N coupled, identical harmonic oscillators, and by analogy with the analysis of molecular\nvibrations and rotations used in infrared spectroscopy, we could think in terms of degrees of\nfreedom and normal coordinates. We have N +1 beads, so we need 3(N + 1) coordinates to\ncompletely specify the instantaneous conformation of the chain. (It turns out that we do not need\nto worry about the 3(N + 1) velocity or momentum values, however, because these equilibrate very\nrapidly due to the very high frequency of collisions between actual solvent molecules and\nmonomers.) We need three coordinates to specify the center of mass position and the remaining\n3N to describe the coordinates relative to the center of mass. That is, the translational diffusion of\nthe whole chain will involve the first three coordinates and the internal motions or conformational\nrelaxations the other 3N. Finally, we recognize a threefold degeneracy, namely that the x, y, and 2\npositions of any bead are uncorrelated; a force exerted on a particular spring in the x direction will\nnot in\ufb02uence the y or z coordinates of a connected bead. Thus, there are really only N distinct\ninternal degrees of freedom. If we choose the right coordinate system, the so-called normal\ncoordinates, we will find N normal modes, or characteristic relaxations, each with its own natural\nfrequency (or inverse relaxation time). Thus the main predictions of the BSM are the values of the\nN relaxation times and the character of the associated modes. The term normal here has the sense\nof orthogonal or linearly independent. Any particular motion of the chain in the laboratory\ncoordinate system can be decomposed into a linear combination of the N normal modes and the\nexcitation of any normal mode by some applied force will be dissipated by that mode only. (For\nthose readers with experience in quantum chemistry, there is a strong analogy between these\nnormal modes and the eigenvectors of the Hamiltonian in the Schrodinger equation; the frequen-\ncies of the BSM correspond to the allowed energy levels or eigenvalues of a quantum mechanical\nsystem, and the normal modes to the stationary states.)\nWe can illustrate the normal mode concept in Figure 11.11. We can draw a vector connecting\nany two beads separated by N/p springs where p is an integer between 1 and N (Figure 11.11a). We\ncan define the relaxation time for this vector in the following way. We take all the vectors\nconnecting the ends of (N/p)-long sections of the chain and do this for many chains. These vectors\nwould have a Gaussian distribution in terms of their length, with an rms value of b\\/ N/p (see\nEquation 6.3.2), and all orientations would be equally likely. Now we apply a step strain. The\ndistribution of orientations would be distorted in some way (Figure 11.11b), as would the distri-\nbution of lengths (Figure 11.11c). The overall distribution would relax back to equilibrium\nexponentially, with a characteristic time constant 7p. Thus the pth relaxation time, 7p, is the average\ntime it takes a section of chain containing N/p springs to recover from a disturbance. There are N\n"]], ["block_2", ["such relaxation times, varying from the first, 71, for the relaxation of the end-to-end vector of the\nentire chain, down to the Nth, m, for the relaxation of a single bead\u2014spring unit. Clearly\n"]], ["block_3", ["1'1 > 1'2 > 1'3 > > m. We have already identi\ufb01ed m as being proportional to 730g above. Note\nthat the roles played by 7N and 756g are essentially equivalent, but m is associated with a speci\ufb01c\nmodel, whereas 756g is a more general concept that can be discussed without reference to a particular\nmodel. The next issue is to find out how 1\",, depends on p, and in particular how 71 depends on N.\nWe can obtain these dependences from a rather simple argument in the case of the Rouse model\nwhere we neglect hydrodynamic interactions. Suppose for the sake of simplicity that the chain is\n"]], ["block_4", ["434\nLinear Viscoelasticity\n"]], ["block_5", ["timescale for motions on this length scale. This latter time, which we can call 733g, turns out to\nbe proportional to bit/k7\".\n5.\nIt is worth pointing out that there are N springs but N + 1 beads and, depending on the property\nin question, a literal solution of the model will involve either N or N + 1. However, we expect\nthe model to be most realistic when N is a large number, so this distinction will not matter; the\nimportant point is that N M.\n"]], ["block_6", ["11.4.2\nPredictions of the Bead-Spring Model\n"]]], "page_446": [["block_0", [{"image_0": "446_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "446_1.png", "coords": [4, 57, 342, 175], "fig_type": "figure"}]], ["block_2", [{"image_2": "446_2.png", "coords": [21, 180, 175, 290], "fig_type": "figure"}]], ["block_3", [{"image_3": "446_3.png", "coords": [29, 51, 200, 172], "fig_type": "figure"}]], ["block_4", ["motions, etc. it might just as well reverse \n"]], ["block_5", ["beads, thereby causing a disturbance. The two neighboring \n"]], ["block_6", ["where b represents the elementary step and 753g the \nthe disturbance. The factor of 2 re\ufb02ects that this is a \nthe average time to travel N/p springs, 7p, will be\n"]], ["block_7", ["and because in fact the beads are constantly undergoing \n"]], ["block_8", ["solvent, we can actually think of the disturbance \n"]], ["block_9", ["element, because in this case no spring can be \n"]], ["block_10", ["Section 9.5. This \u201cinternal\u201d diffusion constant, Dim, \n"]], ["block_11", ["disturbance to propagate down N/p springs? (Note \n"]], ["block_12", ["deformation is not transmitted instantaneously along \n"]], ["block_13", ["exerting forces on the next beads and so on down \n"]], ["block_14", ["distribution of lengths in (c) relax together to their equilibrium \n"]], ["block_15", [{"image_4": "446_4.png", "coords": [33, 634, 183, 679], "fig_type": "molecule"}]], ["block_16", ["lying along the x-axis, with all springs at the rest length \n"]], ["block_17", ["containing N/p\n"]], ["block_18", ["Figure 11.11\nIllustration of the relaxation process for\n"]], ["block_19", ["435\nBead\u2014Spring Model\n"]], ["block_20", ["P(\nN,\nh)\n"]], ["block_21", ["2\n2\n\u00abr, = (if b)\n1\n=\n(5\u2019) \n(11.4.3)\n"]], ["block_22", ["(b2) 2Dint'Tseg\n(11.4.2)\n"]], ["block_23", ["19\n2Dim\np\n"]], ["block_24", ["(0) Distribution \nimmediately after \n"]], ["block_25", ["and at equilibrium \n"]], ["block_26", ["dimensions immediately \n"]], ["block_27", ["(b) Distribution \n"]], ["block_28", ["(dashed) and \n"]], ["block_29", ["5\u2014spring units of \n"]], ["block_30", ["(a) Vectors connecting\n"]]], "page_447": [["block_0", [{"image_0": "447_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "447_1.png", "coords": [27, 448, 160, 493], "fig_type": "molecule"}]], ["block_2", [{"image_2": "447_2.png", "coords": [29, 163, 212, 213], "fig_type": "molecule"}]], ["block_3", [{"image_3": "447_3.png", "coords": [30, 579, 167, 628], "fig_type": "molecule"}]], ["block_4", ["From Equation 1.1.4.3 we can see that the longest relaxation time, 7'], will just be Nzrseg, and the\npth relaxation time will be 71/192. As the local segmental time must be independent of total\nmolecular weight (because a few monomers can rearrange themselves without disturbing the entire\nchain), Tseg does not depend on N. Therefore the Rouse version of the BSM model predicts that the\nlongest relaxation time of the chain will be proportional to M2. Furthermore, the viscoelastic\nresponse will be governed by N different relaxation modes, with relaxation times spaced as 1:1/4:1/\n9: :1/N2.\nThis last result, which we obtained by a rather qualitative argument, is not the exact solution to\nthis model. The correct result is\n"]], ["block_5", ["The last term in Equation 11.4.4 employs the fact that sin x as x for small x. Thus the simple\nscaling of 7}, with (N/p)2 in Equation 11.4.3 is a very good approximation for small p, i.e., for the\nmodes that involve big sections of the chain. It is these modes that we hope to describe well by the\nBSM; the modes for very short pieces of a real chain should depend more on local structural details\nand are therefore less likely to be captured by this approach. Comparing Equation 11.4.4 and\nEquation 11.4.3 we can see that the numerical prefactor for Tseg can be specified:\n"]], ["block_6", ["We have inserted a front factor, GP, which gives the amplitude of each mode; Gp must have the\nunits of a modulus. We now invoke the equipartition theorem of statistical mechanics: each degree\nof freedom, or normal mode, acquires kT of thermal energy. Furthermore, we can assume that the\nmodulus will increase linearly with the number of chains per unit volume because each chain can\nstore the same amount of elastic energy under deformation. The number of chains per unit volume\nis cNav/M, where c is the concentration in g/mL. Therefore we can equate Gp with (cNav/M)\nH\" CRT/M:\n"]], ["block_7", ["You should check that CRT/M does have the units of modulus and compare this quantity to the\nmodulus of the ideal elastomer, pRT/Mx, from Equation 10.5.17. Equation 1 1.4.7 is, in fact, the result\nthat is obtained by a full solution of the model, but that would require a great deal more legwork.\n"]], ["block_8", ["Again, however, it should be emphasized that the BSM is expected to describe the longer range\nmotions of the chain, not the local details, so the numerical prefactor in rseg should not be taken too\nseriously. It is also worthwhile to confirm that the collection of quantities on the right-hand side of\nEquation 11.4.5 gives units of time. The cgs units for g are g/s (recall the beginning of Chapter 9)\nand b is in cm, while kT has units of energy, or g cm2/s2. The net result is therefore (g/s)(cm2)/\n(g cmz/sz) s.\nNow that we have a good representation of the relaxation times of the model, we need to see\nhow they affect G(t), and therefore G*(co) and 17. First, we expect that each mode will contribute an\nexponential decay to G(t), rather like the single mode of the Maxwell model. Because we are\ndealing with normal modes, they are independent of one another, so we just sum them:\n"]], ["block_9", ["436\nLinear Viscoelasticity\n"]], ["block_10", [{"image_4": "447_4.png", "coords": [39, 458, 158, 482], "fig_type": "molecule"}]], ["block_11", [{"image_5": "447_5.png", "coords": [43, 281, 111, 328], "fig_type": "molecule"}]], ["block_12", [",\nN\nG(t) %\nexp(\u2014t/Tp)\n(11.4.7)\n"]], ["block_13", ["N\nG(t) Z GP exp(\u2014t/'rp)\n(11.4.6)\n"]], ["block_14", ["gbl\n1\ngal\nN2\nT : \ng \n\u20180\n6kT 4 sin2 ( par/2N)\n6kT 11's\n(11.4.4)\n"]], ["block_15", ["()2;\n736g 6772\u201d\n(1145)\n"]], ["block_16", ["p=1\n"]], ["block_17", ["p=l\n"]]], "page_448": [["block_0", [{"image_0": "448_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "448_1.png", "coords": [29, 401, 159, 450], "fig_type": "molecule"}]], ["block_2", [{"image_2": "448_2.png", "coords": [31, 556, 206, 646], "fig_type": "figure"}]], ["block_3", [{"image_3": "448_3.png", "coords": [34, 558, 190, 639], "fig_type": "molecule"}]], ["block_4", ["This function may be compared to G(t) for the Maxwell element shown in Figure 11.6. There are\nnow three regimes of behavior: short times, t < TN, where little relaxation occurs; t> r], where the\nstress completely relaxes; TN < t < 71, where each mode contributes. It is this intermediate regime\nthat is new and which is a direct consequence of the multiple relaxation times. Because 71 NQTN,\nthe width of this regime on the time axis should grow as M2. Consequently we have succeeded in\nachieving the two goals identified at the start of this section: we have a microscopic model that\nexhibits a broad range of relaxation processes.\nThe dynamic shear moduli can be obtained from G(t) via Equation 1 1.3.7 (after looking up the\nintegrals of the form [cos (03 exp (\u2014s/rp) ds):\n"]], ["block_5", ["We plot this modulus in Figure 11.12, as MG(I)/CRT versus [/7]. Recalling that rp/rl 1/p2,\nwe have\n"]], ["block_6", ["Figure 11.12\nStress relaxation modulus G(t) for the Rouse model with N 10 beads. The contributions of\nthe nine individual modes as a function of reduced time {/71 are indicated by dashed lines.\n"]], ["block_7", ["where we note that the loss modulus includes an additive contribution from a purely viscous\nsolvent. We could plot these functions as they stand, once values for N and 756g (or b and g) are\n"]], ["block_8", ["Bead\u2014Spring Model\n437\n"]], ["block_9", ["G(t)/CHT\n"]], ["block_10", ["M\n"]], ["block_11", [{"image_4": "448_4.png", "coords": [42, 562, 163, 607], "fig_type": "molecule"}]], ["block_12", [{"image_5": "448_5.png", "coords": [44, 32, 329, 313], "fig_type": "figure"}]], ["block_13", ["M\nN\n2\ngig\u20147:00) Zexp(\u2014i\u2014t)\n(11.4.8)\np=l\n1\n"]], ["block_14", ["RT\nN\nG\u201d (0715 +C_ 2\n"]], ["block_15", [{"image_6": "448_6.png", "coords": [49, 601, 186, 634], "fig_type": "molecule"}]], ["block_16", ["GI\n"]], ["block_17", ["10-1\n"]], ["block_18", ["_cRT\nN\n(mTp)2\nM\n17:] \n"]], ["block_19", ["1 0\u20144\n1 or3\n1 0-2\n1 0-1\n1 00\n"]], ["block_20", ["COT}?\nM\np=l\n1 + (wrp)2\n"]], ["block_21", ["\u201dT1\n"]], ["block_22", ["(11.4.9)\n"]]], "page_449": [["block_0", [{"image_0": "449_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "449_1.png", "coords": [34, 82, 153, 144], "fig_type": "molecule"}]], ["block_2", ["433\nLinear Viscoelasticity\n"]], ["block_3", ["Finally, we can normalize the frequency axis by the longest relaxation time, recalling that\n(rp/71)=1/p2. Thus we arrive at a functional form that depends only on the dimensionless\nvariables N and (071:\n"]], ["block_4", ["speci\ufb01ed. However, we can also form intrinsic functions (by analogy to the intrinsic viscosity in\nSection 9.3) as follows:\n"]], ["block_5", ["and then reduced functions as\n"]], ["block_6", ["These functions are plotted in Figure 11.13 and the results can be compared to those for the\nMaxwell model in Figure 11.9a. At low frequencies, am < 1, the liquid-like response is the same:\nG\u2019 (02, G\u201d w, and G\" > G\u2019. At very high frequencies, (O\u2019TN > 1, the material is a solid and again\n"]], ["block_7", ["Figure 11.13\nReduced intrinsic dynamic moduli [G\u2019]R and [G\u201d]R versus reduced frequency am for \nRouse model for the same lO\u2014bead chain as in Figure 11.12. The limiting slopes are also indicated.\n"]], ["block_8", [{"image_2": "449_2.png", "coords": [36, 164, 196, 206], "fig_type": "molecule"}]], ["block_9", [{"image_3": "449_3.png", "coords": [39, 268, 264, 349], "fig_type": "molecule"}]], ["block_10", [{"image_4": "449_4.png", "coords": [43, 392, 264, 625], "fig_type": "figure"}]], ["block_11", [{"image_5": "449_5.png", "coords": [47, 146, 194, 241], "fig_type": "molecule"}]], ["block_12", ["(1 1.4.10)\n[0\u201d] :1im (G\n M\u201c) \nc\u2014rO\nC\n"]], ["block_13", ["[G\ufb02in*\n]\"Z\ufb02_\u2014_1+(:J%\n"]], ["block_14", ["M\nN\n(an )2\nC}!\nC}!\n__\np\n[\nL?\n"]], ["block_15", ["[G!]R=i\n(w71)2(Tp/T1)2\n2%\n(\u00a30702\n"]], ["block_16", ["101\nI\n1 III\u201c!\nI\ni IIHIq\nI\nIIIIIHL\u2018a\ufb02\nI\nIllHL\n"]], ["block_17", ["0_2\n"]], ["block_18", ["10-1\n1o0\n101\n102\n103\n"]], ["block_19", ["_.\n------- Gr\n"]], ["block_20", ["ERMH\n]\n"]], ["block_21", ["p=1 1+(071)2(7p/71)2\np=1 P4 +0011)2\n(11.4.12)\nN\nN\n2\n_\n(wTi)(7p/Tl)\n__\n(6071)}?\n\u2014\u2014 Z,\nZ\n\u2018|\u2018(w\u201971)2(\u2019Tp/\u2019T1)2\n\u2014\nP4 +0071)2\np21\np=1\n"]], ["block_22", ["I\nllllllll\nl\nI\nlllllll\nl\nllllllll\nl\nI\nLIIIIl\n"]], ["block_23", ["N=10\nI\n"]], ["block_24", ["G;1+(m\u2014P)2\n"]], ["block_25", ["(11.4.11)\nN\n"]], ["block_26", ["m\u20ac1\n"]]], "page_450": [["block_0", [{"image_0": "450_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "450_1.png", "coords": [33, 257, 189, 293], "fig_type": "molecule"}]], ["block_2", ["As we have just seen, the Rouse version of the BSM predicts that [7}] NM and Dt l/M, whereas\nwe saw in Chapter 9 that experimentally [n] M3\"\u20141 and DI M \u2019V, with v = 0.5 in a theta solvent\nand 0.6 in a good solvent. The principal reason for this discrepancy is the phenomenon of\nhydrodynamic interactions. In the context of the BSM, this means that the motion of any one\nbead through the solvent perturbs the \ufb02uid \ufb02ow at the position of every other bead. This effect was\nincorporated into the Rouse model by Zimm, utilizing the approach of Kirkwood and Riseman\nintroduced in Section 9.7. As described there, two simplifying assumptions were necessary to make\nthe solution tractable. First, the description of the hydrodynamic interaction was truncated at the\nleading term (the so-called Oseen tensor), where the effect of bead j on bead k falls off as l/rjk.\nSecond, the instantaneous 1/r\ufb02c was replaced by its equilibrium average, (1/ wk)using the Gaussian\ndistribution; this is called the pre-averaging approximation. After these assumptions it proves\npossible to transform to a set of normal coordinates, just as in the Rouse model, and extract a set of\nN relaxation times, 7p, and associated normal modes. In contrast to the Rouse model, there is no\nanalytical solution for the normal modes and the relaxation times; the exact solution can only be\nobtained numerically. (However, efficient algorithms have been developed to do this [6].)\nThe predictions of the Zimm model are identical to those of the Rouse model in terms of the\nform of the equations (e.g., Equation 11.4.7 and Equation 11.4.9 for G(t), G\u2019, and G\u201d), but differ\n"]], ["block_3", [{"image_2": "450_2.png", "coords": [35, 202, 200, 238], "fig_type": "molecule"}]], ["block_4", ["As we demonstrated above, 1'1 M2, so the Rouse model predicts that [n] M. We saw in Section\n9.3 that this is not experimentally correct. This is, in fact, the clear evidence that the Rouse model\nis missing an important ingredient in solution dynamics, namely the hydrodynamic interactions\ndescribed in Section 9.7. As a final note, we can state that the tracer diffusion coefficient of the\nRouse model is given by\n"]], ["block_5", ["which is what we termed thefreely draining result in Section 9.7. This expression has N + 1 in place\nof N because there are N + 1 beads, but this is of no real consequence as pointed out previously.\n"]], ["block_6", ["The last transformation was taken because\n"]], ["block_7", ["where we have inserted the known result for this infinite sum. Thus\n"]], ["block_8", ["the response is the same: G\u2019 coo, G\u201d (0\u2018 1, and G\u2019 > G\u201d. As with G(t), it is in the intermediate\nregion that the results differ. Now we have N relaxations rather than the single mode of the\nMaxwell element and so G\u2019 and G\" evolve slowly with a) for 1/1'1 < w < \u201dTN; 6\u2019 and G\" both\nincrease approximately as (01/2 (although this is not evident here as N is small).\nThe Rouse model also makes an explicit prediction for the intrinsic viscosity,\n[7}] (see\nSection 9.3). We need to take the solution loss modulus, G\u201d (Equation 11.4.9), use the relation\nbetween G\u201d and 1) (Equation 11.2.24), and take the dual limits of w\u2014> 0 and c\u2014> 0.\n"]], ["block_9", ["11.5\nZimm Model for Dilute Solutions, Rouse Model\nfor Unentangled Melts\n"]], ["block_10", [{"image_3": "450_3.png", "coords": [36, 355, 118, 393], "fig_type": "molecule"}]], ["block_11", ["Zimm Model for Dilute Solutions, Rouse Model for Unentangled Melts\n439\n"]], ["block_12", [{"image_4": "450_4.png", "coords": [38, 140, 301, 185], "fig_type": "molecule"}]], ["block_13", ["[J\nD \n11.4.16\n\u2018\n(N + 1);\n(\n)\n"]], ["block_14", [":31: \n$3.\na, N _. 00\n(11.4.14)\n1'1\n6\n"]], ["block_15", ["1) n5\nRT\u2019TI\u2019JTZ\n_ 1\n__\n2\n11.4.1\n[n]\n6133 (\n11.6 )\nM11156\n(\n5)\n"]], ["block_16", ["G__\u2019:\n"]], ["block_17", ["+c_RT\n+cRT\n"]], ["block_18", [{"image_5": "450_5.png", "coords": [151, 146, 284, 186], "fig_type": "molecule"}]], ["block_19", ["N\n7,,\n"]]], "page_451": [["block_0", [{"image_0": "451_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "451_1.png", "coords": [28, 545, 187, 589], "fig_type": "molecule"}]], ["block_2", ["which was also a prediction for the Rouse model (Equation 11.4.15).\nWe need to be a little careful about something here. The M dependences of [1)], D,, and now 1'1\nare all based on the M dependence of Rg because hydrodynamic interactions are sufficiently strong\nto make the dynamic behavior of the entire chain equivalent to that of a hard sphere. The effect of\nsolvent quality is accounted for through the appropriate value of V and thus this argument works\nequally well for theta solvents and good solvents. However, the Zimm model for the full\nviscoelastic spectrum is only strictly valid for theta chains because (a) the model is freely jointed\nand has no self\u2014avoidance terms and (b) the hydrodynamic interaction is incorporated via pre-\naveraging over a Gaussian distribution. This may seem a little disappointing in the sense that we\nhave simple physical arguments that give the correct M dependences of the global chain dynamics\nin any solvent ([17], D,, and 7'1), but the model for the internal chain dynamics (G\u2019, G\u201d, and 7,, for\np > 1) can only be applied to theta solvents. However, it turns out that there are various approxi-\nmate ways that the effects of varying solvent quality can be incorporated into the Zimm model and\nseveral are sufficiently accurate to describe experimental data very well. We will brie\ufb02y describe\nthe most physically transparent approach known as dynamic scaling; it is in the spirit of the \u201ca coil\nbehaves hydrodynamically like a hard sphere\u201d argument.\nThe algebra is very simple. Returning to Equation 11.5.2, we have\n"]], ["block_3", ["only in the relative values of the relaxation times. Qualitatively, the hydrodynamic interaction\naccelerates the relaxation of the chain because each bead communicates with every other directly\nthrough the solvent, rather than just by the springs. To put it another way, the relaxation times\nbecome more closely Spaced than 1/p2. We saw in Chapter 9 that the diffusion constant of a single\nchain and the intrinsic viscosity were well described by picturing the coil as a sphere with a radius\nproportional to Rg; the same approach is successful here. A relation that plays the same role for\nrotational friction as Stokes\u2019 law does for translational friction (Equation 9.2.4) is called the\nStokes\u2014Einstein\u2014Debye equation. It gives the rotational time for a sphere of radius R as\n"]], ["block_4", ["Thus, in a theta solvent, the longest relaxation time should scale as M\n, be\nexactly the prediction of the Zimm model. We also can see that the longest relaxation time (NRE)\nand the intrinsic viscosity (\u201cIRE/M) are again intimately related:\n"]], ["block_5", ["where again we invoke g\u201c bns. By analogy with Equation 11.4.3, we expect that the relaxation\ntime for a subsection of the chain containing N/p units should scale as (N/p)3\", and thus we assert\nthat any relaxation time can be written\n"]], ["block_6", ["440\nLinear Viscoelasticity\n"]], ["block_7", [{"image_2": "451_2.png", "coords": [35, 630, 123, 679], "fig_type": "molecule"}]], ["block_8", ["We present this result without derivation, but we can observe that it has the same structure as 736g\nin the BSM (Equation 11.4.5), if we recognize that f bm. On this basis we may propose that\n"]], ["block_9", [{"image_3": "451_3.png", "coords": [46, 226, 129, 259], "fig_type": "molecule"}]], ["block_10", ["N\n311\n7\nN\n_\n786\n(11.5.5)\np\n(p)\ng\n"]], ["block_11", ["71 \nRT\n(11.5.3)\n"]], ["block_12", ["Trot \nkT\n(11.51)\n"]], ["block_13", ["3\nRgns\nM31)\";\n71\nrv N \nkT\nkT\n(11.5.2)\n"]], ["block_14", ["Ran-9 N N31) {93$\n[CT\n[CT \n(11.5.4)\nTlm\n"]]], "page_452": [["block_0", [{"image_0": "452_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "452_1.png", "coords": [33, 125, 179, 214], "fig_type": "molecule"}]], ["block_2", ["These expressions are straightforward to evaluate and give predictions for the moduli that are\nessentially indistinguishable from the full numerical evaluation of the Zimm model.\nComparisons of the Zimm model to the viscoelastic properties of \ufb02exible chains are presented\nin Figure 11.14. In these experiments, five high molecular weight 1,4-polybutadiene chains were\ndissolved in diethylhexyl phthalate, which is a theta solvent at 18.0\u00b0C. Measurements were made at\na series of dilute concentrations, and extrapolated to infinite dilution to obtain the intrinsic moduli.\n(The actual measurements were made on a \ufb02ow birefringence apparatus, which determines an\noptical anisotrOpy that is directly proportional to the stress; the Optical experiment is considerably\nmore sensitive than most rheometers, which is important for very dilute solutions where the signals\nare small.) The data are plotted as reduced intrinsic moduli versus reduced frequency, am, and\nthus theory is compared with the data with no adjustable parameters. In Figure 11.14a, the\ntemperature corresponds to the theta point and the fit is to the dynamic scaling result with\nv=0.50, or to the full numerical solution to the Zimm theory; the two curves are identical. In\nFigure 11.14b, the temperature has increased to 30\u00b0C and now there is some degree of excluded\nvolume. The data for G\u201d for the highest M polymer are well described by dynamic scaling with\nv m 0.52-0.53, and furthermore the longest relaxation times used to collapse the data, shown in\nFigure 11.15, have the expected dependence on M3\u201d. Based on many results such as these, we may\nconclude that the Zimm model, modified for the effects of excluded volume as needed, provides a\nquantitative description of the dynamics of isolated, long, \ufb02exible chains.\nFrom the success of the Zimm model, one might be led to conclude that the Rouse model serves\nonly as a conceptual foundation on which to build more elaborate descriptions of solution\nviscoelasticity and that it is not so useful in terms of actual data. However, it turns out that the\nRouse model is very successful in describing the dynamics of low molecular weight polymer melts\nor\nconcentrated\nsolutions.\nThe\nreason\nis\ntwofold.\nFirst,\nat very\nhigh\nconcentrations,\nthe\nhydrodynamic interactions are effectively screened. Any motion of one monomer is transmitted\nby many other monomers before it reaches another monomer on the same chain and thus it imparts\nno through-space coherence to the chain motion. Second, the chains are very nearly Gaussian in the\nmelt, as discussed in Section 7.7, so the freely jointed assumption is appropriate. In the melt there\nis no solvent, so the effect of frictional resistance to motion on the segmental level is subsumed\ninto the friction factor, 5. The Rouse model predicts that nNM, 1'1 ~M2, and DNM \n1, all of\nwhich are at least approximately true for low molecular weight polymers in the melt. (They are not\nexactly true in most cases, because the subchain friction factor ; develops an M dependence at low\nM, due primarily to the M dependence of the glass transition temperature. This effect will be\ndiscussed in Chapter 12, but it can be corrected for, and then the Rouse predictions hold rather\nwell.) We have been careful to emphasize low molecular weight melts here because for higher\nmolecular weights a new effect comes into play and the Rouse model is again inadequate. The new\neffect is called entanglement and we describe it in the next section. To conclude this section, we\nsummarize in Table 11.1 the main predictions for the M dependence of chain dynamics from the\n"]], ["block_3", ["This can be substituted directly into the expressions for the reduced intrinsic dynamic moduli,\nEquation 11.4.12, and thus\n"]], ["block_4", ["Zimm Model for Dilute Solutions, Rouse Model for Unentangled Melts\n441\n"]], ["block_5", [{"image_2": "452_2.png", "coords": [44, 132, 182, 181], "fig_type": "molecule"}]], ["block_6", [":3 p\u20193\"\n(11.5.6)\n"]], ["block_7", ["[GUJR\n"]], ["block_8", ["[01]}?\nZ\n((071)09\n)\n"]], ["block_9", ["NZ\nW\nFl \n"]], ["block_10", ["p=1 \n"]], ["block_11", ["N\n2\n\u20143V 2\n"]], ["block_12", ["(11.5.7)\n"]]], "page_453": [["block_0", [{"image_0": "453_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "453_1.png", "coords": [31, 51, 280, 556], "fig_type": "figure"}]], ["block_2", ["Figure 11.14\nExperimentally obtained reduced intrinsic moduli for five polybutadiene samples in dioctyl\nphthalate, compared to the predictions of the Zimm model. (a) At the theta temperature, 18\u00b0C, with v0.50.\n(b) The data for G\u201d for the highest molecular weight sample only, slightly above the theta temperature, at\n30\u00b0C. The curves represent the Zimm theory with three different scaling exponents, v, and the Rouse theory.\nThe data have been shifted vertically by a factor a, but it is clear that the experimental slope agrees with\nexpectation for modest excluded volume (v > 0.50). (Data obtained from Sahouani, H. and Lodge, T.P.,\nMacromolecules, 25, 5632, 1992. With permission.)\n"]], ["block_3", ["442\nLinear Viscoelasticity\n"]], ["block_4", [{"image_2": "453_2.png", "coords": [35, 58, 304, 302], "fig_type": "figure"}]], ["block_5", ["IE\n"]], ["block_6", ["='\u2014'\n10\u20146\n3\n\u20185\n"]], ["block_7", ["(2.\n"]], ["block_8", [".35\nI\ng\n_\n"]], ["block_9", ["ts\n-\nV2050\n"]], ["block_10", [{"image_3": "453_3.png", "coords": [46, 307, 294, 537], "fig_type": "figure"}]], ["block_11", ["10\u20144\n\u2014l\u2014llllllll\nl\nllllllll\nI\nllllTn'l\nI\nIlllll\u2014I'l\u2014\n1\n\u2018l\u2014IIIIIL\n"]], ["block_12", ["10_9\n"]], ["block_13", ["10\u20143\n\"\nDynamic scaling, :2 0.50\n__\n"]], ["block_14", ["10-5\n5\n"]], ["block_15", ["10'? 5\nE\n"]], ["block_16", ["101\n4.\n\u2018\nJ\n"]], ["block_17", ["100\n\u2014 \u2014 -\nV 0.53\n\u2014:\n"]], ["block_18", ["2\n\u2014l_\nI\u2018l\u2014|_FIII\n\u2014l_\nI\nI_l_l_lll\n\u2014l\u2014l\nIl_l_l_ll\n\u2014l'\nITI\n10\nI\nI\nl\n/\na\nPolybutadiene in DOP\nJ\n"]], ["block_19", ["10._1\nI\nI\nl_[ll\nII\nI\nI\n1\n1\u2014K\nIIl\n_l\nI Ill\nIII\nI\nI\n11\n10-1\n100\n101\n1o2\n103\n"]], ["block_20", ["(a)\nmi\n"]], ["block_21", ["I\nl\nI_IIIIIII\nI\nIIIIIIII\nI\nI IIIIIII\nJ_I_llL\n10-2\n10\u201c\n10\u00b0\n101\n102\n103\n"]], ["block_22", [{"image_4": "453_4.png", "coords": [69, 414, 291, 527], "fig_type": "figure"}]], ["block_23", ["-\n"]], ["block_24", ["\u00e9?\n------- Rouse\n_\n"]], ["block_25", ["Polybutadiene in DOP\nT =18.0\u00b0C (t9)\n"]], ["block_26", ["M: 1.6 x107\n\", o\u2019 ft\n"]], ["block_27", ["a\n.x\n\"l\nT: 300\u00b0C\n'\n"]], ["block_28", ["Zimm theory\n"]], ["block_29", [{"image_5": "453_5.png", "coords": [190, 424, 271, 510], "fig_type": "figure"}]], ["block_30", ["1\n-----\nV 0.59\n\u2018\n"]]], "page_454": [["block_0", [{"image_0": "454_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "454_1.png", "coords": [23, 561, 347, 646], "fig_type": "figure"}]], ["block_2", ["There are several different ways to proceed. For example, from Figure 9.10a we find that\nDt m 10\u2014 cmZ/s, and therefore we can calculate the hydrodynamic radius by the Stokes\u2014Einstein\nequation (Equation 9.5.5) (the viscosity of cyclohexane is about 0.8 cP at 35\u00b0C):\n"]], ["block_3", ["_\nH\nN\n\u20141\n__\nk?\u201c\nN\n_,,\n__ENN\n\u20142\nDI\nDt\n\u2014- m\nM\nD[ \n61T\ufb025Rh\nM\nDt\n\u2014\nN; \nM\n"]], ["block_4", ["Table 1 1.1\nPredictions of Three Models for the Molecular Weight Dependence\nof Various Chain Dynamics Quantities\n"]], ["block_5", ["two forms of the BSM (and also the related predictions of the reptation model to be presented in\nSection 11.7), and give an example of the application of the BSM.\n"]], ["block_6", ["Use the BSM to estimate the longest relaxation time and the segmental relaxation time for polystyrene\nwith M =106 g/mol in cyclohexane at the theta temperature (35\u00b0C). Recall that values of Dt and [n]\nwere given as a function of M for this system in Figure 9.10a and Figure 9.5, respectively.\n"]], ["block_7", ["Pr0perty\nRouse\nZimm\nReptation\n"]], ["block_8", ["Zimm Model for Dilute Solutions, Rouse Model for Unentangled Melts\n443\n"]], ["block_9", ["Figure 1 1.1 5\nInfinite dilution longest relaxation time versus molecular weight at the theta temperature\n(18\u00b0C) and at 30\u00b0C., for the same solutions as in Figure 11.14. The exponents are obtained from the fitted slopes\n(= 31\u201d). (Data obtained from Sahouani, H., and Lodge, T.P., Macromolecules, 25, 5632, 1992. With permission.)\n"]], ["block_10", ["Solution\n"]], ["block_11", ["3\n71\nT1 Nseg M2\nT] NBPTSQg MBV\nT1 N \u00a7ETseg N M3\n"]], ["block_12", ["Example 11.2\n"]], ["block_13", ["n\n\u2014\n\u2014-\nn=\u00a7TIGN~M3\n"]], ["block_14", ["RT\nRT\n_\nin]\n[MAJ-miLNM\n[n]~m\u201dj}~M3\u201d1\n\u2014\n"]], ["block_15", ["\u20199,\u201c 10\u20182?\n_:\nc:\n:\n3\n"]], ["block_16", [{"image_2": "454_2.png", "coords": [39, 45, 276, 272], "fig_type": "figure"}]], ["block_17", ["1.4 X 10'16 x 308\n6x3.14><0.008><10\u20147\n3X\n0\ncm\n30nm\nRh\n"]], ["block_18", ["10\u20143?\n'3\n"]], ["block_19", ["10_1_\nV=0.50\n_\n"]], ["block_20", ["10-4\nI\nIIIIII\u2018\nI\nIIIILIII\n_|\n_I_\nI_lllli\n105\n106\n107\n103\nM (g/mol)\n"]], ["block_21", ["100;\n"]], ["block_22", ["E\n+3000\ni\n:\nV=0.52\n:\n"]], ["block_23", ["'\n+18\u00b0C(6)\nI\n"]], ["block_24", [{"image_3": "454_3.png", "coords": [67, 80, 147, 143], "fig_type": "figure"}]], ["block_25", ["I\n||llllll\nI\nI\nIllllll\nI-\n"]]], "page_455": [["block_0", [{"image_0": "455_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["These two values are in reasonable agreement. (Compare them to the value of the longest\nrelaxation time for molten poly(a-methylstyrene) in Figure 11.2 and Example 11.1: chains relax\na lot more rapidly in dilute solution than in the melt.)\nTo estimate Tsega we can take Equation 11.5.4 with v 0.5 (appr0priate for a theta solvent) and\nN =M/M0 106/104 4 104;\n"]], ["block_2", ["The Rouse and Zimm models for G\u2019 and G\u201d of isolated chains have three distinct regimes of\nbehavior. At low frequencies, our] < 1, the response is that of a liquid, with G\u2019 wwz, G\" Na), and\nG\u201d > G\u2019. This terminal regime is common to all liquids and is therefore common to all models, as\nnoted in Section 11.2. At very high frequencies, (or-N > 1 where TN is the shortest relaxation time,\nthe response is solid-like: G\u2019 (00, G\u201d 00\u20141, and G\u2018r >> G\u201d . It is the third, intermediate zone that is\ncharacteristic of the BSM; the response depends on the detailed nature of the spectrum of\nrelaxation times, i.e., the internal degrees of freedom of the chain.\nWhen we turn to a molten polymer, we find that there are nowfour distinct regions of behavior.\nThis is illustrated in Figure 11.16a for G\u2019 and G\u201d, and in Figure 11.16b for (30?). The latter figure\nmay be compared to Figure 11.2; the results are qualitatively very similar. The long time (or low\nfrequency) terminal regime is just as before (albeit with a considerably different value of 71 and n)\nand the short time (high frequency) regime is solid-like, with a modulus typical of a glass\n(109_1010 Pa). Thus, it is in the internal dynamics that a new feature has emerged. In particular,\nthere is a new regime of solid-like behavior, with G\u2019 wwo % 105\u2014106 Pa. This modulus is\n"]], ["block_3", ["444\nLinear Viscoelasticity\n"]], ["block_4", ["This suggests that motions on the length scale of a styrene monomer take place in 100 ps, which is\nreasonable. It might be more in the spirit of the model to take the segment as a few persistence\nlengths. From Table 6.1 and the discussion in Section 6.4, about 10 styrene monomers correspond\nto four persistence lengths and the newly computed N2106/(10 x 104) a: 1000. This gives a\ncorrespondingly larger value for 7333 of about 3 x 10 s.\n"]], ["block_5", ["In the previous two sections we have emphasized mechanical models for the viscoelastic response\nof polymer chains and we have seen how well the BSM in the Zimm form is able to describe the\nbehavior in very dilute solutions. In large measure, these models were formulated before extensive\nexperimental tests had been performed. When we turn our attention to highly concentrated\nsolutions and melts, however, the situation is reversed. The basic experimental phenomenology\nhas been well known since the 19503 and 19605, but only in the 1970s did a successful, detailed\nmolecular model emerge. This model was initially developed by Doi and Edwards [7], based on the\nreptation hypothesis of de Gennes [8]. We will follow this historical chronology by focusing in this\nsection on the experimental phenomena and then in the next section on the reptation model itself.\n"]], ["block_6", ["We can now insert this value as R in Equation 11.5.1 to obtain the longest relaxation time:\n"]], ["block_7", ["Alternatively, we can use the value of [n] m 84 mL/g from Figure 9.5 in Equation 11.5.3:\n"]], ["block_8", ["11.6.1\nRubbery Plateau\n"]], ["block_9", ["11.6\nPhenomenology of Entanglement\n"]], ["block_10", [{"image_1": "455_1.png", "coords": [41, 220, 146, 258], "fig_type": "molecule"}]], ["block_11", ["10-4\n_,0 S\n1'\n53g \n(104)1.5\n"]], ["block_12", ["84 x 106 x 0.008\nm\nw\n10\u20145\n6x1023x1.4x10\"16><308\n3X\n3\n71\n"]], ["block_13", ["8 x 3.14 x (3 x 10%3 x 0.008 m 10\u20144\n1.4 x 10-16 x 308\nS\nTl\n9\u20185 Trot \n"]]], "page_456": [["block_0", [{"image_0": "456_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "456_1.png", "coords": [13, 46, 313, 252], "fig_type": "figure"}]], ["block_2", ["characteristic of a lightly cross-linked rubber, as we saw in Chapter 10. Qualitatively, if we follow\nG(t) in Figure 11.16b starting from short times, the initial modulus is that of a glass. Rapidly,\nhowever, the stress begins to relax, in what is called the transition zone. This character is similar to\nwhat is seen when a glass is heated through the glass transition temperature (see Chapter 12), hence\nthe terminology.\nThe remarkable new\nfeature\nis that\nat\nsome characteristic\ntime,\nTe in\nFigure 11.16b, the stress stops relaxing before it has decayed to zero. In other words, the material\nthinks it is a solid again, albeit with a much reduced modulus compared to the glassy state. This\nso-called rubbery plateau can persist for some decades in time before finally giving way to flow in\nthe terminal regime. Correspondingly, in G\" we see two peaks, one in the transition zone and one\n"]], ["block_3", ["Figure 11.16\nIllustration of the viscoelastic response of highly entangled polymer melts. (a) G\u2019 and G\u201d\nversus reduced frequency for poly(vinyl methyl ether). (Data from Kannan, RM. and Lodge, T.P., Macro-\nmolecules, 30, 3694, 1997. With permission.) (b) G(t) versus reduced time for polyisobutylene. (Data from\nCatsiff, E. and Tobolsky, A.V., J. Colloid. Sal, 10, 375, 1955. With permission.) The significance of the shift\nfactor or will be discussed in Chapter 12.\n"]], ["block_4", ["Phenomenology of Entanglement\n445\n"]], ["block_5", ["106 r\n1\na\nE\nE:\n-\n4\n"]], ["block_6", ["E5\n105\n,-\n-.\n{5\nE\ni\n"]], ["block_7", ["\u201cE\n_\nTransition\n'\n"]], ["block_8", [".92\n'\n5\n:\u2018i\n.-\n-.\n2 107\n(5\n:\nl\na,\n__\nFlubbery plateau\n1\n.\ni\n"]], ["block_9", [{"image_2": "456_2.png", "coords": [45, 259, 303, 474], "fig_type": "figure"}]], ["block_10", ["r\n4\n_,\n\u2018 o\n\u20180\ng\ng?\nrvrw=1.a>\u00ab<1o6\n1\n"]], ["block_11", ["1 \n.\nI\n.\nI\n.\nI\n.\nI\n.__I\n.\nI\n.\nI\n.\nI\n.\nI _.\nJ_ _.\nI_\n.\n"]], ["block_12", ["1 07\n"]], ["block_13", [".0\n1\nO\n5\n.\n-,\n10\n5\nTerminal .\n5\n"]], ["block_14", ["1\n3\n"]], ["block_15", ["<sub>_-sub>\n<sub>Ysub>\n25\u00b0C\n<sub>15sub>\n109\ns\na\n"]], ["block_16", ["10*12\n10-9\n1o\u20146\n10-3\n1o0\n103\n1o6\n"]], ["block_17", ["\u2014-4\n\u20142\n0\n2\n4\n6\n8\n"]], ["block_18", ["E\n\u00b0\nTo = 91\u00b0C\n'\n"]], ["block_19", ["E\n'\nI\n'\nI\n'\nI\n\"\nI\n'\nl\n'\nl\nT\nI\n_'_\nI\n'\nl\n'\n|\n'\n|\n:\nPoly(viny| methyl ether)\n3\n"]], ["block_20", ["-\nTe\n.\n-\n"]], ["block_21", ["I\n.....I ...III .uul .....I .....l . nnl .II.II \u201dan! Inul Innl .I...I .....I .u..l Inul .....I ...--l Ill-II Inul Inn:\n"]], ["block_22", ["3\nGlass\n:\n"]], ["block_23", ["_ \\\nPolyisobutylene\n\u2014\n"]], ["block_24", ["Log roar\n"]], ["block_25", ["T\n"]], ["block_26", ["\u20193T (5)\n"]]], "page_457": [["block_0", [{"image_0": "457_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "457_1.png", "coords": [13, 521, 358, 667], "fig_type": "figure"}]], ["block_2", ["at the onset of the terminal regime. The former corresponds to motions of a few monomers at a\ntime and represents the transition out of the glassy state. The second corresponds to motions of\nentire molecules, or large subsections, and represents the final transition to liquid-like behavior.\nThe rubbery plateau in G\u2019 and G(t) then arises because of an increased separation on frequency or\ntime between local relaxations and chain relaxations. A plateau in G\u2019 signifies a gap in the\nspectrum of relaxation times, which might arise for any number of reasons. In the particular\ncase of a molten \ufb02exible polymer, the gap in the relaxation time spectrum arises for a purely\ntopological reason, referred to as intermolecular entanglements.\nFor high molecular weight polymers, the individual chains become intertwined with one\nanother. The full relaxation of one chain is thus highly dependent on its surroundings, as one\nchain cannot pass through another. Consequently, the longer relaxation times of the chain are\nseverely increased by this effect, whereas relaxations of a few monomers can still proceed rapidly;\nthis gives rise to the gap in the relaxation time spectrum. The modulus in the plateau region (either\nin G0) or in G\u2019) is called GN and has a magnitude very similar to the solid modulus of the same\npolymer if it were lightly cross-linked like a rubber band. Thus the interchain entanglements can be\nthought of as temporary cross-links; for times shorter than the lifetime of an entanglement, the\nmaterial behaves as a solid, but then at longer times the material \ufb02ows. In the next section we will\nexplore the reptotz'on model, which explains how chains escape from their entanglements, but here\nwe need to examine the nature of the entanglements themselves. As we saw in Chapter 10, to a \ufb01rst\napproximation, the modulus of a lightly cross-linked rubber is given by G pRT/Mx where p is the\ndensity and M, is the molecular weight between cross-links. As with the prefactor for G in the\nBSM (see Equation 11.4.7 and associated discussion), this modulus is an example of the equipar-\ntition theorem: it is the number of network strands per unit volume, ,oN,W MI, multiplied by the\nthermal energy, kT. We now use this equation to define a molecular weight between entangle-\nments, Me:\n"]], ["block_3", ["Source:\nFrom\nFetters,\nL.J.,\nLohse,\nD.J.,\nRichter,\nD.,\nWitten, TA,\nand\nZirkel,\nA.,\nMacromolecules, 27, 4639, 1994.\n"]], ["block_4", ["In other words, Me is the average molecular weight between temporary cross-links. The parameter\nMe has been tabulated for many \ufb02exible polymers; it typically corresponds to about 50\u2014200\nmonomers. Examples are given in Table 11.2. We can now see that the relaxation of the polymer\nin the transition zone corresponds to relaxations of chain segments less than M, long, whereas in the\nterminal zone we have relaxations of chains containing many (i.e., M/Me) entanglement lengths.\nAlthough Me provides a convenient parameterization of entanglements, it is not a fully\nsatisfying explanation of the effect. For example, in a real cross-linked rubber, the cross-links\n"]], ["block_5", ["Table 11.2\nPlateau Modulus, Molecular Weight between Entanglements,\nPacking Length, and Entanglement Spacing for Some Common Polymers\nat 140\u00b0C\n"]], ["block_6", ["445\nLinear Viscoelasticity\n"]], ["block_7", ["Polyethylene\n2.6\n840\n1 .69\n33\nPoly(ethylene oxide)\n1.8\n1,600\n1.94\n37.5\n1,4-Polybutadiene\n1 .2\n1,800\n2.29\n44\n"]], ["block_8", ["Polymer\nGN (MPa)\nM, (g/mol)\np*(A)\nd (A)\n"]], ["block_9", ["1 ,4-Polyisoprene\n0.42\n5,400\n3.20\n62\nPolyisobutylene\n0.32\n7,300\n3.43\n66\nPoly(methyl methacrylate)\n0.31\n10,000\n3.46\n67\nPolystyrene\n0.20\n13,000\n3.95\n76.5\nPoly(dimethylsiloxane)\n0.20\n12,000\n4.06\n79\n"]], ["block_10", ["..\u2014__\n(11.6.1)\n"]]], "page_458": [["block_0", [{"image_0": "458_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "458_1.png", "coords": [27, 335, 172, 375], "fig_type": "molecule"}]], ["block_2", ["From Table 11.2 we can discern a qualitative correlation between Me and the bulkiness of the\nsidegroups. For example, polystyrene, poly(methyl methacrylate), and poly(dimethylsiloxane)\nhave relatively large values of Me and relatively large sidegroups whereas polyethylene,\npoly(ethylene oxide), and polybutadiene have much smaller Me. In general, \u201cthin\u201d chains entangle\nmore easily (have lower values of Me) than \u201cfat\u201d chains. If we suppose that entanglement effects\nare due to the uncrossability and mutual intertwining of different chains, then a crucial parameter\nshould be the amount of chain contour per unit volume. If we may be permitted a culinary analogy,\n"]], ["block_3", ["are identifiable chemical entities, different from ordinary monomers on the chain. In the\nentangled melt, we probably should not think of one monomer in an Me-long section of chain\n"]], ["block_4", ["as being stuck to its surroundings, while all the other ones move happily. Rather, the entangle\u2014\nment phenomenon represents the cumulative effect of many interactions between monomers on\ndifferent chains, with no single monomer behaving differently on average from any other. Thus,\nthe picture of a temporary network with particular points of entanglement, although physically\nappealing, is potentially misleading. An interesting question then arises: can we predict what Mc\nshould be for a given polymer, based on things we already know about the chemical structure?\nThe answer turns out to be yes, and we summarize the explanation given by Fetters and\ncoworkers [9].\n"]], ["block_5", ["Here Mb is the molecular weight per backbone bond, C00 is the characteristic ratio (recall Equation\n6.3.1), and E is the average backbone bond length. Therefore thinner chains will tend to have a\nsmaller packing length because they have a smaller Mb. Note, however, that more \ufb02exible chains\nwill have a smaller COO, which tends to increase p*.\nNow consider the volume of space pervaded by a chain, Vp ARE, where A is a constant of\norder unity. The pervaded volume is considerably larger than the occupied volume for a long chain.\nThe number of different chains, 11, that pack into a volume Vp is given by\n"]], ["block_6", ["Note that because Vp increases as R: M3/2, It increases with M2. Now we introduce a\nhypothesis: the onset of entanglement effects occurs when n :3 2, i.e., when M is just big enough\nthat a test chain and one other intertwine to pack a volume VP. Certainly we would not expect\nentanglement effects if n were less than 2 and maybe the onset really occurs for some larger value,\nbut that would just affect the proportionality constant in the following analysis.\nWe rewrite the radius of gyration as\n"]], ["block_7", ["a bowl of cooked cappellini has many more entanglements than the same volume of fettuccini. To\nbe more quantitative, we introduce the packing length,p*, which is the ratio ofthe volume occupied by\n"]], ["block_8", ["a chain (M/pNav) to its mean square end-to\u2014end distance; values are also listed in Table 1 1.2.\n"]], ["block_9", ["and thus\n"]], ["block_10", ["Phenomenology of Entanglement\n447\n"]], ["block_11", ["11.6.2\nDependence of Me on Molecular Structure\n"]], ["block_12", [{"image_2": "458_2.png", "coords": [36, 630, 171, 672], "fig_type": "molecule"}]], ["block_13", [{"image_3": "458_3.png", "coords": [40, 575, 115, 615], "fig_type": "molecule"}]], ["block_14", ["PNav (h2>0\n\u2014\n1\u20183'Navcoof2\n(11.6.2)\np*\n"]], ["block_15", ["R\n: \n1\n1.694\ng\n,Amn,\n(\n)\n"]], ["block_16", ["\u20191vNV\na\np M\n(11.6.3)\n"]], ["block_17", ["n:2:A\u2014\u2014-\u201499\u2014\u2014\u2014\u2014\\/L\ufb01mvav\n(men\n(61%)\n"]], ["block_18", ["M\nl\n_\nMb\n"]], ["block_19", ["COOME\n"]]], "page_459": [["block_0", [{"image_0": "459_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "459_1.png", "coords": [26, 373, 195, 413], "fig_type": "molecule"}]], ["block_2", ["We now equate M6 with the value of M for which n 2:\n"]], ["block_3", ["This relation is compared to an extensive compilation of experimental data in Figure 11.17.\nThe dependence on (19*)\"3 is very clear and the resulting value of the unknown constant A is about\n1.7. Thus we can assert that entanglement is a universal property of \ufb02exible chains and that the spacing\nof entanglements for a given polymer is determined entirely by the density (p) and the \ufb02exibility (C00).\n"]], ["block_4", ["There are two aspects to tack: forming a bond quickly and developing mechanical strength.\nThe former requires molecular motion to be rapid; higher GN means smaller Me (or M, in the\n"]], ["block_5", ["448\nLinear Viscoelasticity\n"]], ["block_6", ["Figure 11.17\nDependence of the plateau modulus, GN (and thus the molecular weight between entangle-\nments, Me) on the packing length, p*, for various polymers. (From Fetters, L.J., Lohse, D.J., Richter, D.,\nWitten, T.A., and Zirkel, A., Macromolecules, 27, 4639, 1994. With permission.)\n"]], ["block_7", ["A well-known rule of thumb in adhesion science is the Dahlquist criterion, which states that in\norder for a pressure sensitive adhesive (PSA) to have good tack, it should have a plateau modulus\nbelow 3 x 105 Pa. Tack describes the ability of an adhesive to form a bond of measurable strength,\nquickly, under a light load (e.g., Post-it Notes). Give a qualitative explanation for this rule, and\ncomment on whether the polymers in Table 11.2 might be useful as PSAs.\n"]], ["block_8", [{"image_2": "459_2.png", "coords": [35, 44, 294, 294], "fig_type": "figure"}]], ["block_9", ["where A\u2019 is a new constant A/ (12%). Using the empirical de\ufb01nition of Me (Equation 11.6.1),\nwe can express the plateau modulus as\n"]], ["block_10", ["Example 11.3\n"]], ["block_11", ["Solution\n"]], ["block_12", ["\u201dES\u201c\n0.\nE.\nz\n(5\n5\n"]], ["block_13", [{"image_3": "459_3.png", "coords": [44, 443, 127, 478], "fig_type": "molecule"}]], ["block_14", ["4M363\npNa\nM \nb\n_\n*3\n.zizcgo\ufb01\ufb01pzNgv\n(A!)\n(11.6.6)\n"]], ["block_15", ["(11.6.7)\nGN (A\u2019)\n"]], ["block_16", ["0.1\n\u2014\n_\nSloe=\u20143\n,\nL\n"]], ["block_17", ["10_\n"]], ["block_18", ["0\n_|_\nl\n"]], ["block_19", ["1'\u2014\n'.\n\u2018\n"]], ["block_20", ["1\n1O\np*(i\\)\n"]], ["block_21", ["2\nkT\n"]], ["block_22", ["(19*)3\n"]], ["block_23", [{"image_4": "459_4.png", "coords": [98, 73, 280, 238], "fig_type": "figure"}]], ["block_24", ["O\n"]], ["block_25", ["0\n"]]], "page_460": [["block_0", [{"image_0": "460_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Before proceeding to the reptation model, two more phenomenological effects of entanglement\ndeserve comment. The first is the molecular weight dependence of the viscosity. Recall from\nEquation 11.3.4 that the viscosity is the integral over G(t). Thus, when entanglement effects set in,\nthe rubbery plateau grows in extent and the area under G(t) increases markedly. The viscosity\nitself, plotted for several polymers in Figure 11.18, shows a very strong but universal dependence\non M, namely a3\u20184i0\u20182, when M > Me. This was illustrated earlier, in Figure 11.1, for one\nparticular polymer. For shorter polymers, M S Me, the dependence is much weaker because there is\nno entanglement and is more consistent with the Rouse model (17 M). The molecular weight for\nwhich the dependence changes is called MC and generally MC 2: 2\u20143 Me. The exponent of 3.4 is\nsomething that any successful theory of polymer melt dynamics must explain. From the practical\npoint of view, it means that the processing of molten polymers will be very dependent on the\naverage molecular weight; as noted in the context of Figure 11.1, an increase by a factor of two in\nM results in an increase by a factor of 10 in 17.\nThe second important phenomenological effect is that of large-scale elastic recovery. This is a\nnonlinear response and therefore not something we will cover in detail, but it serves to illustrate a\ncrucial point. The most remarkable feature of a rubber band is that not only can it be deformed by\n500% or more without breaking, but that it recovers its original shape upon letting go. The same is\ntrue of a molten polymer liquid. If a very large strain is applied quickly and then released before\nthe stress has had a chance to relax much, the liquid will also snap back to its original shape.\n(How quickly is quickly? From Figure 11.16b we can see that the polymer will behave as a rubbery\nsolid for times significantly less than 71.) It is this ability to show large scale elastic recovery that\n"]], ["block_2", ["cross-linked case) and thus more restricted chain motions. A typical PSA is soft and squishy to the\ntouch and distinctly softer than a rubber band (GN :3 106 Pa or more). At the molecular level, this\nmeans that the molecules are able to \ufb02ow and conform to the surfaces to be adhered. The\nmechanical strength part of the problem is a little more subtle. It turns out that the strength of an\nadhesive is much more than the sum of the attractive forces between the glue and the substrate;\nmost of it comes from energy dissipation in deforming the adhesive. This dissipation is achieved\nby molecular motion, so softer materials are favored for this reason too. However, this argument\ndoes not mean that the modulus should be made as small as possible; then there would be no\nstrength at all. All adhesives are polymers because the many modes of viscoelastic relaxation\nprovide many modes of dissipation, and thus good adhesive strength.\nFrom Table 11.2 we can see that the last five polymers might be candidates for PSAs (poly-\nisoprene, polyisobutylene, poly(methyl methacrylate), polystyrene, and poly(dimethylsiloxane)),\nalthough their plateau moduli are very close to the limit of 3 x 105 Pa. In order to reduce the\nmodulus of a PSA, it is common to add a lower molecular weight diluent or oil, known as a\ntackz'fier. Experimentally, GN decreases at least as c2, so adding 50% of a tackifier should drop the\nmodulus by at least a factor of (1/2)2 1/4. With this additional degree of freedom, we could make\na PSA out of almost any polymer, except for one crucial factor: the bond also needs to form\nquickly. How quick is quick enough? Well, common experience tells us that we do not want to\nhold Scotch tape in place for 10 3 while it sets, so quickly means within 1 s or less. This brings into\nplay the time dependence of the modulus: at very short times, the material will behave as a glass,\nwith G as 109 Pa or more. So, we need to make sure that timescales of say 0.1\u20141 3 fall within the\nrubbery plateau of response. Figure 11.2 tells us that poly(a-methylstyrene) would not work;\nthe rubbery plateau does not begin until 1\u201410 3 (and then only at 186\u00b0C). Figure 11.16a, on the\nother hand, suggests that polyisobutylene would be a good candidate at room temperature. In fact,\nout of our list of five polymers from Table 11.2, polystyrene and poly(methyl methacrylate) can\nbe eliminated; at room temperature for timescales of\n1 3, they behave as hard, glassy solids\n(e.g., Plexiglas). This demarcation depends on a crucial parameter\u2014the glass transition temperature\u2014\nwhjch is the subject of the next chapter. In addition to smaller values of GN, PSAs tend to have\nglass transition temperatures below room temperature.\n"]], ["block_3", ["Phenomenology of Entanglement\n449\n"]]], "page_461": [["block_0", [{"image_0": "461_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "461_1.png", "coords": [28, 37, 248, 373], "fig_type": "figure"}]], ["block_2", ["truly is the hallmark of entanglements in polymer liquids. Many systems can show a plateau in\nG\u2019 or G(t) that has nothing to do with entanglements; examples include dense colloidal suSpensions\nnear their glass transition, and densely packed, roughly spherical polymer objects (e.g., block\ncopolymer micelles, hyperbranched chains, dendrimers, and microgels). One might be tempted to\nargue on this basis that the viscoelastic response of \ufb02exible polymers has nothing much to do with\nentanglements, because other nonentangling systems also show a plateau in G\u2019 with a similar\nmagnitude to ON. This argument is fallacious, however. A plateau in G\u2019, as noted above, only\nrequires a wide separation between relaxation processes. Large\u2014scale elastic recovery in a liquid, in\ncontrast, demands some kind of interchain entanglement. Materials comprising roughly spherical\nobjects such as those noted above can usually not be extended by even 5% or 10% without falling\napart, let alone the 500% that \ufb02exible polymer liquids can sustain.\n"]], ["block_3", ["The reptation model was originally developed by de Gennes for a single, \ufb02exible chain trapped in a\npennanent network (reptation was coined from the Latin reptare, to creep) [8]. It was then\nextended to a theory for linear and nonlinear rheological response of polymer liquids by Doi and\nEdwards [7]. Here, as throughout the chapter, we will emphasize the linear response, but it is worth\nnotn in passing that the success of the reptation model in capturing nonlinear phenomena is at\nleast as impressive as its description of linear viscoelasticity. We will begin with a simple\n"]], ["block_4", ["450\nLinear Viscoelasticity\n"]], ["block_5", ["Figure 11.18\nDependence of the melt viscosity on molecular weight for a variety of polymers c and k are\narbitrary constants to shift the data. (Compiled by Berry, G.C. and Fox, T.G, Adv. Polym. Sci, 5, 261, 1968.\nWith permission.)\n"]], ["block_6", ["11.7\nReptation Model\n"]], ["block_7", ["_8_\u2019\n10 \nPoly(dimethyl\n+\n\u2014\nsiloxane)\n(J\n3\n__\nPolyisobutylene\n"]], ["block_8", ["F:-\n"]], ["block_9", ["18\n\u2014\n"]], ["block_10", ["16\n\u2014\n_\n6\n14\n\u2014\n"]], ["block_11", ["12\n\u2014\n"]], ["block_12", ["4 \nsiloxane)\n"]], ["block_13", ["5\n\u2014\nPolyethylene\nPolybutadlene\n_\nPoly(tetrarnethyl\n"]], ["block_14", ["2 \nPoly(ethylene\nglycol)\n\u2014 P0ly(vinyl acetate)\n"]], ["block_15", ["_\n__I3.4\n"]], ["block_16", ["PolystyreJn_e t i\nJ\n_|\nl\n0\n1\n2\n3\n4\n5\n6\nk+logM\n"]], ["block_17", ["Poly(rnethyl\nmethacrylate)\n"]]], "page_462": [["block_0", [{"image_0": "462_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "462_1.png", "coords": [19, 24, 188, 347], "fig_type": "figure"}]], ["block_2", [{"image_2": "462_2.png", "coords": [20, 318, 176, 424], "fig_type": "molecule"}]], ["block_3", [{"image_3": "462_3.png", "coords": [28, 182, 172, 404], "fig_type": "figure"}]], ["block_4", [{"image_4": "462_4.png", "coords": [29, 40, 195, 169], "fig_type": "figure"}]], ["block_5", [{"image_5": "462_5.png", "coords": [30, 165, 181, 301], "fig_type": "figure"}]], ["block_6", [{"image_6": "462_6.png", "coords": [32, 42, 197, 163], "fig_type": "molecule"}]], ["block_7", [{"image_7": "462_7.png", "coords": [34, 290, 179, 432], "fig_type": "figure"}]], ["block_8", ["as far as d, the obstacles have no effect. We may associate this regime of behavior with the\ntransition zone of viscoelastic response, 2\u2018 < 76; segments of the chain up to N6 units long can relax\nreadily. On very long times, such that the chain has completely lost contact with the obstacles it\nwas surrounded by at time t: 0, it has fully relaxed; this is the terminal regime t> 1-1 as far as our\ntest chain is concerned. The intermediate regime, 11, < r g 1-1, is the one we must concentrate on,\n"]], ["block_9", ["Imagine a chain trapped in a field of obstacles, as in the two-dimensional representation in\nFigure 11.19. The obstacles have an average spacing d, which we will subsequently associate\nwith the average distance between entanglements (see Table 11.2). Therefore d % bx/NQ, where N6\nis the number of monomers in M3; d is typically 30\u201480 A, i.e., longer than the persistence length\nbut less than Rg. How do the obstacles affect the relaxation and diffusion of the chain? There are\nthree regimes of behavior. On short times, such that individual monomers on average do not move\n"]], ["block_10", ["Figure 11.19\nA single chain trapped in an array of obstacles, spaced an average distance d apart. By\nsnaking backward and forward through the obstacles, i.e., by reptating, the chain eventually escapes from its\noriginal set of obstacles.\n"]], ["block_11", ["argument for the molecular weight dependence of the longest relaxation time and the diffusion\ncoefficient, before proceeding to the stress relaxation modulus.\n"]], ["block_12", ["11.7.1\nReptation Model: Longest Relaxation Time and Diffusivity\n"]], ["block_13", ["Reptation Model\n451\n"]]], "page_463": [["block_0", [{"image_0": "463_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "463_1.png", "coords": [26, 439, 259, 650], "fig_type": "figure"}]], ["block_2", ["when the chain feels the con\ufb01ning effects of the obstacles. In particular we want to find the longest\nrelaxation time, which defines when the chain has fully escaped from its initial surroundings.\nThe reptation hypothesis is that the chain ultimately escapes from the obstacles by snaking\nalong its own contour. Under the in\ufb02uence of Brownian motion, one end of the chain comes out of\nthe obstacles and explores space at random, while the other end penetrates further into the\nobstacles. However, this is a random process, so the chain is just as likely to move back in the\nopposite direction; the snake has two heads. The key point is that when the chain moves back, it is\nunder no obligation to retrace its previous path through the obstacles; it will follow a random\n"]], ["block_3", ["course. In this way it will escape the first set of obstacles, as illustrated in Figure 11.19. Now, as\ntime goes on, random motion will drive the ends of the chain further and further into the initial set\nof obstacles and then subsequent reverse motion will erase the memory of where the chain used to\nbe. The longest relaxation time will correspond to the moment when the middle segments of the\nchain finally escape from the con\ufb01nes of the initial set of obstacles.\nWe can describe this process in the following way. The set of fixed obstacles with spacing d will\nbe replaced by a confining tube of diameter d, as in Figure 11.20. This tube is defined by the\nconformation of the test chain at time t: 0', it is itself a random walk in three dimensions. The real\npolymer segments will move freely within the tube due to the rapid segmental motions, but the\nwhole chain will only escape the tube by reptating out of the ends. We replace the real chain\n(N monomers, statistical segment length b, R: s/6) with an average chain that has N/Ne\nentanglement lengths (with d2 =Neb2) trapped in the tube. When one end of this coarse-grained\nchain diffuses out of the end of the tube and the other end moves a distance d into the tube, that\nportion of the tube is erased (see Figure 11.20). The entire tube will be erased when the chain\ncenter-of-mass has diffused a distance proportional to the length of the tube, L. We can de\ufb01ne a\ndiffusion coef\ufb01cient for the motion within the tube, Dmbe, by\n"]], ["block_4", ["Figure 11.20\nThe obstacles in Figure 11.19 can be replaced by a tube with diameter d; the process of\nreptation gradually erases the tube from the ends inward.\n"]], ["block_5", ["where Trap is the reptation time, the time needed to completely erase the tube. This diffusion\nprocess is postulated to have a Rouseelike dependence on chain length, i.e., the friction of the chain\n"]], ["block_6", ["452\nLinear Viscoelasticity\n"]], ["block_7", ["(L2) 2r),ut,.;rrep\n(11.7.1)\n"]], ["block_8", [{"image_2": "463_2.png", "coords": [54, 406, 229, 465], "fig_type": "molecule"}]]], "page_464": [["block_0", [{"image_0": "464_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This equation can be recast in terms of molecular weight by recognizing that N M[M0, where M0\nis the monomer molecular weight, N/Ne M/Me, and rseg bzgo/kT:\n"]], ["block_2", ["The crucial feature of this result is that the longest relaxation time of the chain is predicted to vary\nwith M3. This dependence is much stronger than the Zimm (r, ~M3\u20192) or Rouse (1'1 ~M2) results\nand arises because of the fact that the chain conformation (or stress) is relaxed only by the motion\nof the chain ends, rather than by the concurrent relaxation of all segments of the chain. As noted at\nthe beginning of the chapter, describing the molecular weight dependences of the various dynamic\nproperties is the first goal of molecular models.\nIn addition, we can consider the long-time translational diffusion of the chain. At the instant the\nchain finally escapes the initial tube, one of its ends must just be touching the initial tube at some\npoint. The initial tube and the new tube are uncorrelated random walks, except that they touch in\nthis manner. Therefore in the time Twp the center of mass of the chain moves a mean-square\ndistance of approximately (hz), and therefore we can write\n"]], ["block_3", ["Thus Dt decreases as M\u20182, which is also a much stronger dependence than the Zimm (Dt M2)\nand Rouse (Dtr-vM\"1) models. This particular prediction of the reptation model inspired a great\ndeal of experimental activity, including the development and application of several new ap-\nproaches to measuring 0,. The results overall are in reasonable agreement with Equation 11.7.6,\nexcept that the M exponent (about \u20142.3) is slightly stronger than anticipated. An example is\nshown in Figure 11.21 for hydrogenated 1,4-polybutadiene (essentially polyethylene) measured by\ndifferent researchers using several different techniques. Two possible reasons for the difference\nbetween the experimental and theoretical M dependences will be discussed in Section 11.7.3.\n"]], ["block_4", ["moving in its tube is proportional to the number of tube segments, N/Ne, and the friction factor of\neach tube segment, \u00a36. The latter quantity is just the number of monomers in the segment, Ne, times\nthe monomeric friction factor, \u00a30. (The monomeric friction factor is similar to the friction factor of\n"]], ["block_5", ["Now that we have the reptation prediction for the longest relaxation time, we can also develop\npredictions for G(t) and therefore G\u2019, G\", and 17. At short times (in the transition zone, I < 11,) the\nchain sections between entanglements relax in a Rouse-like fashion before the relaxation effect-\nively ceases with G(t > re) ON. The reptation model does not really address this regime.\n"]], ["block_6", ["so that we can solve Equation 11.7.1 for Twp\nZ.1: z 1: 211. z mm,\nZDtube\nNe\n2kT\nZkTNe\n"]], ["block_7", ["a bead in the Rouse model, but is calculated to correspond to the friction per repeat unit.) Using this\nargument we find\n"]], ["block_8", ["Of course, the contour length of the tube is directly related to the length of the chain, namely\n"]], ["block_9", ["Reptation Model\n453\n"]], ["block_10", ["11.7.2\nReptation Model: Viscoelastic Properties\n"]], ["block_11", [{"image_1": "464_1.png", "coords": [36, 274, 123, 310], "fig_type": "molecule"}]], ["block_12", [{"image_2": "464_2.png", "coords": [41, 164, 143, 201], "fig_type": "molecule"}]], ["block_13", [{"image_3": "464_3.png", "coords": [45, 434, 141, 493], "fig_type": "molecule"}]], ["block_14", ["N\nN\nL=Ed=EJMb\n(11.7.3)\n"]], ["block_15", ["Trep\ngm'rseg\n(11.15)\n"]], ["block_16", ["(112) W 60.71,,\n"]], ["block_17", ["1 N, H\"\n(11.7.6)\nWanna;\n"]], ["block_18", ["M3\n"]], ["block_19", ["friction\n(N/NBMe\n(N/Ne)(Neg\u20190)\nNg\u2019o\n"]], ["block_20", ["(11.7.2)\n"]], ["block_21", ["(11.7.4)\n"]]], "page_465": [["block_0", [{"image_0": "465_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "465_1.png", "coords": [24, 518, 187, 605], "fig_type": "molecule"}]], ["block_2", ["These predictions are compared with experimental data for a hydrogenated polybutadiene in\nFigure 11.22. The main discrepancy occurs in G\u201d(w) for frequencies in the plateau zone where\nexperimentally the decrease in G\u201d with a) is rather less rapid than predicted. It will turn out that this\ndifference can be accounted for in much the same way as the difference in M exponents for D,\nidenti\ufb01ed above, but again we will defer this discussion to the next section.\n"]], ["block_3", ["454\nLinear Viscoelasticity\n"]], ["block_4", ["The time constants for the modes have a p2 dependence, which echoes the Rouse model, but the\namplitude of each mode is further attenuated by a factor of I/pz. Furthermore, only odd numbered\nmodes contribute due to the symmetry of the reptation process; the center of mass does not move\nfor even numbered modes. Consequently, Equation 11.7.7 is not much different from a single\nexponential decay, dominated by the longest relaxation time. The dynamic moduli, G\u2019 and G\", can\nbe obtained from Equation 11.7.7 through the sine and cosine Fourier transforms (Equation\n"]], ["block_5", ["Figure 11.21\nExperimental data for the diffusion of hydrogenated (or deuterated) polybutadienes from\nvarious sources. (Compiled by Lodge, T.P., Phys. Rev. Lett., 83, 3218, 1999. With permission.)\n"]], ["block_6", ["At longer times, in the rubbery plateau and in the terminal regime, the escape from the tube also\ninvolves a spectrum of Rouse-like modes, except that the chain is confined to the tube. The\nargument above only addressed the longest relaxation time, but to avoid a good deal more\nmathematics we will just state the result:\n"]], ["block_7", ["1 1.3.7):\n"]], ["block_8", ["10\u20148 g\nj\na\ni\n?\nNE\n10\u20189\nJ\n.9.\nE\n3\nc:\nE\n_\n"]], ["block_9", [{"image_2": "465_2.png", "coords": [39, 395, 322, 442], "fig_type": "molecule"}]], ["block_10", [{"image_3": "465_3.png", "coords": [39, 52, 283, 283], "fig_type": "figure"}]], ["block_11", [{"image_4": "465_4.png", "coords": [43, 528, 186, 570], "fig_type": "molecule"}]], ["block_12", ["8\n1\nz\ns\n1\nG(t)zGN Z EECXPC\u2014E) =GNZ 7536x136 p2:\n)\n(11.7.7)\n"]], ["block_13", ["8\n1\n(or\ng\u201d : GN'\u2014\n_ _____B___\n"]], ["block_14", ["1\n10\u201410%\n3\nE\n3\n"]], ["block_15", ["10\u201411:\ng\nE\na\n"]], ["block_16", ["T\n10\"\n2\n_I_\nl\nI\n_I_Ll LlJ.\nl\nJ\nI\nI lllll\nI\nI\nl\nI llll\n103\n104\n105\n106\n"]], ["block_17", ["8\n1\n2\nG, = GN\"\u20142 Z 7 \u201c((0719)2\n7T\nOdd P p\n1 + ((0713)\n"]], ["block_18", ["10\u20146 :\nl\u2014l_|'\nl\u2018|_|\u2019ll\u2018|\u2014\n_I_\n_l\u2014l\nI [III[\nI\nI\nI\nI_l_l_l'_\n"]], ["block_19", ["E\ni\n-\nh(d)PB Melts at 175\u00b0C\n1\n10\u20147 E\na\nE\nj\n\u2014\nD: 18.7 x M?30\nj\n"]], ["block_20", ["'1\n"]], ["block_21", ["l\u2014\n"]], ["block_22", ["772 \u00ab1;; P2\n1 + (p)2\n"]], ["block_23", ["odd P\nodd p\nTrep\n"]], ["block_24", ["M (g/mol)\n"]], ["block_25", ["(11.7.8)\n"]]], "page_466": [["block_0", [{"image_0": "466_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Consider the following polymer processing problem. Imagine you are making a plastic hoop by\ninjecting molten polymer into a mold. The molten polymer travels in two directions from the\ninjection point and meets on the opposite side of the hoop. How long do you need to wait for\nthe two liquid streams to merge after they first meet and develop full mechanical strength at the\njunction? Assume the polymer is polystyrene, with M >> M, and a viscosity of 105 P at the\nprocessing temperature, and use the reptation model. (The answer is of practical importance, in\nthe sense that the faster the part sets, the faster it can be ejected from the mold and the more parts\ncan be made per unit time. And, although the proposed geometry is rather artificial, the issue of\nhealing a polymer\u2014polymer interface is a very general one.)\n"]], ["block_2", ["Lastly we turn to the steady \ufb02ow viscosity. This can be obtained by integrating over the\nmodulus (Equation 11.3.4):\n"]], ["block_3", ["Note that in performing this integration we ignore the modulus for times shorter than re, or in other\nwords, I: 0 for the stress relaxation process corresponds to the onset of the rubbery plateau. This\napproximation is fine, because the time ranges are so different that the early time portion\ncontributes essentially nothing to the integral for a high M chain. The main result of Equation\n11.7.9, and arguably the main result of the basic reptation model, is that the viscosity is predicted\nto scale as M3. This should be compared with the universal experimental result of M\u201c, illustrated\nin Figure 11.1 and Figure 11.18. On the one hand, the fact that such a simple idea can get a\nnontrivial answer (1; ~M3), which is not too different from reality, is very encouraging. On the\nother hand, the difference is significant. Accordingly, after the following example we brie\ufb02y\nconsider two omissions of the reptation model that are thought to account for the discrepancies\nwith the experimental results for DL, G\", and n. The main predictions of reptation were also\nsummarized in Table 1 1.1.\n"]], ["block_4", ["Figure 11.22\nComparison of the basic reptation theory with experiment for the dynamic moduli of\nhydrogenated polybutadiene melts. The main discrepancy is in the slope of G\u201d at higher frequencies. (Data\nfrom Tao, H., Huang, C.-I., and Lodge, T.P., Macromolecules, 32, 1212, 1999.)\n"]], ["block_5", ["Example 11.4\n"]], ["block_6", ["Reptation Model\n455\n"]], ["block_7", ["(Pa)\n"]], ["block_8", ["G\"\n"]], ["block_9", ["G',\n"]], ["block_10", [{"image_1": "466_1.png", "coords": [42, 53, 276, 228], "fig_type": "figure"}]], ["block_11", [{"image_2": "466_2.png", "coords": [42, 336, 111, 380], "fig_type": "molecule"}]], ["block_12", ["103\n"]], ["block_13", ["7T2\n11 E,repGN\n(11.7.9)\n"]], ["block_14", ["107\nI\nIIIIIHI\nI\nllllllll\nl\u2018llllilll\n[lllltlll\nI.\nIllllI-ll\n"]], ["block_15", ["10-2\n10-1\n100\n101\n102\n103\n"]], ["block_16", ["lIll'l\u2018l\u2018l\nL\n"]], ["block_17", ["lllllll'\n.I\n"]], ["block_18", ["1111\u201d]\n"]], ["block_19", ["I\n"]], ["block_20", ["llll\n"]], ["block_21", ["I\nI]\nlllll\nI\n1Il[[ll]\nl\nllllllll\nI\nllllllll\nI\n1\n111111\n"]], ["block_22", ["G!\n"]], ["block_23", ["co (radls)\n"]], ["block_24", ["ll\"\n"]], ["block_25", ["|IIIIII|\n"]], ["block_26", ["_I\n"]], ["block_27", ["_LIIlLIlJ_|_Il|LLLI_I_l\n"]]], "page_467": [["block_0", [{"image_0": "467_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The experimental scaling laws (D,~M_2'3, a3'4) have stronger M dependences than pre-\ndicted by the reptation model. Numerically, the experimental results indicate more rapid chain\nmotion than by reptation alone. This is illustrated in Figure 11.23 for the viscosity. For chains\ncontaining only a few Me the viscosity is significantly lower than expected by the model. However,\nas M increases, the steeper experimental dependence suggests that the experiments might converge\non the reptation prediction at very high M. One obvious candidate for this more rapid relaxation is\nthe fact that the reptation model was developed for a chain moving in an array of fixed obstacles. In\nreality, of course, the obstacles are a way of accounting for the entanglements with other chains,\nand because all chains are moving, some entanglements should disappear while a test chain is\ntrying to reptate. In the tube language, we could imagine that the tube develops occasional leaks,\nthereby allowing the chain to escape through the leaks rather than the ends (see Figure 11.24a).\nThis process is known as constraint release, and because it accelerates the chain relaxation, it\ncould explain why 77 is lower and D[ is higher compared to pure reptation. However, a quantitative\ndevelopment of constraint release is complicated, and it is not yet clear how much of the\ndiscrepancy between experiment and theory can be accounted for in this manner.\nA second correction, termed contour length \ufb02uctuations, can actually reproduce the experi-\nmental M dependences of n and 0,, plus the frequency dependence of G\u201d(w), rather well. This\nprocess applies even to a chain in an array of fixed obstacles. In the discussion preceding Equation\n11.7.8, we argued that relaxation modes for which the center of mass does not move down the tube\ndo not contribute to G(t). In fact, this is not quite true. Imagine that we pin the center of the chain,\nbut let the ends wiggle around. There are \u201caccordion\u201d modes, in which the two ends of the chain\npenetrate into (and out of) the average tube by a certain amount. If they do so, they can select a\ndifferent path on their way out, and thus they accelerate the relaxation of the chain ends. This is\nillustrated in Figure 11.24b. (Of course, the ends must also occasionally \ufb02uctuate out of the tube,\n"]], ["block_2", ["The mechanical strength of a polymer part will depend on achieving full interpenetration or\nentanglement of the various molecules. Assume that we start with two \ufb02at polymer surfaces that\nare brought into contact at time t: 0. To achieve full interpenetration, we need to wait until the\naverage chain has diffused a distance of about R3. We have not been given direct information about\nM for our polymer, so estimating Rg and Dt (from 77) could be tedious. However, in the reptation\nmodel Twp is nothing more than the time to diffuse Rg, and Twp itself is simply related to n and the\nplateau modulus, GN, through Equation 11.7.9. We can take the value of GN from Table 11.2; it is\nabout 2 x 105 Pa. This gives\n"]], ["block_3", ["where we ignore the numerical prefactor of 772/12 in this estimate, and are careful to make the units\nof viscosity and modulus match.\nThe answer suggests that interfacial healing will not be a problem in this processing operation,\nas it will presumably take several seconds to fill the mold, and several more to cool the part\nsufficiently to remove it from the mold. Note also that the form of this solution is exactly the same\nas in Example 11.1, where we estimated the flow time from the modulus and the viscosity using the\nMaxwell model. As a last comment, the issue of interfacial healing is much richer than this\nanalysis reveals. For example, significant mechanical strength at an interface will actually deve10p\nafter the chains have interpenetrated a distance of only about the tube diameter or entanglement\nspacing and then it will continue to grow slowly as the chains become completely intermixed.\nA full description of the evolution of interfacial strength with time therefore requires the complete\nspectrum of relaxation times.\n"]], ["block_4", ["456\nLinear Viscoelasticity\n"]], ["block_5", ["11.7.3\nReptation Model: Additional Relaxation Processes\n"]], ["block_6", ["Solution\n"]], ["block_7", ["Trepg\n10\nP\nlPaSmOOS s\n2 x 105 Pa\n10 P\n"]], ["block_8", ["5\n"]]], "page_468": [["block_0", [{"image_0": "468_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "468_1.png", "coords": [21, 338, 424, 465], "fig_type": "figure"}]], ["block_2", [{"image_2": "468_2.png", "coords": [31, 354, 310, 614], "fig_type": "figure"}]], ["block_3", ["Figure 11.23\nDiscrepancy between experimental results and the reptation prediction for the viscosity. The\ndata are for hydrogenated polybutadiene at 140\u00b0C. (Reported in Tao, H., Lodge, T.P., and von Meerwall, E.D..\nMacromolecules, 33, 1747, 2000. With permission.)\n"]], ["block_4", ["distance into the tube by Brownian motion, without moving the center of mass, and therefore relax some stress\nwithout reptation; these movements are called contour length \ufb02uctuations.\n"]], ["block_5", ["Figure 11.24\nIllustration of additional relaxation processes that can bring reptation predictions into closer\nagreement with experiment. (a) The two entangling chains indicated by solid circles diffuse away, allowing\nthe test chain to move sideways by the process of constraint release. (b) The ends of the chain can move some\n"]], ["block_6", ["Reptation Model\n457\n"]], ["block_7", ["102 \n\u2018n(Pa!\ufb02\nl\n"]], ["block_8", [{"image_3": "468_3.png", "coords": [38, 468, 280, 595], "fig_type": "figure"}]], ["block_9", [{"image_4": "468_4.png", "coords": [42, 33, 260, 262], "fig_type": "figure"}]], ["block_10", [{"image_5": "468_5.png", "coords": [45, 337, 254, 459], "fig_type": "figure"}]], ["block_11", ["10\n;\u2014\nReptation\n4 \n"]], ["block_12", ["100-\nn=eoo><10-1\u2018WI3-43\n"]], ["block_13", ["108\nI\nI\n|\nI\n| I\nI I\n|\nI\n|\nI\nl\nI\nI II\nI\n|\nI\nI\n"]], ["block_14", ["104 ;\nI\u00e9g\ufb01jg\n"]], ["block_15", [";\n(n=56x1041M%\n,:\n"]], ["block_16", ["104\n105\n106\n"]], ["block_17", [{"image_6": "468_6.png", "coords": [111, 341, 326, 460], "fig_type": "figure"}]], ["block_18", ["A4(ghnon\n"]], ["block_19", [{"image_7": "468_7.png", "coords": [192, 428, 455, 608], "fig_type": "figure"}]], ["block_20", ["y\n"]], ["block_21", [{"image_8": "468_8.png", "coords": [237, 343, 439, 586], "fig_type": "figure"}]], ["block_22", [{"image_9": "468_9.png", "coords": [238, 346, 444, 462], "fig_type": "figure"}]], ["block_23", ["nnm\n"]], ["block_24", ["IIIIIII\n"]], ["block_25", ["llllll\n"]], ["block_26", ["II\n"]], ["block_27", ["Innul\n"]], ["block_28", ["n\n"]], ["block_29", ["IIIIIII\n"]], ["block_30", ["IIIIIIIII\n"]], ["block_31", ["IIIIIIII\n"]], ["block_32", ["Illllllll\n"]], ["block_33", ["ml\n"]], ["block_34", ["lIlIIIlI\n"]], ["block_35", [{"image_10": "468_10.png", "coords": [253, 482, 447, 597], "fig_type": "figure"}]], ["block_36", [{"image_11": "468_11.png", "coords": [261, 345, 436, 453], "fig_type": "molecule"}]]], "page_469": [["block_0", [{"image_0": "469_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "469_1.png", "coords": [32, 433, 187, 471], "fig_type": "molecule"}]], ["block_2", ["So in this case, both the shear rate and the shear strain are oscillatory functions of time. The shear rate\nis not, however, a function of position in the gap, and therefore by Equation 9.1.3 or Equation ll.l.l\nthe stress is the same everywhere in the sample. This geometry is less useful for lower viscosity\nfluids because they tend to leak out of the gap.\nThe primary limitation of the shear sandwich is the exclusion of steady flow experiments.\nTo overcome this, rotational rheometers are the common solution. The Couette geometry\ndescribed in Section 9.4.2 is one approach; another is the cone and plate, illustrated in Figure\n11.25b. In this apparatus, the sample is confined in the narrow gap between a flat, fixed plate,\nand a conical piece, which makes a small angle 6 (typically 10\u201430) with the flat plate. The cone\nrotates around the vertical axis with rotation rate .Q (rad/s), sweeping out an angle (35 with time.\nThe key feature of this design is that the shear rate is homogeneous throughout the sample. To\nsee how this comes about, consider the strain rate at any radial distance r. The width of the gap\nd increases linearly with r, namely d r tan 6 m r sin 6 % r6 for small 6 (also in radians). The\ninstantaneous linear velocity of the moving cone in the tangential direction, vg, also increases\nlinearly with r: v\u00a2,: r0. The net result is that for any value of r the instantaneous shear rate\nis vd,/d:(2/9.\n"]], ["block_3", ["but that has no effect on the escape from the tube if we consider the center to be pinned.) As this\nprocess is a random fluctuation, we expect that the characteristic distance the ends can penetrate\nshould be roughly the square root of the total, i.e., a distance of about (dN/N6)U2. This is a\nnegligible fraction of the tube for very long chains, but turns out to be significant for most\nexperimentally accessible chain lengths.\n"]], ["block_4", ["In Chapter 9 we described two flow geometries in common use for the measurement of the\nviscosity in steady flow, namely the capillary (Poiseuille flow) and the concentric cylinder\n(Couette flow). Although both of these can serve for transient and dynamic measurements, at\nleast two other geometries are also commonly employed, and particularly for the higher viscosities\nassociated with molten polymers. These are represented by the parallel plate (shear sandwich)\nrheometer for dynamic measurements and the cone-and\u2014plate rheometer capable of dynamic,\ntransient, and steady shear measurements. Both of these will be described brie\ufb02y below and then\nwe conclude by identifying several important general issues that arise in rheometry.\n"]], ["block_5", ["The shear sandwich geometry is illustrated in Figure 11.25a. A central \ufb02at plate is driven up and\ndown between two fixed, parallel plates and the sample is contained within the two narrow gaps on\neach side of the moving plate. This arrangement is nothing more than an experimental realization\nof the parallel surface configuration used in Figure 9.1 and Figure 11.1. It is made possible by the\nuse of an oscillatory strain; clearly, steady flow cannot be achieved with this design. As drawn in\nFigure 11.25a, the plates have an area A =Lh, and the total volume of sample contained in the two\ngaps is 2dLh. If the moving plate has a displacement along the x axis of x0 sin out, its velocity v,[\nwill be xow cos wt. Furthermore, the fluid velocity at each \ufb01xed plate will be zero (under the no-\nslip assumption). If the velocity profile across each gap is linear, as is the case for sufficient small\ngap widths (1, then the velocity at any point across the gap is (vx/d)y, where y is the distance from\nthe fixed surface. Recalling the discussion in Section 9.1 and Section 11.1, the shear rate in the gap\nis therefore given by\n"]], ["block_6", ["458\nLinear Viscoelasticity\n"]], ["block_7", ["11.8.1\nShear Sandwich and Cone and Plate Rheometers\n"]], ["block_8", ["11.8\nAspects of Experimental Rheometry\n"]], ["block_9", ["d\ndvx\nv,\nyzdit\u2019: dy \n(11.8.1)\n"]]], "page_470": [["block_0", [{"image_0": "470_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "470_1.png", "coords": [28, 101, 308, 232], "fig_type": "figure"}]], ["block_2", [{"image_2": "470_2.png", "coords": [31, 52, 234, 266], "fig_type": "figure"}]], ["block_3", ["There exist several texts in the field of rheometry that interested readers may consult for much\nfuller treatments of this important topic. There are also several commercial ventures that specialize\nin rheological equipment, so that a wide variety of instruments are readily accessible. We conclude\nthis section with a brief listing of some of the important issues that can arise in choosing and using\n"]], ["block_4", ["a rheometer:\n"]], ["block_5", ["Aspects of Experimental Rheometry\n459\n"]], ["block_6", ["(a)\n"]], ["block_7", ["(b)\n"]], ["block_8", ["Figure 11.25\nIllustration of the (a) shear sandwich and (b) cone and plate geometries.\n"]], ["block_9", ["11.8.2\nFurther Comments about Rheometry\n"]], ["block_10", ["1.\nAll rheometers measure\na force (or torque) related to the\nstress\nand\na displacement\n(or velocity) related to the strain (or strain rate). Some rheometers are strain\u2014controlled,\nmeaning that the Operator inputs a desired strain amplitude and the instrument measures\nthe stress required to achieve that strain; others are stress\u2014controlled and impose a stress\nand monitor the\nstrain.\nStrain\ncontrol\nis\nnecessary\nto measure\nthe\nstress relaxation\n"]], ["block_11", [{"image_3": "470_3.png", "coords": [37, 110, 172, 221], "fig_type": "figure"}]], ["block_12", ["Y\nH'\nX0 Sin cut\n"]], ["block_13", [{"image_4": "470_4.png", "coords": [114, 32, 318, 265], "fig_type": "figure"}]], ["block_14", [{"image_5": "470_5.png", "coords": [134, 314, 206, 378], "fig_type": "molecule"}]], ["block_15", [{"image_6": "470_6.png", "coords": [183, 102, 317, 224], "fig_type": "figure"}]]], "page_471": [["block_0", [{"image_0": "471_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this chapter we have examined the linear viscoelastic properties of \ufb02exible polymers, both in\ndilute solution and in the molten state. We have defined the basic concepts and experimental\napproaches, emphasizing shear \ufb02ow. Molecular models have been presented for both dilute\nsolutions and melts and they are quite successful in describing the experimental phenomena.\nThe main points are as follows:\n"]], ["block_2", ["modulus, whereas stress control is required for a creep experiment. In the linear viscoelastic\nregime, either approach should be suitable, given the interrelationships among the various\nviscoelastic functions outlined in Section 11.3. However, in many cases it is the nonlinear\nproperties that are of interest and then the choice of control mode becomes more important.\n2.\nIn the measurement of force and displacement, transducers are required to convert the\nmeasured quantity into an electrical current or voltage. The force transducer is particularly\nimportant, in terms of sensitivity, range, speed of response, linearity, and reproducibility. The\ndata in Figure 11.1, Figure 11.2, Figure 11.16, and Figure 11.18, for example, illustrate how\nthe modulus and viscosity can vary over many orders of magnitude and no single transducer\ncan cover this entire range. The measured force for a low viscosity \ufb02uid can be increased\nsomewhat by increasing the area of the moving surface and by increasing the strain rate\n(and vice versa for a force that is too high to measure reliably); but in the end, a range of\ntransducers or even a range of instruments are required to cover the full spectrum of materials\nfrom dilute solutions to highly entangled melts.\n3.\nIn the dynamic mode, the accessible frequency range is also important. Most commercial\ninstruments can manage approximately 0.001\u201410 Hz, and some custom instrumentation can\nachieve kHz or even MHz frequencies. However, the data in Figure 11.14 extend over 10\u201415\norders of magnitude along the time or frequency axis, well beyond the capability of any single\ninstrument. How are such measurements accomplished? The answer will be given in the next\nchapter, but it involves the use of variable temperature, and the principle of time\u2014temperature\nsuperposition.\n4.\nWe have not discussed the temperature dependence of the viscoelastic properties in this\nchapter; that subject, too, is deferred to Chapter 12. However, the temperature dependence\nis generally very strong, and so accurate rheometry requires good temperature control.\nUnfortunately, this is not so easy to attain in many cases, particularly far above or far below\nroom temperature. Part of the difficulty arises from having the sample and moving parts\nsurrounded by air or inert gas, neither of which have good heat transfer characteristics.\nA related problem is viscous heating, which arises from the substantial dissipation of energy\nat high viscosities or high flow rates (see Equation 9.1.7).\n5.\nRheometric experiments integrate the response over the entire sample, in the sense that only a\nsingle measurement of force is used to determine the stress. For example, if the strain field is\ninhomogeneous for any reason, the stress will be different at different locations in the sample,\nbut the measurement will not take that into account. The inhomogeneity might arise from\ntemperature gradients, sample nonuniformity, or secondary flows induced at high flow rates,\nbut in addition it is always present under the category of edge effects. In every geometry there\nis a sample surface that is not in contact with either the fixed or moving surfaces of the\napparatus, and the \ufb02ow profile in the vicinity of this extra surface must differ from that\nassumed in the calculation of the stress. Schemes have been developed for partially correcting\nedge effects in all of the common rheometer geometries.\n"]], ["block_3", ["460\nLinear Viscoelasticity\n"]], ["block_4", ["11.9\nChapter Summary\n"]], ["block_5", ["1.\nPolymer liquids generally show viscoelastic behavior, i.e., a response to an imposed deformation\nthat is intermediate between the viscous \ufb02ow of liquids and the elastic deformation of solids. This\nresponse can be characterized by a variety of material functions, such as the steady flow viscosity,\nstress relaxation modulus, creep compliance, and dynamic modulus or dynamic viscosity.\n"]]], "page_472": [["block_0", [{"image_0": "472_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n461\n"]], ["block_2", ["Problems\n"]], ["block_3", ["I\u2018W.W. Graessley, T. Masuda, J.E.L. Roovers, and N. Hadjichristidis, Macromolecules, 9, 127 \n"]], ["block_4", ["1.\nThe following are approximate 0' (in dyn/cmz) versus 4/ data for three different samples of\npolyisoprene in tetradecane solutions of approximately the same concentration?[\n"]], ["block_5", [{"image_1": "472_1.png", "coords": [40, 465, 316, 583], "fig_type": "figure"}]], ["block_6", ["From plots of these data, estimate the Newtonian viscosity of each of the solutions and the\napproximate rate of shear at which non-Newtonian behavior sets in. Are these two quantities\nbetter correlated with the molecular weight of the polymer or the molecular weight of the arms?\n"]], ["block_7", ["M... (g/mol)\n1.61 x 106\n1.95 x 106\n1.45 x 106\nLinear\nFour-armed star\nSix-armed star\n6 (g/cm3)\n0.0742\n0.0773\n0.0078\nids\u20141)\n0'><10_3\n0'><10_3\nO'XIO_3\n"]], ["block_8", ["The basic character of viscoelastic response is revealed by the simple Maxwell and Voigt\nmechanical models. In general, a polymer liquid behaves more like an elastic solid at short\ntimes or high frequencies and more like a viscous liquid at long times or low frequencies. The\ndemarcation between these limits is determined by the relaxation times of the material.\nIn the limit of linear response, i.e., sufficiently small strain amplitudes and strain rates such\nthat the material functions do not depend on the strain amplitude or rate, the Boltzmann\nsuperposition principle provides a direct route to calculate the viscosity, dynamic moduli, and\nrecoverable compliance from the stress relaxation modulus.\nThe BSM of Rouse and Zimm provides a molecular explanation for the viscoelastic\nresponse of polymers. The Zimm version, which includes intramolecular hydrodynamic\ninteractions, is very successful in dilute theta solutions, and variable solvent quality can be\nincorporated by a simple dynamic scaling argument. The Rouse version, without hydro-\ndynamic interactions, applies very well to low molecular weight molten polymers.\nFor high molecular weight polymers in concentrated solutions or melts, the phenomenon of\nentanglement dominates the viscoelastic properties. The moduli exhibit four regimes of\nbehavior denoted glassy, transition zone, rubbery plateau, and terminal, as a function of\nincreasing time or decreasing frequency. In the rubbery plateau the liquid behaves as a soft\nsolid, with a modulus similar to a lightly cross-linked rubber (Chapter 10). A characteristic\nmolecular weight between entanglements is inferred, and may be predicted based solely on\nknowledge of the chain \ufb02exibility and density.\nThe reptation model provides a physically appealing description of chain motion and stress\nrelaxation in entangled polymers. The theoretical predictions for diffusion and viscosity do not\nquite match the experimental results, but good agreement can be obtained when the additional\nprocesses of constraint release and contour length \ufb02uctuations are included.\n"]], ["block_9", ["0.6\n0.7\n\u2014\n\u2014\n0.8\n0.9\n\u2014\n\u2014\u2014\n"]], ["block_10", ["1\n1\n0.3\n\u2014\n2\n2\n0.6\n\u2014\n4\n4\n1\n0.15\n8\n5\n2\n0.3\n10\n6\n3\n0.4\n20\n7\n4\n0.8\n60\n7\n2\n100\n8\n4\n"]]], "page_473": [["block_0", [{"image_0": "473_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["462\nLinear Viscoelasticity\n"]], ["block_2", ["2.\nNewtonian viscosities of polystyrene samples of different molecular weights were determined\nat 200\u00b0C by Spencer and Dillon.\u2019r Use these data to determine the exponent of M in the\nrelationship between 7; and M.\n"]], ["block_3", ["(a) Evaluate the power loss per cycle if the material is a Hookean solid: 0' Gy.\n(b) Evaluate the power loss per cycle if the material is a Newtonian liquid: 0 n (d'y/dt).\n(c) Brie\ufb02y comment on the signi\ufb01cance of these results.\n4.\nUsing complex notation, derive the Maxwell model predictions for the dynamic shear modu-\nlus, G*(w), and the dynamic viscosity, n*(w), and the relations between the elastic and viscous\ncomponents of each; assume a dynamic strain, 31* yo exp(iwt).\n5.\nFor polystyrene (M 600,000) at 100\u00b0C, the following values describe the creep compliance,\nJ(t), at long timeszi\n"]], ["block_4", ["Use Equation 11.2.8 to evaluate the viscosity of the polymer at this temperature. Then use\nEquation 11.7.5 and Equation 11.7.9 and Table 11.2 to estimate the segmental relaxation time\nfor polystyrene at this temperature.\n6.\nFind the relation between the phase angles 5 and 11: given in Equation 11.2.21 and Equation\n11.2.23.\n7.\nEstimate the high molecular weight values of the dimensionless group 071 /RE for dilute\n\ufb02exible chains in a theta solvent, and for the same polymer in the melt, using appropriate\ntheoretical models.\n8.\nSketch a careful log\u2014log plot of the Zimm theory prediction for [n\u2019]R and [1)\u201d]R versus MN,\nwhere TN is the shortest relaxation time of the model. Mark the decades on the axes to make the\ncurves realistic. Perform two pairs of curves on the same axes, one for a large N, i.e., a high\nmolecular weight polymer and one for a small N polymer to indicate the effects of molecular\nweight on the response.\n"]], ["block_5", ["3.\nIn a dynamic experiment with y(t) yo sin(cot) the power loss per cycle of oscillation is given\nby\n"]], ["block_6", ["1R.S. Spencer and RE. Dillon, J. Colloid Sci, 4, 241 (1949).\ntD.J. Plazek and V.M. O\u2019Rourke, reported in JD. Ferry, Viscoelastic Properties ofPolymers, 3rd ed., Wiley, New York, 1980.\n"]], ["block_7", [{"image_1": "473_1.png", "coords": [41, 107, 151, 230], "fig_type": "figure"}]], ["block_8", [{"image_2": "473_2.png", "coords": [45, 90, 225, 245], "fig_type": "figure"}]], ["block_9", ["wt=21r\nJ\n0' d7\n(of: 0\n"]], ["block_10", ["86\n3.50 x 103\n162\n4.00 x 104\n196\n6.25 X 104\n360\n4.81 x 105\n490\n1.89 x 106\n508\n1.00 X 106\n510\n1.64 x 106\n560\n3.33 x 106\n710\n6.58 x 106\n"]], ["block_11", ["M x 10\"3\nn (P)\n"]], ["block_12", ["logJ(t)(m2/N)\n\u20141.8\n\u20141.4\n\u20141.0\n\u20140.6\n\u20140.2\n+0.1\nlog1\u2018(S)\n12.6\n13.0\n13.4\n13.8\n14.2\n14.6\n"]]], "page_474": [["block_0", [{"image_0": "474_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Problems\n453\n"]], ["block_2", ["1\u2018D.S. Pearson, et al., Macromolecules, 20, 1133 (1987).\n"]], ["block_3", ["9.\nPropose two possible reasons why the Zimm model fails to exhibit shear-thinning behavior at\nhigh shear rates.\n10.\nEstimate the longest relaxation time, 1'1, for polystyrenes with M =105 and 10\u00b0, in cyclo\u2014\nhexane at 35\u00b0C and toluene at 25\u00b0C. There is no single correct way to do this, so be sure to\nidentify any assumptions you make.\n11.\nPearson, et al.Jr reported measurements of the viscosity and diffusivity of narrow distribution\npolyethylenes at 175\u00b0C. The data are given below. Prepare log\u2014log plots of 71 versus M and D\nversus M and compare with expectations based on the Rouse and reptation models. How well\ndo the data agree with these theories? What do you propose for the origin of any discrepan\u2014\ncies? What is MC for PE? Prepare a plot of the product D17 versus M. Does this agree better\nwith theory in the putative Rouse regime? Why is this the case?\n"]], ["block_4", ["12.\nConsider the tracer diffusion coefficient of polystyrenes in the melt, at 176\u00b0C. At this\ntemperature, Me: 13,000 g/mol, M0: 104 g/mol, and lcT/{210\u20189 cmz/s. What is Dt for\nM 13,000 under these conditions? What is Dt for M 65,000 dissolved as a tracer in a melt\nwith M 13,000? Suppose the M 13,000 matrix polymer was end-functionalized at both\nends, so that it could be cross-linked. Suppose that a small amount of bifunctional agent was\nadded and reacted to completion; the M 13,000 polymers were therefore all end-linked to\nform very long linear chains. What would Dt be for the M 65,000 tracer then? Suppose that\nthe cross\u2014linking agent was tetrafunctional, so that a complete network was formed. What\nwould Dt be for the M 65,000 tracer in this case?\n13.\nImagine you have a narrow distribution sample of 1,4-polybutadiene with Mw=54,000\ng/mol. On one set of logarithmic axes, sketch the shape of the stress relaxation modulus\nG(t) versus time that you would expect to see for this polymer in the melt. Extend your curve\nto cover the full range of viscoelastic response, and estimate any numerical values that you\n"]], ["block_5", ["can. (For polybutadiene, p=0.9 g/mL, b 26.9 151, Me: 1800 g/mol.) Then add two more\ncurves, corresponding to the expected response after cross-linking: (i) 0.05% of the mono-\nmers and (ii) 1% of the monomers. Indicate clearly which curve is which, and brie\ufb02y explain\nwhy the three curves differ from each other (if they do), and why in some respects they are\nthe same. (It may be helpful to recall Section 10.1.)\n14.\nUse the correlation of plateau modulus and packing length developed in Section 11.6.2 to\npredict the molecular weight between entanglements, Me, for poly(viny1 acetate). Table 6.1\nprovides useful data for the chain dimensions, and the density is approximately 1.08 g/cm3.\nHow does your value compare with the experimentally reported value of 7000 g/mol?\n"]], ["block_6", [{"image_1": "474_1.png", "coords": [47, 196, 243, 359], "fig_type": "figure"}]], ["block_7", ["506\n6.6\n0.0185\n590\n5.4\n0.0232\n618\n4.8\n0.0248\n695\n3.5\n0.035\n1,280\n1.4\n0.092\n2,390\n0.35\n0.338\n3,310\n0.15\n0.873\n4,100\n0.093\n1.63\n13,600\n0.012\n37.7\n32,100\n0.0020\n1,100\n"]], ["block_8", ["M (g/mol)\n10\u00b0 D (c/s)\nn (P)\n"]], ["block_9", ["1 19,600\n0.00013\n125,000\n"]]], "page_475": [["block_0", [{"image_0": "475_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Bird, R.B., Armstrong, RC, and Hassager, 0., Dynamics ofPolymer Liquids, Volume 1, 2nd ed.; Bird, R.B.,\nCurtiss, C.F., Armstrong, RC, and Hassager, 0., Dynamics of Polymer Liquids, Volume 2, 2nd ed.,\nWiley, New York, 1987.\nDoi, M. and Edwards, S.F., The Theory of Polymer Dynamics, Clarendon Press, Oxford, 1986.\nFerry, J.D., Viscoelastic Properties of Polymers, 3rd ed., Wiley, New York, 1980.\nGraessley, W.W., Polymeric Liquids and Networks: Structure and Properties, Garland Science, New York,\n2003.\nMacosko, C.W., Rheology: Principles, Measurements, and Applications, VCH, New York, 1994.\nMorrison, F.A., Understanding Rheology, Oxford University Press, New York, 2001.\nRubinstein, M. and Colby, R.H., Polymer Physics, Oxford University Press, New York, 2003.\nShaw, M.T. and MacKnight, W.J., Introduction to Polymer Viscoelasticity, 3rd ed., Wiley, Hoboken, NJ,\n2005.\nWalters, K., Rheometry, Chapman and Hall, London, 1975.\n"]], ["block_2", ["6.\nSammler, R.L. and Schrag, J.L., Macromolecules 21, 1132 (1988); Lodge, A.S., and Wu, Y.-J., Rheol.\nActa 10, 539 (1971).\nDoi, M. and Edwards, S.F., J. Chem. Soc, Faraday Trans. 2, 74, 1789, 1802, 1818 (1978).\nde Gennes, P.G., J. Chem. Phys. 55, 572 (1971).\n9.\nFetters, L.J., Lohse, D.J., Richter, D., Witten, T.A., and Zirkel, A., Macromolecules 27, 4639 (1994).\n"]], ["block_3", ["464\nLinear ViscoeIasticity\n"]], ["block_4", ["1.\nMaxwell, J.C., Phil. Trans. 157, 49 (1867).\n2.\nVoigt, W., Ann. Phys. 47, 671 (1892).\n3.\nBoltzmann, L., Ann. Phys. Chem. 7, 624 (1876).\n4.\nRouse, P.E., Jr., J. Chem. Phys. 21, 1872 (1953).\n5.\nZimm, B.H., J. Chem. Phys. 24, 269 (1956).\n"]], ["block_5", ["901\u201c]\n"]], ["block_6", ["References\n"]], ["block_7", ["Further Readings\n"]], ["block_8", ["15.\nUse the reptation model to estimate the longest relaxation time for the linear polyisoprene in\nProblem 1. How does the inverse relaxation time compare with the onset of non\u2014Newtonian\nresponse? Explain.\n"]]], "page_476": [["block_0", [{"image_0": "476_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Here we resume a sequence of three chapters that treat polymers in the solid state. In Chapter 10\nwe examined the formation of polymer networks and rubber elasticity. In Chapter 13 we will\nconsider crystallinity in polymers and the associated crystallization transition from the high\ntemperature, liquid state. In this chapter we take up the subject of the glass transition, whereby a\npolymer liquid is cooled in such a way as to solidify without adopting a crystalline packing.\nAmong the three classes of polymer solid\u2014\u2014network, crystal, and glass\u2014the glassy state is the\nmost universal; relatively few polymers are used to form networks; a significant fraction can never\ncrystallize, but all can form glasses. Furthermore, all three topics are central to understanding the\nutility of polymer materials. In Chapter 9 and Chapter 11 we covered some properties of polymer\nliquids, especially those pertaining to \ufb02ow. In almost all cases, polymers are synthesized, charac-\nterized, and processed in the liquid state, and consequently the material in Chapter 9 and Chapter\n"]], ["block_2", ["11 represents a foundation for many diverse areas of polymer science. However, most polymer\napplications rely on the properties in the solid state; consequently, Chapter 10, Chapter 12, and\nChapter 13 provide the background for understanding how polymers are chosen or developed for\none application or another.\n"]], ["block_3", ["To begin, we need a working definition of a glass. A reasonable one may be simply stated: a glass\nis an amorphous solid. By amorphous, we mean that there is no long-range order or symmetry in\nthe packing of the molecules. In this sense, the structure of a glass looks very much like the\nstructure of a liquid. However, a glass is a solid: it does not flow over relevant timescales.\nAlthough the preceding statement is apparently innocent, the phrase \u201crelevant timescales\u201d hints\nat a fundamental complexity: if we are talking about an equilibrium state, time should play no role.\nWhat we will see is that glasses are generally nonequilibrium states, metastable in a sense, and\nkinetic issues will be of central importance. When a polymer liquid is cooled, the density increases\nand the molecular relaxation times increase. Over some range of temperature, the molecular\nmotion will become so slow that an equilibrium packing of the molecules cannot be attained\nduring the experiment. When this happens we say that the sample has undergone the glass\ntransition, or has vitri\ufb01ed, and we associate with each polymer a glass transition temperature,\n"]], ["block_4", ["as a solid, but below any temperature at which we intend to process the polymer as a liquid. As we\nwill see, in practice Tg is not a thermodynamic property of the polymer, and it can adopt a range of\nvalues for a given polymer, but nevertheless it is an extremely useful parameter.\nAlthough we will cover crystallization in polymers in the next chapter, a few pertinent points\nare appropriate here. Upon cooling from the liquid state, a polymer might either crystallize or turn\ninto a glass; thus these two transitions are, in a sense, in competition. Polymers with irregular\nmicrostructure, such as atactic vinyl polymers or mixed-microstructure polydienes, cannot crys\u2014\ntallize, because the lack of a regular structure at the monomer length scale prevents the formation\n"]], ["block_5", ["Tg. The value of Tg is the single most important characteristic in choosing a polymer for a given\napplication. It must lie significantly above any temperature at which we intend to use the polymer\n"]], ["block_6", ["12.1\nIntroduction\n"]], ["block_7", ["12.1.1\nDefinition of a Glass\n"]], ["block_8", ["Glass Transition\n"]], ["block_9", ["12\n"]], ["block_10", ["465\n"]]], "page_477": [["block_0", [{"image_0": "477_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Tg, which leads into the fourth section concerning kinetic models of the transition. The most\npopular of these, the free volume model, provides a simple basis for understanding one of the\nmost dramatic consequences of the\nglass transition,\nnamely the\nvery\nstrong\ntemperature\ndependence of molecular relaxation above Tg. This temperature dependence underlies the utility\nof the principle of time\u2014temperature superposition, which is of central importance to the\nexperimental characterization of viscoelasticity. This topic is covered in Section 12.5. The chapter\nconcludes with a discussion of how Tg can be modified (Section 12.6), and an introduction to\nthe properties of glassy polymers (Section 12.7).\n"]], ["block_2", ["In the preceding chapter, time (or frequency) was the primary independent variable under consid-\neration. We saw that at short times of observation, polymers exhibit high values of the modulus,\nroughly three or four orders of magnitude higher than those shown in the rubbery state of these\nmaterials. The transition between the two values of the modulus occurs over a range of time at\nfixed temperature in the so-called transition zone of viscoelasticity. It turns out that temperature\nvariation at fixed time can produce changes in mechanical properties that parallel those resulting\nfrom shifts of timescale. This change in mechanical behavior signals the glass transition, and when\nmonitored during temperature variation it occurs near the glass transition temperature. We shall\nreturn to an examination of the equivalency of time and temperature with respect to effects on\nmechanical properties later in this chapter. For now, however, it is desirable to consider some other\nproperties of matter that change near Tg; although a variety of observables are available, we shall\nemphasize volume.\nFigure 12.1 illustrates schematically the range of possibilities for the variation in specific\nvolume, Vsp, with temperature. Remember that V3], is the reciprocal of the density; Vsp, rather\nthan density, is chosen to describe these changes in anticipation of the \u201cfree volume\u201d interpret-\nation to be presented in Section 12.4. Path ABDG in Figure 12.1 shows how V5,, changes upon\nfreezing\na low molecular weight compound. (A few substances\u2014water is the best known\nexample\u2014occupy a larger volume per unit mass in the solid state than in the liquid, and for\nthese the transition at the melting point Tm would correspond to a jump rather than a drop in\nvolume.) What is significant is that this transition occurs at a single temperature, the melting point\nTm. The slopes of AB and DG re\ufb02ect the coefficients of thermal expansion of the liquid, 0:1, and the\ncrystalline solid, ac, respectively. These coefficients are approximately independent of tempera-\nture (i.e., AB and DG are nearly straight lines) but the main point is that VSp is different for solids\nand liquids; it shows a discontinuity at the melting point.\nAn entirely different pattern of behavior is shown along lines ABHI. In this case there is no\ndiscontinuity at Tm. The line AB, which characterizes the liquid, changes slope at Tg to become HI.\nActually, the change in slope occurs over a range of temperatures (about 200C), as suggested by\nFigure 12.1, but extrapolation of the two linear portions permits a single Tg to be defined. The\nregion HI characterizes the glassy state, and the threshold for its appearance is the glass transition\ntemperature. In the region BH, the liquid is said to be supercooled.\n"]], ["block_3", ["of a unit cell. Therefore crystallization is not an option for many polymers. For polymers that do\ncrystallize, the kinetics associated with finding the correct packing are sufficiently slow that only a\nfraction of the molecules succeed; consequently the material is termed semicrystalline. The\nremaining fraction is amorphous, and can undergo a glass transition. Thus the glass transition\ntemperature plays an important role even for crystallizable polymers.\nThe remainder of the chapter is organized as follows. We conclude this introductory section\nwith a more detailed comparison of the glass\u2014liquid and crystal\u2014liquid transitions, focussing on the\ntemperature dependence of the specific volume (or density). In the second section, we consider the\nglass transition from a thermodynamic point of view, especially emphasizing the fundamental\nquestion of whether there is actually a thermodynamic transition hidden beneath the kinetically\ndominated Tg. The third section describes the most common experimental routes to characterizing\n"]], ["block_4", ["466\nGlass Transition\n"]], ["block_5", ["12.1.2\nGlass and Melting Transitions\n"]]], "page_478": [["block_0", [{"image_0": "478_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "478_1.png", "coords": [29, 45, 285, 250], "fig_type": "figure"}]], ["block_2", ["Each of the two paths we have discussed could describe the behavior of either high or low\nmolecular weight compounds. This is not to say, however, that each is equally probable for the two\nclasses of compounds. For most low molecular weight materials, special effort must be made to\nsuppress crystallization and achieve glass formation. With polymers, on the other hand, the glassy\nstate is always obtained, whether a particular polymer is crystallizable or not. The mere fact that\nmolecular structure allows the possibility of crystal formation does not mean that the latter occurs\nrapidly or completely. The line ABCEFG in Figure 12.1 describes the situation of a partly\ncrystalline, partly amorphous polymer. At Tm crystallization begins and the characteristic discon-\ntinuity in specific volume occurs. The sharpness of Tm is not as pronounced for polymers as for low\nmolecular weight compounds, as evidenced by the trailing off between C and E. In the region EF\nthe volume contraction reflects the supercooling of the amorphous portion of the polymer. The\nchange in slope between EF and FG occurs at T , just as it would in the absence of crystallization.\nIf partial crystallization occurs, the amount of amorphous material is decreased and the change in\nslope at Tg may be harder to detect in this case.\nThe line ABJK in Figure 12.1 is a displaced variation of ABHI in which AB is a liquid, BI is a\nsupercooled liquid, and JK is a glass. The experimental variable that causes region JK to be offset\nfrom H1 is the cooling rate, ABJK being the route for the more rapidly cooled polymer. Since Tg is\nidentified from the change in slope, it is apparent that Tg is also displaced, appearing at a higher\ntemperature (T; in the figure) for higher rates of cooling. The change in slope that defines Tg may\nalso be viewed as the first departure from the behavior extrapolated from the liquid state. In other\nwords, although the supercooled liquid is not at thermodynamic equilibrium, its specific volume\nfollows the same T dependence as the equilibrium liquid. Thus the glass transition on cooling\nrepresents the first obvious departure from equilibrium, and it is intuitively reasonable that the\nhigher the rate of cooling, the sooner this departure will become apparent.\nTo summarize some basic observations for polymers:\n"]], ["block_3", ["1.\nAbove Tm the material is liquid. The zero-shear viscosity depends strongly on the molecular\nweight of the polymer (see Chapter 11) and temperature (see Section 12.4) but it would be\nconsidered high by all standards.\n2.\nBetween Tm and Tg, depending on the regularity of the polymer and on the experimental\nconditions, this domain may be anything from almost 100% crystalline to 100% amorphous.\nThe amorphous fraction, whatever its abundance, behaves like a supercooled liquid in this region.\n"]], ["block_4", ["Figure 12.1\nSchematic illustration of possible changes in the specific volume of a polymer with tempera-\nture. See text for a description of the significance of the various lettered features.\n"]], ["block_5", ["Introduction\n467\n"]]], "page_479": [["block_0", [{"image_0": "479_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "479_1.png", "coords": [6, 70, 360, 328], "fig_type": "figure"}]], ["block_2", ["The kinetic nature of the experimental glass transition was noted in the previous section, but it is\nnevertheless instructive to consider the possibility of a thermodynamic description of the transition\n"]], ["block_3", ["The foregoing description introduces the phenomena with which we shall be dealing in this\nchapter. As noted above, both high and low molecular weight compounds are capable of displaying\nthese effects, but the chain structure of the polymer molecules is responsible for the reversed\nprobabilities. The specific identity of the polymer anchors these transitions to some particular\nregion of the temperature scale; a list of representative values of Tg and Tlm is provided in\nTable 12.1. The regularity of the microstructure of the polymer molecule, along with experimental\nconditions, determines the extent of crystallization. The glassy state is thus seen as a lowest\ncommon denominator shared by all polymers, because 100% crystallinity is virtually impossible.\nThis promotes Tg to the position of importance assumed by Tm for low molecular weight\ncompounds. The fact that the mechanical properties undergo such profound change at Tg also\ncontributes to the significance of this parameter.\n"]], ["block_4", ["3.\nBelow Tg the material is hard and rigid with a coef\ufb01cient of thermal expansion equal to\nroughly half that of the liquid. With respect to mechanical properties, the glass is closer in\nbehavior to a crystalline solid than to a liquid. In terms of molecular order, however, the glass\nmore closely resembles the liquid. In this temperature region, the noncrystalline fraction\nacquires the same glassy properties it would have if the crystallization had been suppressed\ncompletely.\n4.\nThe location of Tg depends on the rate of cooling. In principle the location of TIm is not subject\nto this variability, but in fact, the degree of crystallinity does depend on the conditions of the\nexperiment, as well as on the nature of the polymer. For example, if the rate of cooling exceeds\nthe rate of crystallization, there may be no observable change at Tm, even for a crystallizable\npolymer (see Chapter 13).\n"]], ["block_5", ["Poly(hexamethylene adipamide)\n50\n265\nPoly(ethylene terephthalate)\n70\n265\nPoly(vinyl alcohol)\n90\n240\nPoly(vinyl chloride)\n90\n270\nPolystyrene\n100\n240\nPoly(methyl methacrylate)\n110\n183\nPoly(tetrafluoroethylene)\n130\n330\nPolycarbonate of bisphenol A\n150\n330\nPoly(oxy-2,6-dimethyl-1,4-pheny1ene)\n210\n3 10\nPoly(p-phenylene terephthalamide)\n240\n325\n"]], ["block_6", ["Poly(dimethylsiloxane)\n\u2014 123\n\u2014 40\nPolyethylene\n\u2014\u2014 120\n135\n1,4-Polybutadiene (cis)\n\u2014\u20141 12\n12\nPolyisobutylene\n\u201475\n44\n1,4-Polyisoprene (655)\n-\u201470\n28\nPoly(ethylene oxide)\n\u201470\n65\nPolypropylene\n\u2014 10\n188\nPoly(vinyl acetate)\n30\n\u2014\n"]], ["block_7", ["Table 12.1\nRepresentative Values of the Glass Transition Temperature and the\nMelting Temperature (for Stereoregular Forms, where Applicable) for Some\nCommon Polymers\n"]], ["block_8", ["453\nGlass Transition\n"]], ["block_9", ["12.2\nThermodynamic Aspects of the Glass Transition\n"]], ["block_10", ["Polymer\nTg (\u00b0C)\nTm (\u00b0C)\n"]]], "page_480": [["block_0", [{"image_0": "480_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "480_1.png", "coords": [34, 454, 192, 495], "fig_type": "molecule"}]], ["block_2", ["There is no discontinuity in volume at the Curie point, but there is a change in the temperature\ncoefficient of V, as evidenced by a change in slope. To understand why this is called a second-order\ntransition, we begin by recalling the de\ufb01nition of some relevant physical quantities:\n"]], ["block_3", ["Thermodynamic Aspects of the Glass Transition\n469\n"]], ["block_4", ["that occurs near T3. Most phase equilibria in common experience, such as boiling and melting, are\nexamples of what are called first-order transitions. There are other, less familiar but also well-\nknown transitions in nature that are not first order. The disappearance of ferromagnetism at a\nparticular temperature (called the Curie point) is an example of such a transition. Rather than the\ndiscontinuities in S, V, and H characteristic of first\u2014order transitions, these variables merely exhibit\n"]], ["block_5", ["Since V experiences a change of slope at the second-order transition, i.e., (EN/6\u20197\"\u00bb, and (8V/3p)7\nhave different values on each side of the transition, it is a and K that show the discontinuities at the\nsecond-order transition rather than V itself. The term second order comes about because the\nquantities may be written as second derivatives of the free energy, G, as follows. Recalling\nEquation 7.1.3 and Equation 7.1.4, here applied to a one-component system\n"]], ["block_6", ["Figure 12.2a and Figure 12.2b describe a second\u2014order transition schematically, in terms of V, S, a,\nand K. By extension, an nth-order phase transition is associated with discontinuities in nth-order\nderivatives of the free energy.\n"]], ["block_7", ["2.\nThe isothermal compressibility K2\n"]], ["block_8", ["a change in slope with increasing temperature. Since this is similar to the behavior near T , it is\nimportant to consider whether the glass transition is actually a second-order phase transition, or,\nperhaps, whether the kinetically affected experimental transition actually masks an underlying\nthermodynamic transition.\n"]], ["block_9", ["we can expand Equation 12.2.1 and Equation 12.2.2 to obtain\n"]], ["block_10", ["and\n"]], ["block_11", ["where the last form exploits Equation 12.2.3. From this relation it is apparent that the heat capacity\nshould also be discontinuous at a second-order transition, whereas the enthalpy (H) should behave\n"]], ["block_12", ["Another useful quantity in this context is the heat capacity at constant pressure, Cp:\n"]], ["block_13", ["12.2.1\nFirst-Order and Second-Order Phase Transitions\n"]], ["block_14", ["1.\nThe coefficient of thermal expansion 0::\n"]], ["block_15", [{"image_2": "480_2.png", "coords": [37, 604, 228, 642], "fig_type": "molecule"}]], ["block_16", [{"image_3": "480_3.png", "coords": [40, 401, 178, 438], "fig_type": "molecule"}]], ["block_17", [{"image_4": "480_4.png", "coords": [41, 506, 183, 547], "fig_type": "molecule"}]], ["block_18", [{"image_5": "480_5.png", "coords": [45, 300, 128, 337], "fig_type": "molecule"}]], ["block_19", ["1\n8\n[(6%)]\na=\u2014\u2014\n\u2014\u2014\n=\u2014\u2014\n\u2014\n(12.2.4)\nV\n61\" p\nV 6T\n8p r p\n"]], ["block_20", ["8H\n85\n826\nC, (ET); TCa\u2014f); \u2014T(W)p\n(12.2.6)\n"]], ["block_21", ["36\n36\nv- (a) S__(3_T)p\n(12.2.3,\n"]], ["block_22", ["_\n1\n3V\nor =\n17(5Tl\n(12.2.1)\n"]], ["block_23", ["1\n8V\nE \n\u2014\n12.2.2\nK\nV (\u201819P)\n(\n)\n"]], ["block_24", ["1\nav\n1\n826\n_ __ _\n\u2014_- __ _\n12.2.\nv13\u00bb. vs\u00bb.\n(\n5)\n"]], ["block_25", ["1\n(8V)\n"]]], "page_481": [["block_0", [{"image_0": "481_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "481_1.png", "coords": [15, 25, 369, 190], "fig_type": "figure"}]], ["block_2", [{"image_2": "481_2.png", "coords": [18, 43, 263, 137], "fig_type": "figure"}]], ["block_3", [{"image_3": "481_3.png", "coords": [30, 434, 129, 478], "fig_type": "molecule"}]], ["block_4", [{"image_4": "481_4.png", "coords": [34, 333, 219, 374], "fig_type": "molecule"}]], ["block_5", ["470\nGlass Transition\n"]], ["block_6", [{"image_5": "481_5.png", "coords": [35, 558, 211, 599], "fig_type": "molecule"}]], ["block_7", ["The behavior of these various thermodynamic functions at \ufb01rst\u2014order and second-order transitions\nis compared schematically in Figure 12.2.\nThe Clapeyron equation is a well\u2014known thermodynamic relation that applies to first\u2014order\ntransitions:\n"]], ["block_8", ["like V and S and show only a change in slope. In the following section the importance of the heat\ncapacity in the calorimetr\u2018ic determination of Tg will become apparent. The entropy term in\nEquation 12.2.6 can be inverted to provide another useful relation that shows how measurements\nof heat capacity versus temperature can be used to determine the entropy:\n"]], ["block_9", ["phase transition, and of (c) a, K, and Cp at a second\u2014order transition.\n"]], ["block_10", ["This expression relates the variation of the pressure\u2014temperature coordinates of a first-order\ntransition (i.e., the phase boundary) in terms of the changes in S and V that occur there. The\nClapeyron equation cannot be applied to a second\u2014order transition because AS and AV would be\nzero and their ratio unde\ufb01ned. However, we may apply L\u2019H\u00e9pital\u2019s rule to both the numerator and\ndenominator of the right\u2014hand side of Equation 12.2.8 to establish a limiting value of dp/dT. In this\nprocedure we may differentiate either with respect to p,\n"]], ["block_11", ["(a)\nT\n(b)\nT\n"]], ["block_12", ["or T\n"]], ["block_13", ["Figure 12.2\nSchematic illustration of the behavior of V, S, and H at a (a) first\u2014order and (b) second-order\n"]], ["block_14", ["maH\nL////\nmaH\n"]], ["block_15", [{"image_6": "481_6.png", "coords": [38, 622, 228, 663], "fig_type": "molecule"}]], ["block_16", [{"image_7": "481_7.png", "coords": [46, 54, 179, 125], "fig_type": "figure"}]], ["block_17", ["S(T2) 5m) :[97-51 dT JCP d lnT\n(12.2.7)\n"]], ["block_18", ["(d2) \n(93:)\n(12 2 9)\ndT 2nd\n(aAV/ap)T,2nd\nAK 2nd\n"]], ["block_19", ["dp\nAS\n\u2014\n:\n\u2014\u2014~\u2014\n1\n.2.\n(dT)lst\n(Av)lst\n(2\n8)\n"]], ["block_20", ["98\nI =\nMp\n(12 2 10)\ndT 2nd\n(aAv/aT)p,2nd\nTVA\u201c 2nd\n"]], ["block_21", [{"image_8": "481_8.png", "coords": [103, 331, 203, 381], "fig_type": "molecule"}]], ["block_22", [{"image_9": "481_9.png", "coords": [111, 50, 270, 214], "fig_type": "figure"}]], ["block_23", [{"image_10": "481_10.png", "coords": [112, 132, 282, 212], "fig_type": "figure"}]], ["block_24", ["TI\nT1\n"]], ["block_25", ["T2\nT2\n"]], ["block_26", ["(C)\nT\n"]], ["block_27", [{"image_11": "481_11.png", "coords": [219, 56, 360, 121], "fig_type": "figure"}]], ["block_28", ["1\nI\n"]], ["block_29", [".\n"]]], "page_482": [["block_0", [{"image_0": "482_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["On the assumption that the glass transition is a second\u2014order thermodynamic transition, estimate\nthe pressure dependence dTg/dp of\nT5\nusing the following data for poly(vinyl chloride): Tg :347\nK, v,, 0.75 cms/g, Aa :3.1 X 10\n\u2014\nK\u20141 and AC,, 0.068 cal/ K/g.T\n"]], ["block_2", ["The following example illustrates how results like these can be applied.\n"]], ["block_3", ["Thermodynamic Aspects of the Glass Transition\n471\n"]], ["block_4", ["This quantity has been measured directly to be 0.016 K/atm. Note that a pressure change of about\n60 atm is required to change Tg by 1 K. Note also that the stated value of Tg is somewhat different\nfrom that given in Table 12.1, underscoring the variability from sample to sample or from\ntechnique to technique.\n"]], ["block_5", ["Despite these useful thermodynamic relationships, it is clear that the glass transition in practice\nis not truly a second-order phase transition. A moment\u2019s re\ufb02ection reveals that the source of this\nreservation is the doubt about the state of equilibrium for the glass transition. Implicit throughout\nthe thermodynamic arguments above has been the notion that the phases on either side of a\ntransition are in thermodynamic equilibrium. This enters the mathematical formalism from the\nstart: Equation 12.2.3 assumes that the free energy of a phase is described by only two variables, in\nthis case p and T. Although the glass transition is certainly affected by p and T, it is also dependent\non the time of observation, as indicated in Figure 12.1. Because of this time dependence, the\nexperimental glass transition must involve more than a simple second\u2014order transition. In terms of\nstability (recall Section 7.5) the equilibrium liquid (above Tm) and the crystalline state (below Tm)\nare stable, meaning that the free energy is at a minimum. The glassy state is unstable, i.e., the\nsystem is constantly evolving into a state of lower free energy (although this evolution may\nproceed at a glacial pace). The supercooled liquid can be viewed as metastable, because the free\nenergy is in a local minimum but not the global minimum (i.e., the crystal); but for a noncrystalliz-\nable polymer, the supercooled state is effectively at equilibrium.\n"]], ["block_6", ["to generate some additional useful expressions. All of the A\u2019s in these equations refer to the\ndifference in the value of the variable from one side (prime) of the transition temperature to\nthe other (double prime):\n"]], ["block_7", ["Invert Equation 12.2.10 and substitute. The ratio of gas constants is convenient for unit conver\u2014\nsions:\n"]], ["block_8", ["The preceding discussion indicates what should be expected for a true thermodynamic second\u2014\norder transition, and therefore underlines how the glass transition both resembles and is distinct\nfrom such a transition. One way to reconcile these characteristics of TE is to invoke an underlying\nthermodynamic transition that in practice is masked by kinetic effects. For example, if we were to\nperform a cooling experiment at progressively slower rates, would the obtained values of Tg\ncontinue to decrease steadily, or would they converge to a limiting value? In the latter case, we\n"]], ["block_9", ["Solution\n"]], ["block_10", ["TData from J.M. O\u2019Reilly, J. P01ym.Sci., 57,429 (1962).\n"]], ["block_11", ["Example 12.1\n"]], ["block_12", ["12.2.2\nKauzmann Temperature\n"]], ["block_13", ["A0: = or\" or\u201d;\nAK K\u2019 K\";\nACp ;,- Cg\n(122-11)\n"]], ["block_14", ["dTg _ TgVAar _ (347 K)(0.75 cm3 g\u20141)(3.1 x 10-4 K\u201d) X\n1.99 cal\n"]], ["block_15", ["dp \nACp \n(0.068 cal g\u20141 K\u20141)\n82 atm cm\n"]]], "page_483": [["block_0", [{"image_0": "483_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "483_1.png", "coords": [31, 66, 250, 231], "fig_type": "figure"}]], ["block_2", ["We will not develop this theory in detail, but its physical content and basic conclusions can be\nreadily appreciated. The treatment begins with a lattice, similar in spirit to that employed in the\nFlory\u2014Huggins theory of mixing in Chapter 7. In this case, however, the solvent is replaced by\nvoids, or vacancies on the lattice, which will play a role qualitatively similar to that of the \u201cfree\n"]], ["block_3", ["Figure 12.3\nIllustration of the \u201cKauzmann paradox\u201d: the speci\ufb01c entropy of the glass (dashed line) would\nbe less than that of the crystal (solid line) below the Kauzmann temperature, TK.\n"]], ["block_4", ["would have located an apparently thermodynamic transition. It turns out that experimentally it has\nnot proven possible to answer this question de\ufb01nitively, but an argument can be made that some\nkind of a transition must intervene even for infinitely slow cooling. This argument was originally\nmade by Kauzmann [l], and has since become known as the \u201cKauzmann paradox\u201d or the \u201centropy\ncrisis.\u201d Kauzmann showed that if the entropy of the supercooled liquid were to continue to follow\nthe temperature dependence seen just above Tg, then eventually the entropy of the glass would be\nless than the entropy of the crystal. This in itself violates no laws of thermodynamics, but does\nseem highly counterintuitive. However, if this state of affairs persisted, the entropy of the glass\nwould go to zero at a temperature above 0 K, which would violate the third law of thermodynam-\nics. On the other hand, if a transition intervened, at which the heat capacity dropped sufficiently,\nthen neither of these difficulties would arise.\nKauzmann\u2019s argument can be seen from the schematic diagram in Figure 12.3. The specific\nentropy (entropy per gram of material) drops with decreasing T in the liquid state, down to Tm.\nAlong the crystal branch, there is a discontinuous drop in entropy at the first-order transition.\nBelow Tm the entropy goes smoothly to zero at 0 K (by the third law of thermodynamics), and may\nbe computed from the experimental heat capacity by Equation 12.2.7. Along the supercooled liquid\nbranch, the heat capacity is larger than in the crystal, and thus the entropy drOps more rapidly with\ndecreasing T than for the crystal. By extrapolation, therefore, the supercooled liquid entropy will\nequal the crystal entropy at a finite temperature, the Kauzmann temperature, TK. Experimentally,\nsuch extrapolations give values of TK that are about 50\u00b0 below the measured Tg. On the other hand,\nif a thermodynamic transition intervenes, then the heat capacity of the glass would be lower than\nthat of the supercooled liquid, perhaps very close to that of the crystal, and the problem would be\naverted. Note, however, that a smooth reduction in heat capacity with decreasing temperature\ncould also avoid the Kauzmann paradox without the necessity for an intervening phase transition.\n"]], ["block_5", ["472\nGlass Transition\n"]], ["block_6", ["12.2.3\nTheory of Gibbs and DiMarzio [2]\n"]], ["block_7", ["0\nTrn\nT\n"]], ["block_8", ["I\n"]], ["block_9", ["I\n"]], ["block_10", ["I\nl\n"]], ["block_11", ["I\n"]], ["block_12", ["I\n"]], ["block_13", ["I\n"]], ["block_14", ["|\n"]]], "page_484": [["block_0", [{"image_0": "484_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Given all of these characteristics, it is tempting to interpret T2 as corresponding to the Tg that would\nbe obtained in the limit of infinitely slow cooling. In so doing, the Kauzmann paradox would also\nbe resolved.\nDespite these successes, it is probably fair to say that the Gibbs\u2014DiMarzio theory has not been\nwidely adopted as a general description of the glass transition. One reason for this lies in the fact\nthat many different classes of materials can exhibit a glass transition, such as inorganic networks\n(e.g., \u2014Si02\u2014~, our everyday glass\u201d), ionic liquids, small organic molecules (e.g., glycerol,\n0-terphenyl), and even colloidal particles. Consequently, a description based on chain \ufb02exibility\n"]], ["block_2", ["where z is the number of nearest neighbors on the lattice, wzz is the energy of interaction between\ntwo polymer segments (species 2), m1 is the total number of voids (species 1), and (152 is the\npolymer volume fraction. This result may be compared to Equation 7.3.l2a for the enthalpy of\nmixing in the Flory\u2014Huggins theory, as z|w22]/2 plays the role of Add\". A key feature of m1 is that it\ncan vary with temperature; as T decreases, the energy penalty for having empty space plays an\nincreasingly important role, and so the material will contract. As m1 decreases, so will the entropy\nof placement of the chains on the shrinking lattice.\nThe second, crucial modification to the Flory\u2014Huggins approach is to assign different energies\nto various nearest neighbor conformations of the chain on the lattice. In the simplest case, an\nenergy 31 is assigned to one, lowest energy conformation, and 82 is assigned to other possibilities.\nThis is analogous to having one energy for a trans conformer, and a higher energy for either gauche\nplus or gauche minus states in polyethylene, as discussed in Chapter 6. As temperature decreases,\nthe chains will tend to adOpt more and more 81 conformations, which also serve to reduce the\nentropy of placement of the chains on the lattice.\nThe main calculation in this theory involves enumerating the number of ways the chains may\noccupy the lattice. This is similar in spirit to calculation of the entropy of the Flory\u2014Huggins theory\nin Section 7.3.2, but is complicated by including the different energies of each state (via both E,\nand A8 81 82) and by a more accurate accounting of the effects of chain ends. The inclusion of\nthe energy terms means that the result (called the partition function in statistical mechanics) can be\nused to \ufb01nd the free energy, and not just the entropy. The central consequence is the emergence of\n"]], ["block_3", ["are discontinuous, so this represents a second-order transition.\nThe results of this calculation exhibit several features that are in general agreement with the\ncharacteristics of the experimental glass transition (the corresponding experimental dependences\nwill be discussed in Section 12.6):\n"]], ["block_4", ["T2 increases with degree of crosslinking.\n9\u201854\"???)\n"]], ["block_5", ["volume\u201d to be discussed in Section 12.4. There is an energy associated with the voids, because a\npolymer segment adjacent to a void will have lost the interaction energy that it might have had with\nanother segment. In the notation of Section 7.2 and Section 7.3, the total energy associated with the\nvoids Ev will be given as\n"]], ["block_6", ["a temperature, T2, at which the number of possible states shrinks to l, and thus where the entropy\nvanishes. The value of T2 depends only on N, E,, and A8. The calculation shows the free energy,\nentropy, and volume to all be continuous at T2, whereas the thermal expansivity and heat capacity\n"]], ["block_7", ["Thermodynamic Aspects of the Glass Transition\n473\n"]], ["block_8", ["1.\nT2 increases with Ev, just as the experimental Tg tends to increase with cohesive energy density\n(recall Equation 7.6.5 to relate the cohesive energy density to W22).\nT2 increases with A3, just as Tg tends to increase with chain stiffness.\n"]], ["block_9", ["ZW22\ntam |2| \u201d11\u2018352\n(12.2.12)\n"]], ["block_10", ["T2 increases with N at low N, but approaches a limiting value as N \u2014> 00.\n"]], ["block_11", ["T2 depends on the number average molecular weight for polydisperse samples.\n"]], ["block_12", ["T2 decreases with added solvent.\n"]]], "page_485": [["block_0", [{"image_0": "485_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "485_1.png", "coords": [27, 332, 189, 648], "fig_type": "figure"}]], ["block_2", ["Figure 12.4\nSchematic illustration of a dilatometer. The polymer is in bulb A, the height of the polymer plus\nHg is determined in capillary B, and C contains extra Hg. The cell can be sealed at D and E. G is a calibration\ncapillary sealed at F. (Reproduced from Sperling, L.H., Physical Polymer Science, Wiley, New York, 1986.\nWith permission.)\n"]], ["block_3", ["This is about as unglamorous an experiment as one can imagine. As a prOperty of matter, we take\ndensity very much for granted. The fact that it is conceptually simple, readily accessible in\nhandbooks for many materials, and relatively monotonous in its variations all contribute to this\nattitude. Yet the phenomena represented schematically in Figure 12.1 require careful experimen\u2014\ntation on well\u2014defined samples to yield reproducible results. The device that is used to follow\nvolume changes with temperature is called a dilatometer; an example is shown in Figure 12.4. The\n"]], ["block_4", ["In this section we brie\ufb02y describe three approaches to characterizing the glass transition in polymers.\nThere are, in fact, many other possible experimental probes, but these three represent the most\ncommonly employed and also serve to illustrate several of the important aSpects of the glass\ntransition. The first is to measure the density or volume directly as a function of temperature, by\ndilatometiy; this ties directly to the introduction to the glass transition in Section 12.1. The second is to\nuse di\ufb01\u2018erential scanning calorimetry (DSC) to determine the heat capacity versus temperature. This\nis probably the most commonly employed method, as it combines speed, ease of use, and potentially\nquantitative thermodynamic information. The third method is mechanical analysis, which is nothing\nmore than a particular application of the viscoelastic properties described in Chapter 11.\n"]], ["block_5", ["is unlikely to provide a universal description of the glass transition. However, a successful,\nuniversal description of the glass transition has not yet been achieved even 40 years after the\nGibbs\u2014DiMarzio theory, so the successes of this theory should not be taken lightly.\n"]], ["block_6", ["474\nGlass Transition\n"]], ["block_7", ["12.3.1\nDilatometry\n"]], ["block_8", ["12.3\nLocating the Glass Transition Temperature\n"]], ["block_9", ["Vacuum\ni\ni\n"]], ["block_10", ["0\n"]], ["block_11", [{"image_2": "485_2.png", "coords": [81, 368, 155, 619], "fig_type": "figure"}]]], "page_486": [["block_0", [{"image_0": "486_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "486_1.png", "coords": [30, 275, 338, 661], "fig_type": "figure"}]], ["block_2", ["sample is placed in a bulb that is then filled with an inert liquid, generally mercury. Mercury is\nsuitable because it has negligible solubility in most polymers, and because it does not undergo any\ntransitions of its own over the relevant temperature range. The bulb is connected to a capillary\nso that changes in volume register as variations in the height of the mercury column, just as in a\nthermometer. For a constant temperature experiment, say, monitoring crystallization at Tm, the\nvolume changes in the capillary correspond identically to changes occurring in the sample. When\ntemperature variation is involved, the expansion of the mercury due to the temperature change is\nsuperimposed on the expansion of the specimen and must be taken into account. To obtain\nmeaningful results it is necessary to standardize the rate at which temperature changes are made\nand, of course, to have an accurately measured and uniform temperature in the bath surrounding\nthe dilatometer. The sample must be degassed to prevent entrapment of air; a gas bubble can really\nraise havoc in this kind of experiment. This experimental protocol favors slow rates of heating and\ncooling, to allow for temperature stabilization throughout the bath. For this reason, measurements\nduring cooling are relatively straightforward to make. This turns out not to be the usual case in\nother common methods where measurements during heating are the norm. Measurements on\ncooling have one fundamental advantage, namely, the sample is initially at equilibrium, and Tg\nrepresents the first observable departure from equilibrium. Experiments conducted during heating\nbegin with a nonequilibrium sample, and therefore the results generally depend on sample history.\nAn example of dilatometric data is provided in Figure 12.5, taken from the classic study of\nKovacs [3]. Two sets of speci\ufb01c volume versus temperature data for poly(vinyl acetate) are shown.\n"]], ["block_3", ["Figure 12.5\nSpecific volume versus temperature for poly(vinyl acetate), measured at 0.2 and 100 h after\ncooling rapidly from well above the glass transition temperature. (Reproduced from Kovacs, A.J., J. Polym.\nSal, 30, 131, 1958. With permission.)\n"]], ["block_4", ["Locating the Glass Transition Temperature\n475\n"]], ["block_5", ["(cm3/g)\n"]], ["block_6", ["100\n"]], ["block_7", ["vsp\nx\n"]], ["block_8", [{"image_2": "486_2.png", "coords": [47, 411, 240, 620], "fig_type": "figure"}]], ["block_9", ["85\n"]], ["block_10", ["03 .5\n"]], ["block_11", ["83\n\u201425\n0\n25\n50\nTemperature (\u00b0C)\n"]]], "page_487": [["block_0", [{"image_0": "487_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "487_1.png", "coords": [29, 464, 232, 625], "fig_type": "figure"}]], ["block_2", ["Figure 12.6\nSchematic illustration of heat \ufb02ow into the material versus temperature, showing the glass\ntransition, crystallization, and melting.\n"]], ["block_3", ["Differential Scanning Calorimetry (DSC) is a common example of a thermal analysis method.\nA small quantity of the sample is con\ufb01ned within an aluminum pan and subjected to controlled\ntemperature variation. A reference material is placed in an equivalent pan and the two pans are\nheated simultaneously. The temperatures of the two pans are monitored continuously and the rate\nof heat \ufb02owing to the sample is adjusted to keep the temperatures of the two pans equal. The heat\n\ufb02ow is proportional to an electrical current in a resistive heating element, so it is straightforward\nboth to control and to monitor. Whenever the sample undergoes a thermal transition, so that there\nis a change in heat capacity, the DSC registers both the amount and the direction of the additional\nheat \ufb02ow. A schematic example is shown in Figure 12.6. The first feature upon heating is the\nincrease in heat \ufb02ow required by the increase in heat capacity at Tg. This would be a step function\nfor a genuine second-order transition examined at very slow rates of heating. The second feature\nindicated is an exothermic peak, indicating that some of the material has crystallized on heating, at\na crystallization temperature To. This often occurs in crystallizable polymers because insufficient\ntime was allowed on cooling for extensive crystallization; but once the glass transition is traversed,\nchain segments acquire enough mobility to crystallize. As will be discussed in Chapter 13, the area\nunder this peak is prOportional to the amount of crystallized material if the enthalpy of fusion is\nknown and the DSC has been calibrated with a standard. The final feature in the DSC trace is an\nendothermic peak, corresponding to the melting of all crystalline material in the sample near the\nmelting temperature, Tm.\nFigure 12.7 shows an experimental DSC trace for a polystyrene sample with molecular weight\n13,000. In this case there are no peaks associated with crystallization, because the polymer is\n"]], ["block_4", ["In each case the measurements were taken after a direct temperature quench from an initial T >> Tg,\nThe upper curve corresponds to measurements taken 0.02 h after the quench, and the second set\n100 h later. The two traces closely resemble the schematic diagram in Figure 12.1 and in particular\nthe longer time data lead to a lower value of Tg, as expected. However, there is an important\ndistinction to be noted. The 100 h data were acquired not upon cooling more slowly, but rather\nafter waiting for a longer time after cooling below Tg. The process of isothermal volumetric\ncontraction with time after cooling below Tg is called aging. It can play an important role in the\nlongtime stability of the mechanical properties of glassy polymers, because densification can lead\nto undesirable changes in dimensions, increases in brittleness, and even failure.\n"]], ["block_5", ["476\nGlass Transition\n"]], ["block_6", ["12.3.2\nCalorimetry\n"]], ["block_7", ["flow\n"]], ["block_8", ["Heat\n"]], ["block_9", ["1\u201d\u201c?!\n\\\n"]], ["block_10", ["Temperature (\u00b0C)\n"]], ["block_11", ["TC\n"]]], "page_488": [["block_0", [{"image_0": "488_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "488_1.png", "coords": [33, 46, 288, 250], "fig_type": "figure"}]], ["block_2", ["atactic. There is a peak associated with T3, which, however, is often observed. It is sometimes\ncalled an \u201centhalpy overshoot,\u201d and is attributed to a \u201csuperheating\u201d of the glassy state, i.e., as\ntemperature increases and the enthalpy increases along the glassy branch, it crosses the equilibrium\nline at Tg but does not recover immediately. Consequently there is an extra increment of enthalpy\nrequired, beyond that dictated by Tg; this leads to a more rapid change with time in the enthalpy,\nand thus an overshoot in the DSC trace. Another notable feature of the curve in Figure 12.7 is that\nthe transition itself extends over an interval of 15\u00b0C\u201420\u00b0C, which is quite typical. There are various\nconventions for extracting a single value of Tg from such a trace, but the most common is to take\nthe midpoint, as indicated in the figure.\nAs alluded to above, the DSC measurement is generally made on heating, and therefore the\nsample begins in a nonequilibrium state. If we return for a moment to Figure 12.1, and suppose\nthat the sample has just been cooled along path ABJK. If immediately reheated, this path would\nbe retraced, giving T; as the result. Suppose, however, the sample sat overnight at the temperature\ncorresponding to point K. The volume would actually contract towards the extrapolation of the\nequilibrium curve AB, as illustrated by the volumetric data in Figure 12.5. We can see that if\nthe sample had contracted as far as point I in Figure 12.1, then on reheating it would follow the\npath IHBA, and give a different (lower) value of Tg. In general, then, the observed value of Tg in a\nheating experiment will depend not only on the heating rate, but also on how long the sample was\nheld below Tg, and also at what temperature it was held. In this way one could obtain many\ndifferent values of Tg from a single sample, even without varying the heating rate. To avoid this\ncomplication, the standard measurement protocol is as follows. The sample is loaded, usually at\nroom temperature, and is then heated above any suspected transition. A few minutes at elevated\ntemperature are sufficient to anneal the sample, i.e., erase all memory of past thermal history\nand stress. The sample is then rapidly cooled or quenched to the beginning temperature, and a\nsecond heating scan begun immediately. This second scan is taken as the measurement; a typical\nheating rate is 10\u00b0C/min. Examination of Figure 12.1 and Figure 12.5 suggests that this protocol\nwill tend to give relatively high values of Tg because of the rapid cooling, but often the more\nimportant issue is to make the result reproducible.\n"]], ["block_3", ["Figure 12.7\nCalorimetric determination of Tg for a polystyrene with M 13,000 at a heating rate of 10\u00b0C/\nmin. (Data described in Milhaupt, J.M., Lodge, T.P., Smith, S.D., and Hamersky, M.W., Macromolecules, 34,\n5561, 2001.)\n"]], ["block_4", ["Locating the Glass Transition Temperature\n477\n"]], ["block_5", ["flow\n"]], ["block_6", ["Heat\n"]], ["block_7", ["ub>uub>b>2ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>cub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>aub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>.\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>1ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>aub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>Iub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>ub>ub>nub>ub>ub>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>ub>uub>b>\nub>eub>\nub>uub>\n"]], ["block_8", ["40\n60\n80\n100\n120\n140\n160\nTemperature (\u00b0C)\n"]]], "page_489": [["block_0", [{"image_0": "489_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 12.8\nThe dynamic moduli G\u2019 and G\u201d at 1 Hz, measured as a function of increasing temperature, for a\npoly(styrene-ran\u2014butadiene) copolymer. (Reproduced from Nielsen, L.E., Mechanical Properties of Poly\u2014\nmers, Reinhold, New York, 1962. With permission.)\n"]], ["block_2", ["As a polymer sample is cooled through Tg, the motions of individual segments undergo a dramatic\nslowing down. Consequently, any experimental measurement of such local relaxation times should\nbe sensitive to Tg. As commercial rheometers are widely available, measurements of G\u2019 and G\u201d\nversus temperature are commonly employed to locate Tg. An example of the dynamic moduli\nmeasured versus temperature at a fixed frequency of\n1 Hz is shown in Figure 12.8. Note the\nsimilarity in shape between this G\u2019 curve and those in Figure 11.2 and Figure 11.16b for G(t).\nThe former re\ufb02ects the modulus at fixed frequency (or time) as a function of temperature, whereas\nthe latter shows the variation with time at fixed temperature. The origin of this time\u2014temperature\nequivalence will be explored in Section 12.5, but for now we can consider it intuitively. Below T ,\nall relaxation modes are frozen, so the material behaves as a solid with a very high modulus. Upon\nheating, all relaxation processes accelerate, and when a process can occur in 1 s or less, it will show\nup as a relaxation if the modulus measurement is made at about\n1 Hz. The first processes to\nbecome fast enough upon heating a glass correspond to motions of small pieces of chain, in what\nwe called the transition zone of viscoelastic response in Chapter 11. Gradually, on further heating\nwe enter the rubbery plateau, and, eventually, at temperatures far above Tg the molecules can fully\nrelax and the material can flow within\n1 s, and the moduli enter the terminal regime. During the\ntransition zone G\u201d exhibits a peak at a particular temperature (as does the loss tangent, tan 5, given\n"]], ["block_3", ["478\nGlass Transition\n"]], ["block_4", ["12.3.3\nDynamic Mechanical Analysis\n"]], ["block_5", ["(dyn/cme)\n"]], ["block_6", ["Shear\n"]], ["block_7", ["modulus\n"]], ["block_8", [{"image_1": "489_1.png", "coords": [49, 275, 310, 617], "fig_type": "figure"}]], ["block_9", ["107\n"]], ["block_10", ["106\nl\nI\n_|\nI\nJ\nT\nF\u2014\n"]], ["block_11", ["109\n"]], ["block_12", ["108\n"]], ["block_13", ["\u2014-20\n0\n20\n40\n"]], ["block_14", ["Temperature (\u00b0C)\n"]]], "page_490": [["block_0", [{"image_0": "490_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1 rad/s and 2\u00b0C/min. (Data described in Milhaupt, J.M., Lodge, T.P., Smith, S.D., and Hamersky, M.W.,\nMacromolecules, 34, 5561, 2001.)\n"]], ["block_2", ["In this section we set aside the issue of thermodynamic equilibrium and simply consider why\nthe viscosity, or other dynamic and mechanical properties of polymers, should undergo a dramatic\nchange over a relatively narrow range of temperatures, even though the structure of the material\nremains liquid\u2014like. The essence of the argument is based on the concept of \ufb01'ee volume, which\nis intuitively very appealing although difficult to pin down precisely. The actual volume of a\nsample can be written as the sum of the volume \u201coccupied\u201d by the molecules (subscript occ)\nand the free volume (subscript f). Acknowledging that each of these is a function of temperature,\nwe write\n"]], ["block_3", ["The variation in VOCC with T arises from changes in the amplitude of molecular vibrations with\nchanging T, a variation that affects the excluded volume of the molecules. The free volume, on the\nother hand, may be viewed as the \u201celbow room\u201d between molecules, and is required for molecules\nto undergo rotation and translational motion. As the average kinetic energy increases with\nincreasing temperature, so the associated free volume is also expected to increase with T. Free\nvolume plays a similar role to the \u201cvoids\u201d in the Gibbs\u2014DiMarzio theory.\n"]], ["block_4", ["by G\u201d/G\u2019). By analogy with the Maxwell model in Section 11.2, this means that the frequency\nequals an inverse relaxation time, so the location of this peak corresponds to the temperature at\nwhich segmental motion occurs in 1 s. This can be taken as an alternative, empirical definition of\nTg. Figure 12.9 shows an example for the same polystyrene sample as in Figure 12.6, with the data\nobtained while heating at 2\u00b0C/min. The values of Tg obtained by the two methods are quite\ncomparable (although there is no fundamental reason why they should agree to better than\nabout 10\u00b0C).\n"]], ["block_5", ["Free Volume Description of the Glass Transition\n479\n"]], ["block_6", ["Figure 12.9\nDetermination of T3 for polystyrene with M = 13,000 from the maximum in G\" obtained at\n"]], ["block_7", ["12.4\nFree Volume Description of the Glass Transition\n"]], ["block_8", ["g\n0.6\n'-\n:\n.\n_\n0\n.\n'.\n.\n"]], ["block_9", ["=E\nI\no\nd\n"]], ["block_10", [{"image_1": "490_1.png", "coords": [36, 47, 316, 261], "fig_type": "figure"}]], ["block_11", ["\u00a7\n0\n.\n"]], ["block_12", ["V(T) Vocc(T) + Via\")\n(12-4-1)\n"]], ["block_13", ["0.2\n'1'\u201d\nI\n.\n:3\n\u2018\nI\n"]], ["block_14", ["-\n- -\n1\n08 \nTg=104.5\u00b0c\n.\no\nA\n'\n_\no\nO\n.\n-\n"]], ["block_15", ["_\nI\nO\n4\n0.4 -\na\n.4\n:\n'-\n'\n"]], ["block_16", ["0\n"]], ["block_17", ["1\n-\n\ufb01\n_\n3 .\n1\n"]], ["block_18", ["70\n80\n90\n100\n110\n120\n130\n"]], ["block_19", ["a\nU\n"]], ["block_20", ["I\n"]], ["block_21", ["\u2018l\nl\nI\nI\n\u2018I_l\u2014l\nI\nI\nl\u2014I_r\n\"I\nl'\nI\n\"I\nl\u2019\nI\nI\nI\n\"I\nr\nT\nI\ni\nI\nT\n"]], ["block_22", ["l\nl\nI\n.l\nI.\n.l\nl\nl\nI\nl\nI\nI\nI\nI\nI\nI\nI.\nt\nl\nl\nI\nI\n"]], ["block_23", [{"image_2": "490_2.png", "coords": [151, 65, 292, 244], "fig_type": "figure"}]], ["block_24", ["T(\u00b0C)\n"]], ["block_25", ["-\n"]]], "page_491": [["block_0", [{"image_0": "491_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "491_1.png", "coords": [27, 418, 286, 616], "fig_type": "figure"}]], ["block_2", ["Figure 12.10\nSchematic representation of the actual volume (solid line) and occupied volume (dashed line)\nversus temperature. The hatched area represents the free volume.\n"]], ["block_3", ["The subscripts g and\n1 on the coefficient of volume variation with T indicate that these are\ndetermined for the glass and liquid states, respectively. Each of the differences used to describe\nthe second-order phase transition as given by Equation\n12.2.9 or Equation 12.2.10 can be\n"]], ["block_4", ["2.\nAt T,\n"]], ["block_5", ["These concepts may be represented schematically as shown in Figure 12.10 where V is plotted\nagainst T. The solid line represents the actual volume, which shows a change in slope at T,. The\ndashed line indicates the increase in VOCC with T. In all cases the lines are straight, i.e., a linear\ndependence of V on T, which is a very reasonable approximation for our purposes. The shaded area\nrepresents the free volume, Vf. The key feature of Vf is that it decreases upon cooling from the\nliquid state, until at Tg it reaches some critical, small value. Physically the idea is that once Vf\nbecomes too small, molecular rearrangements are effectively frozen out and the system can no\nlonger continue contracting. Below T,, V and V0,, have approximately the same T dependence and\nso Vf becomes roughly independent of T.\nOn the basis of these ideas, we can write the following expression for the volume of the sample:\n"]], ["block_6", ["3.\nAbove T,\n"]], ["block_7", ["430\nGlass Transition\n"]], ["block_8", ["12.4.1\nTemperature Dependence of the Free Volume\n"]], ["block_9", ["v\n"]], ["block_10", ["1.\nBelow T,\n"]], ["block_11", [">\nFree volume\n"]], ["block_12", [{"image_2": "491_2.png", "coords": [50, 325, 258, 360], "fig_type": "molecule"}]], ["block_13", [{"image_3": "491_3.png", "coords": [52, 217, 281, 252], "fig_type": "molecule"}]], ["block_14", [{"image_4": "491_4.png", "coords": [52, 270, 263, 306], "fig_type": "molecule"}]], ["block_15", ["d V,\nv\nV(T > T,) = V(T,) + (34:?) (T T,)\n(12.4.4)\n"]], ["block_16", ["dvocc\nV(T < T,) V0,,(T 0) 4 W0\u201c 0) +\n\u2014dT_\nT\n(12.4.2)\n"]], ["block_17", ["dVocc\nV(T,) V,.,, (T 0) + Vf(T 0) +\n(H\nT,\n(124.3)\n"]], ["block_18", ["\\\\\\\\\\\\\\\\\\\n~\n\\\\\n"]], ["block_19", ["Tg\nT\n"]], ["block_20", ["l\n"]], ["block_21", ["I\nl\nI\n"]], ["block_22", ["I\nI\nl\nI\nl\nl\nl\nI\n"]], ["block_23", ["1\n"]], ["block_24", ["g\n"]], ["block_25", ["8\n"]]], "page_492": [["block_0", [{"image_0": "492_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["qualitatively traced to the sort of change implied by Figure 12.10. Around Tg, for example, the\ncoefficients of expansion of the liquid and glassy states are, respectively,\n"]], ["block_2", ["The preceding introduction to the free volume interpretation of the glass transition provides a\nqualitative picture, but is not immediately of much utility because, as discussed in the preceding\nsection, volume is not the generally measured quantity. Furthermore, even if it were we would be\nfaced with the tricky problem of resolving the measured V into the contributions from Vf and VOCC.\nInstead, we shall tum our attention to the steady \ufb02ow viscosity, 1), which will serve many purposes.\nFirst, it will illustrate the dramatic effect approaching Tg can have on the \ufb02ow properties. Second, it\nwill show how powerful the free-volume concept can be in describing experiments. Third, it will lead\nus naturally to the principle of time\u2014temperature superposition, which forms the topic of Section 12.5.\n"]], ["block_3", ["where A is a prefactor with units of viscosity, and E, is the activation energy. In this view, the\nlimiting process to flow is the energetic barrier to molecules sliding past one another. For toluene at\nroom temperature, for example, E, is approximately 9 kJ/mol. (This may be compared with the\nenergy of a hydrogen bond, approximately 15\u201420 kJ/mol, or a carbon\u2014carbon bond in polyethylene,\nabout 350 kJ/mol. These numbers indicate that there is little energetic resistance to flow in a small\nmolecule \ufb02uid.) If the viscosity of polystyrene followed Equation 12.4.7, then the data would fall on\n"]], ["block_4", ["a straight line when plotted as log 7; versus l/T, as in Figure 12.1 lb; clearly they do not. One can take\nthe slope of a small portion of the curve, and extract an apparent 15,. This gives a value almost 10\ntimes greater than toluene in the high temperature range, and 60 times higher near Tg (see Problem\n12.8). The local energy of interaction between styrene monomers cannot be too different from that\nbetween toluene molecules, so this factor of 60 must have a different origin. It is not a simple result of\nmolecular weight, because the T dependence of n is found to be more or less independent of chain\nlength (see Section 12.5). The reason is that \ufb02ow in glass-forming liquids is impeded primarily by a\nlack of free volume, rather than by an energy barrier (although such barriers should still contribute).\nDoolittle [4] studied the viscosity of n\u2014alkanes in detail and found that the following equation\nwas able to describe the data:\n"]], ["block_5", ["We begin with some experimental results. Figure 12.1 1a and Figure 12.1 lb show the viscosity of an\noligomeric polystyrene as a function of temperature, from about 180\u00b0C down to 375\u00b0C (the nominal\nTg for this sample). The most remarkable feature of the data is the smooth, 12 orders of magnitude\nvariation in 11 over an otherwise unremarkable temperature interval. This behavior is actually\ntypical of all polymers as Tg is approached from above, although 7) is rarely measured over such a\nwide range. For most small molecule \ufb02uids we would expect an Arrhenius temperature dependence:\n"]], ["block_6", ["and therefore A0: a! org is a measure of the \u201copening up\u201d of V1: at Tg. The additional volume\nabove Tg also accounts for the increase in compressibility AK, and the emergence of the associated\nmodes of energy storage. i.e., additional degrees of motional freedom, accounts for ACP.\n"]], ["block_7", ["Free Volume Description of the Glass Transition\n481\n"]], ["block_8", ["12.4.2\nFree Volume Changes Inferred from the Viscosity\n"]], ["block_9", [{"image_1": "492_1.png", "coords": [36, 82, 149, 153], "fig_type": "molecule"}]], ["block_10", [{"image_2": "492_2.png", "coords": [38, 635, 147, 671], "fig_type": "molecule"}]], ["block_11", ["#\n1\nd(Vocc + V,)\non \n"]], ["block_12", ["_\n1\ncit/0,,\n03g \nVg\n(\n(17\u2018 1\n(12.4.6)\n"]], ["block_13", ["n(T) A exp (1%,)\n(12.4.7)\n"]], ["block_14", ["1? A\u2019exp(\nm)\n(12.4.8)\nVr\n"]], ["block_15", ["Vg\n(\n(1T\n1\n(12.4.5)\n"]]], "page_493": [["block_0", [{"image_0": "493_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "493_1.png", "coords": [30, 40, 329, 313], "fig_type": "figure"}]], ["block_2", ["482\nGlass Transition\n"]], ["block_3", ["(a)\nT(\u00b0C)\n"]], ["block_4", ["Figure 12.11\nViscosity of oligomeric polystyrene (M 1100) plotted (a) versus temperature and (b) as log\nviscosity versus inverse temperature. The smooth curve in (b) represents the fit to the VFTH or WLF\n"]], ["block_5", ["equations. (Data described in Plazek, DJ. and O\u2019Rourke, V.M., J. Polym. Sci. A2, 9, 209, 1971.)\n"]], ["block_6", ["(b)\n1/T(K\u201c)\n"]], ["block_7", [",... 107\nE\n.\ng\nEL\n:\n:\n\"\u2019\n3\nM =11\n:\n:-\n5\n.\n00\n'5\n5\no\nE\n105\nE\u2018\n\u201c3;;\n:\n.\n_\n"]], ["block_8", ["E\"\nM = 1100\n1.\n1;\nE \n2'\n\ufb02\n2\nE.\u201c\n-\nI\u2018\n"]], ["block_9", [{"image_2": "493_2.png", "coords": [37, 337, 313, 590], "fig_type": "figure"}]], ["block_10", ["1013\nl;\n"]], ["block_11", ["1011\nQ-\n.\n1:\n"]], ["block_12", [":\no\n\u2014;\nI\n.\n:\n9\n=-\n\u2014-I_\n10\n.\n"]], ["block_13", ["3\nE\n.\n:\n10\n1\nE:\n0\ng\n"]], ["block_14", ["10\u201c\nE.-\n,'\ng\n:\nr\n:\n"]], ["block_15", ["E\nO .\nf\n\u20180\u2018\n. .\n"]], ["block_16", ["10\u20141\n"]], ["block_17", ["1013\n=\nI\nl\n|\n|\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\u2014\nI\nI\n5\nE\na\n,'\n1:\n"]], ["block_18", ["103\n;\u2014\nx.\ng\ng\nI.\n:\n"]], ["block_19", ["o\n105\nE-\nI/\n_\u00a7\n:\nx.\n:\n"]], ["block_20", ["101\nE-\n\u201c\n-\u00a7\n"]], ["block_21", ["109\n\u00a3-\nPolystyrene\n\u2018.\n\u00e9\n"]], ["block_22", ["10\u20141\n-I\nI\nI\nI\nI\nI\nI\nI\nl\nl\nI\nI\n|\nI\n|\nI\nI\nl\nI\nI\nI\nI\nI\nI\nI\nI\nl\u2014\n"]], ["block_23", ["2 10-3\n2410-3\n2.810\u20183\n3.210-3\n"]], ["block_24", ["\u00e9-\n\u00b0\n1;\n"]], ["block_25", ["E'\n.\nPolystyrene\n\"g\n"]], ["block_26", ["E\n\u201ca\n"]], ["block_27", ["2\u2019\n\u00b0\n0\n\"2\n"]], ["block_28", ["-\n"]], ["block_29", [":\u2014\n0\nE\n"]], ["block_30", ["if\ni\"\n\u2018E\n"]], ["block_31", [":\nI\nE\n"]], ["block_32", ["\u00a7\n.,0.\ng\n"]], ["block_33", ["2'\n0\u2019\n1;\n"]], ["block_34", ["E\"\nO\nE\n\u20184\u2019\n"]], ["block_35", ["|\nI\nI\nI\nI\nI\nI\nI\nI\nl\nI\nI\nI\n|\nI\n1\nE\n"]], ["block_36", ["I\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nl\nl\n|\n-\n"]], ["block_37", ["40\n80\n120\n160\n200\n"]], ["block_38", [{"image_3": "493_3.png", "coords": [123, 167, 303, 283], "fig_type": "figure"}]], ["block_39", [{"image_4": "493_4.png", "coords": [130, 61, 297, 289], "fig_type": "figure"}]], ["block_40", ["_\n"]]], "page_494": [["block_0", [{"image_0": "494_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where the new parameters are B :B\u2019/af and To :Tg \u2014fg/a:f. The last form of Equation 12.4.12 is\noften known as the Vogel\u2014Fulcher\u2014Tammann\u2014Hesse (VFTH) equation and the parameter To is\nreferred to as the Vogel temperature [5]. By comparison with Equation 12.4.7, we can see that the\nVFTH equation reduces to the Arrhenius equation when To 0 K. However, experimentally one\nfinds that To > 0, and therefore the viscosity is predicted to become infinite at T0, or some fg/af\ndegrees below Tg.\nThe data in Figure 12.1 1 have been fit to the VFTH equation, resulting in the smooth curve shown\nin Figure 12.11b. Clearly the data follow this form very well. The resulting parameter values are\nA :5.1 x 10 P, B 1743 K, and To 265 K. Thus in this case To is 450 below Tg, which turns out to\nbe quite typical. This is intriguing, because the Vogel temperature is therefore rather close to the\nKauzmann temperature discussed in Section 12.2; To, based purely on a semiempirical fit to a\ndynamic property, lends some support to the concept of TK, a quantity anticipated on thermodynamic\ngrounds. Furthermore, recalling the derivation above, we can equate 45\u00b0 with the ratio fg/af. Given\nthat af should be close to Ace, and that this quantity is on the order of 5 x 10 DC l, the implication is\nthatfg % 0.023. It is sometimes suggested that the glass transition corresponds to a particular value of\nthe fractional free volume. We will consider this issue further later in this section.\n"]], ["block_2", ["where A\u2019 and B\u2019 are empirical constants. Although they relate 77 to different variables, both the\nDoolittle and Arrhenius equations have the same functional form. Just as the activation energy 1'3,l\nrepresents the height of a barrier relative to thermal energy, RT, the Doolittle equation compares\nthe space needed for a molecule, VOCC, to the space available, Vf. In a sense Equation 12.4.8 implies\nthat the primary impediment to molecular motion is entropic in nature (is there space available to\nmove?) rather than energetic as in Equation 12.4.7 (is there enough thermal energy to overcome\nthe activation barrier?). As we shall see shortly, the Doolittle equation describes the remarkable T\ndependence of 17 shown in Figure 12.11 very well.\nTo proceed, we define the fractional free volume, f: Vf/V, and then Equation 12.4.8 can be\nwritten as\n"]], ["block_3", ["simply by incorporating a factor of exp(\u2014B\u2019) into A. To proceed further, we need to insert an\nexpression forf, particularly in terms of its T dependence. On the basis of Figure 12.10, we propose\na linear dependence forf(T), for temperatures at and above Tg:\n"]], ["block_4", ["The VFTH equation is capable of describing rather well the temperature dependence of n or, in\nfact, of any relaxation time of a polymer chain, based on three parameters. It is often the case that\n"]], ["block_5", ["We can simplify this expression as follows:\n"]], ["block_6", ["In this expression (If is the coefficient of expansion of the free volume only, but it should be close\nto Aar. Equation 12.4.11 can be inserted into Equation 12.4.10 as follows:\n"]], ["block_7", [{"image_1": "494_1.png", "coords": [35, 174, 146, 215], "fig_type": "molecule"}]], ["block_8", ["Free Volume Description of the Glass Transition\n483\n"]], ["block_9", [{"image_2": "494_2.png", "coords": [35, 356, 240, 429], "fig_type": "molecule"}]], ["block_10", ["12.4.3\nWilliams-Landel-Ferry Equation\n"]], ["block_11", [{"image_3": "494_3.png", "coords": [39, 229, 204, 261], "fig_type": "molecule"}]], ["block_12", [{"image_4": "494_4.png", "coords": [44, 377, 221, 419], "fig_type": "molecule"}]], ["block_13", ["fg+af(T_Tg)\n(12.4.11)\n"]], ["block_14", ["1\nn = A'exp[B\u2019 (IT \u2014 1)]\n(12.4.9)\n"]], ["block_15", ["B\u2019\nB\u2019\n77 =A\u2019exp(}\u2014\u2014B\u2019) =Aexp(}~)\n(12.4.10)\n"]], ["block_16", ["Bf\nerratum\u2014Ts)\nBf/Otf\nB\nexPSource: Data collected in Ferry, J.D., Viscoelosric Properties of Polymers, 3rd ed., Wiley, New York, 1980; Sperling, L.H-,\nPhysical Polymer Science, Wiley, New York, 1986.\n"]], ["block_5", ["Poly(dimethylsiloxane)\n\u2014 130\n8.5\n4.5\n0.071\n6.1\n69\n1,4\u2014Polybutadiene\n\u2014 95\n7.8\n2.0\n0.039\n11.2\n60.5\nPolyisobutylene\n\u2014 75\n6.2\n1.5\n0.026\n16.6\n104\n1,4\u2014Polyisoprene\n\u2014- 70\n6.2\n2.1\n0.021\n10.8\n51.1\nPoly(viny1 acetate)\n30\n6.0\n2.1\n0.028\n15.6\n46.8\nPolystyrene\n100\n5.5\n1.8\n0.032\n13.7\n50\nPoly(methy1 methacrylate)\n110\n4.6\n2.2\n0.013\n34\n80\n"]], ["block_6", ["Table 12.2\nRepresentative Values of VFTH/WLF Equation Parameters for Various Polymers\n"]], ["block_7", [{"image_3": "495_3.png", "coords": [35, 164, 143, 207], "fig_type": "molecule"}]], ["block_8", ["a in units of 10*4 \u00b0C' 1.\n"]], ["block_9", ["Polymer\nTg (\u00b0C)\na?\n01;\nf3\nCi\u201d\nCE (\u00b0C)\n"]], ["block_10", ["one has measurements over a range of temperatures, and it is convenient to choose some tempera\u2014\nture as a reference. The value of the measured quantity at the reference temperature, T,, can then be\nused to eliminate the front factor A from the VFTH equation, thereby reducing the number of\nparameters to two. This can be seen by taking the ratio of two values of r) with two different values\noff(subscript r denotes the property at T.) from Equation 12.4.10, remembering that B\u2019 is a constant:\n"]], ["block_11", ["OI'\n"]], ["block_12", ["We now convert to base 10 logarithms, and collect the various constants and rename them:\nn_\u201c_QU\u2014E)\n10g (\"7) \nC2 +(T Tr)\n(12.4.16)\n"]], ["block_13", ["484\nGlass Transition\n"]], ["block_14", ["Now we reintroduce the linear temperature dependence off as in Equation 12.4.11, except that we\nchoose Tr as the reference instead of Tg: f:fr+ of (T# Tr). This is legitimate as long as T, > Tg,\n"]], ["block_15", [{"image_4": "495_4.png", "coords": [37, 317, 172, 354], "fig_type": "molecule"}]], ["block_16", [{"image_5": "495_5.png", "coords": [47, 234, 217, 303], "fig_type": "molecule"}]], ["block_17", ["11\n1\n1\n1\n_.._ 2\nB\n_ _ a\n12.4.\nnr\nexpi:\n(f\nfr):i\n(\n13)\n"]], ["block_18", ["ln (\u201d3) =B\u2019 (11; \u2014%)\n(12.4.14)\n"]], ["block_19", ["IL\n_\n;\n1\n_l\nln(nr)_B(fr+af(T\u2018\u2014Tr)\nfr)\n"]], ["block_20", [":H(r4#\u00abmr\u2014nqzh\u00a3(:n#mr\u2014n))\nfr2 +fraf(T \n(fr/05f) + (T Tr)\n"]], ["block_21", ["(12.4.15)\n"]]], "page_496": [["block_0", [{"image_0": "496_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["From Chapter\n11 we recall that n~M\u00b0\u20184, so a 20% increase in MW increases\n77 by about\n(1.2)3'42 1.86. We can write two WLF equations (Equation 12.4.16) for the two temperatures\nT; 150\u00b0C and T2, the unknown, using the parameters from Table 12.2:\n"]], ["block_2", ["1.\nAlthough Equation 12.4.12 or Equation 12.4.16 describes the variation of viscosity over a wide\nrange of conditions quite well, they tend to break down both very far above Tg, where free\nvolume is not so important, and very close to Tg. In the latter range some workers advocate a\npower law relationship in (T To), where To is an adjustable critical temperature analogous to\nthe Vogel temperature; such power laws are the rule for the divergence of experimental\nquantities approaching a phase transition. This idea will be explored further in Problem 12.11.\n2.\nAlthough it is easy to discuss free volume, it is necessary to come up with a numerical value\nfor this quantity in order to test these concepts. There is considerable disagreement as to which\nof several different methods of computation gives the best value for Vf and Vocc, and its\nextraction from experimental data requires assuming a particular model.\n3.\nIt is possible to derive an expression equivalent to Equation 12.4.12 or Equation 12.4.16\nstarting from entropy rather than free\u2014volume concepts. We have emphasized the latter\napproach, since it is easier to visualize and hence to use for qualitative predictions about T3.\n4.\nCompletely aside from the theories that attempt to explain it, the empirical usefulness of the\nVFTH and WLF equations is beyond doubt. We shall examine this in detail in the next section.\n"]], ["block_3", ["A process for molding transparent plastic cups from polystyrene has been optimized to run at\n150\u00b0C. When the supplier of the raw material introduces a new resin with a 20% higher Mw, the\nincreased viscosity slows the process down. At what temperature should the process be run in order\nto recover the viscosity of the original raw material?\n"]], ["block_4", ["where the factor of 1.86 is inserted in the second equation in order to find the temperature T; at\nwhich the viscosity has been lowered by that amount. We now take the difference between these\ntwo equations to obtain\n"]], ["block_5", ["polymers, and the WLF coefficients C1 and C2 when Tg is taken as the reference temperature. The\nextracted value of fg varies significantly, and the sensitivity of the VFTH and WLF equations to\nthe choice of parameters is rather strong. Conversely, when fitting experimental data to extract\nthese parameters, the exact location of Tg plays an important role. And, although a, and org can be\nmeasured, the expansion of free volume af is not necessarily exactly equal to al ag, and one is\nstill left with B as an adjustable parameter in extracting a value for fg. Some further observations\nabout the free volume approach are listed below:\n"]], ["block_6", ["which can be solved to give T2 154\u00b0C. This calculation illustrates the remarkable in\ufb02uence of the\nglass transition on dynamics. In this instance, although the experimental temperature is about 50\u00b0\nabove Tg, only a 4\u00b0 increase in temperature is sufficient to cut the viscosity almost in half.\n"]], ["block_7", ["Solution\n"]], ["block_8", ["Free Volume Description of the Glass Transition\n485\n"]], ["block_9", ["Example 12.2\n"]], ["block_10", ["13.7(T2 100)\n.\n=\n.2\n= \u20146.\nlog(186)\n0\n69\n85 +\nT2 \n"]], ["block_11", ["log a z _ \n= _ = _6.85\n77s\nC5 + (T1 \n100\n"]], ["block_12", ["10g\n771\nz _\n= \n"]], ["block_13", [{"image_1": "496_1.png", "coords": [54, 467, 291, 541], "fig_type": "molecule"}]], ["block_14", ["186%\nC5 + (T2 \nT2 50\n"]]], "page_497": [["block_0", [{"image_0": "497_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "497_1.png", "coords": [28, 284, 199, 315], "fig_type": "molecule"}]], ["block_2", ["One of the main points of the previous section was that the viscosity of a polymer can change by many\norders of magnitude over only a few tens of degrees in temperature. This feature is routinely exploited\nto extend the range of measurement in any viscoelastic experiment. As an example, consider Figure\n11.16b, where the stress relaxation modulus of polyisobutylene at 25\u00b0C was plotted against time.\nNote that the time axis was actually day, where the significance of the factor aT will emerge shortly.\nThe range of the time axis extended from a few picoseconds to megaseconds\u2014nearly 18 orders of\nmagnitude. There are no rheometers capable of measuring the picosecond response of materials, or\neven the nanosecond response. Some custom instrumentation has been developed that can extend\ndown to microseconds, but that was not the case for these measurements. Similarly, at the long time\nend, the measurements extend to almost two weeks\u2014hardly experimentally convenient. What was\ndone in fact was to make measurements over a range of temperatures, and to reduce these to one\nmaster curve using the time\u2014temperature superposition (TTS) shift factor, aT. The underlying\nprinciple is one of corresponding states; a measurement at a certain temperature and time (or\nfrequency) is equivalent to a measurement at a lower temperature and longer time.\nThis correspondence can be understood quite simply. We begin with a generic expression for\nthe stress relaxation modulus, 00,7), which also happens to be consistent with the Rouse, Zimm,\nand reptation models (see Chapter 11):\n"]], ["block_3", ["The significance of Equation 12.5.3 is this: a measurement of G at a particular combination of (t,T)\nis exactly equivalent to a measurement at a new time, t/aT, and new temperature, Tr. (The front\nfactor involving p and T contributes a small correction that is often ignored in practice; note that p\ndecreases as T increases, so the net effect is even smaller than that due to p alone.)\nNow we can see how TTS is used in practice. The actual measurements of Catsiff and Tobolsky\n[7] are shown in Figure 12.12. They employed an instrument for which they could resolve\nmeasurements from a few seconds up to a few hours, and they varied the temperature from\n\u201480.8\u00b0C to 50\u00b0C. At the lowest temperatures the modulus is very high, characteristic of a\nglass. As temperature increases, the modulus decreases at fixed time. Near \u2014\u201420\u00b0C the modulus\nseems to be independent of time and the value is characteristic of a lightly crosslinked rubber.\nFinally, by 50\u00b0C the material \ufb02ows on the measurement timescale. The data may be shifted\n"]], ["block_4", ["The crucial assumption here is that all relaxation times have the same temperature dependence and\nthus that one value of (IT applies to any dynamic property including the viscosity. Now we insert\nthis definition into Equation 12.5.1 and rearrange:\n"]], ["block_5", ["486\nGlass Transition\n"]], ["block_6", ["where we have explicitly indicated the quantities that depend on temperature: density, p, and the\nrelaxation times, 7p. Of these, the dependence of p is rather weak; recall that the thermal expansion\nfactor is usually less than one tenth of 1% per degree. The relaxation times, however, follow the\nVFTH or WLF dependence derived in the previous section, and might change by an order of\nmagnitude over a few degrees (see Figure 12.11). Now we define the TTS shift factor, dry, as the\nratio of any relaxation time at one temperature to its value at the chosen reference temperature, T,\n"]], ["block_7", ["12.5\nTime\u2014Temperature Superposition\n"]], ["block_8", [{"image_2": "497_2.png", "coords": [37, 392, 138, 431], "fig_type": "molecule"}]], ["block_9", ["a \nT \"7pm\nncr.)\n"]], ["block_10", ["t\nG(t,T) :\nex\n\u2014\nM 2\np\n7pm\n"]], ["block_11", ["__ \nt\n__ \n_\nt\nG(t,T) \nM\nZexp(\u2014 Tp(T))\n\u2014-\nM\nZexp<\namTqrD\n"]], ["block_12", ["= \np(Tr)Tr\n"]], ["block_13", ["p ( T)RT\n ( \n"]], ["block_14", [{"image_3": "497_3.png", "coords": [111, 458, 346, 517], "fig_type": "molecule"}]], ["block_15", ["G(t/aT,T,)\n(12.5.3)\n"]], ["block_16", [{"image_4": "497_4.png", "coords": [153, 469, 337, 505], "fig_type": "molecule"}]], ["block_17", [")\n(12.5.1)\n"]], ["block_18", ["(12.5.2)\n"]]], "page_498": [["block_0", [{"image_0": "498_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "498_1.png", "coords": [27, 498, 158, 531], "fig_type": "molecule"}]], ["block_2", ["Time-Temperature Superposition\n487\n"]], ["block_3", ["horizontally by an arbitrary shift factor, until they overlap to produce a master curve; a reference\ntemperature of 25\u00b0C was selected. (A small vertical shift corresponding to the pa?\" front factor was\nalso applied.) The result is shown in Figure 12.13 (and previously as Figure\n11.16b). This\nempirical generation of a master curve produces a set of shift factors, aT, and we can ask how\n"]], ["block_4", ["Figure 12.12\nStress relaxation modulus for polyisobutylene at the indicated temperatures. Data are actually\nYoung\u2019s modulus, which is approximately equal to 3G(t) (see Section 10.3). (Reproduced from Catsiff, E. and\nTobolsky, A.V., J. Colioid Sci, 10, 375, 1955. With permission.)\n"]], ["block_5", ["The shift factors are used to generate the master curve in Figure 12.13 and plotted against\ntemperature in Figure 12.14, along with a fit to Equation 12.5.4. They follow the WLF equation\nvery well except for the lowest temperatures. The reason for this discrepancy is simple; some of the\ndata were obtained below Tgsa\u2014 680C. The WLF equation relies on a linear variation of free\nvolume with temperature, which should not hold when traversing Tg, as Figure 12.10 indicates.\nThe principle of TTS applies to any dynamic property, a generality that can be understood from\nthe discussion in Section 11.3, where we showed that any linear viscoelastic function is derivable\nfrom G(t). In particular, TTS is often used in dynamic experiments, and G\u2019 and G\" may be\nsuperposed to form master curves versus reduced frequency maT. This approach was used, for\nexample, to produce the dynamic moduli for poly(viny1 methyl ether) shown in Figure ll.l6a. In\nthis case the superposed data extend over almost 10 orders of magnitude in reduced frequency,\n"]], ["block_6", ["(21- depends on T. The answer should come as no surprise: it follows the WLF equation, as in\nEquation 12.4.16:\n"]], ["block_7", [{"image_2": "498_2.png", "coords": [36, 28, 270, 361], "fig_type": "figure"}]], ["block_8", ["g\n_49.60\n_65.4OC\n"]], ["block_9", ["'2\n\u20144OJ\u00b0Q\\\\HHHHHEH*\u2018:E\u00a7B%3\n>~\nk\n3;\nLu\n"]], ["block_10", ["C1(T Tr)\n_m\n(12.5.4)\nlog arr \n"]], ["block_11", [{"image_3": "498_3.png", "coords": [50, 46, 276, 204], "fig_type": "figure"}]], ["block_12", ["1010 \nX80800\n"]], ["block_13", ["104-\n~\n"]], ["block_14", ["106 \n-\n"]], ["block_15", [{"image_4": "498_4.png", "coords": [59, 182, 261, 344], "fig_type": "figure"}]], ["block_16", ["108 \n\u201470.6\u00b0C\n"]], ["block_17", ["I\nI\n001\nto\n100\nThne(h)\n"]], ["block_18", [";::::::::::?\\767%3\n\u2014\u201474.1\u00b0C\n"]], ["block_19", ["N.B.S. Polyisobutylene\n"]], ["block_20", ["_\n"]]], "page_499": [["block_0", [{"image_0": "499_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "499_1.png", "coords": [30, 48, 314, 269], "fig_type": "figure"}]], ["block_2", ["Figure 12.14\nTemperature dependence of the shift factors used in Figure 12.13, and the corre3ponding fit to\nthe WLF equation.\n"]], ["block_3", ["488\nGlass Transition\n"]], ["block_4", ["Figure 12.13\nMaster curve of modulus versus reduced time for the polyisobutylene data [7] in Figure 12.12.\n"]], ["block_5", ["log\n"]], ["block_6", ["NE\nE\nE\n0\n-\n.\nE 10\"\nE\nz\n:\nUJ\n'\n.\n"]], ["block_7", ["3T,25\u00b0c\n"]], ["block_8", ["A\n;\nTransition\n\u20180.\n1;\n"]], ["block_9", [{"image_2": "499_2.png", "coords": [41, 355, 307, 611], "fig_type": "figure"}]], ["block_10", ["18\n"]], ["block_11", ["1011\nEmml \"'E\n"]], ["block_12", ["103\n"]], ["block_13", [": \nPolylsobutylene\n'5\ns\n-.\n25\u00b0C\n5\n109\ng\n.\nE\n"]], ["block_14", ["E\na\n\u201805\n.\n'\nTerminal \n"]], ["block_15", ["-\u201480\n-60\n\u201440\n-20\nO\n20\n4O\n60\n"]], ["block_16", ["10-12\n10-9\n10*6\n10\u20143\n100\n103\n108\n"]], ["block_17", ["-\nmml mm! Inml IIIIIIIJIOIIII mm! mml mm] mm! mml \u201duni nun] \n"]], ["block_18", [":,\nRubbery plateau\n'5\n"]], ["block_19", ["=-\n.\n1\n"]], ["block_20", ["Temperature (\u00b0C)\n"]], ["block_21", ["\u201d81' (5)\n"]]], "page_500": [["block_0", [{"image_0": "500_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Time\u2014Temperature Superposition\n439\n"]], ["block_2", ["The dynamic moduli for a polyisoprene sample (M: 80,000) were measured at five temperatures:\n"]], ["block_3", ["even though the rheometer employed could only be used from about 0.001\u201410 Hz. Another\nillustration of how TTS is applied in practice is provided by the following example.\n"]], ["block_4", ["The preceding discussion illustrates how TTS can be used to expand the accessible range of\ntime (or frequency) scales dramatically. Furthermore, once the temperature dependence of a; has\nbeen determined for a particular property and a particular polymer, it should apply to all dynamic\nproperties and all reasonably high molecular weights of the same polymer. In general, this is the\ncase, but we still have one complication if we hope to just look up the WLF parameters C1 and C2\nfor a given polymer in some handbook. The problem is that the values of C1 and C2 depend on\nthe choice of reference temperature and there are obviously many possible choices of T,. Suppose,\nhowever, we make the particular selection Tr: Tg. We may find it experimentally inconvenient\nto make the measurement at Tg, but that does not matter; we can extrapolate our data using\nthe WLF function. When this choice of reference is made, the corresponding parameters are\ndesignated C? and Cg, as in Example 12.2. It turns out that the values of these parameters\nare very approximately universal, as illustrated in Table 12.2. For many systems C? is about\n"]], ["block_5", ["To obtain the shift factors, one can simply play with the values of aT in a spreadsheet until the\ndata superpose nicely, or each set of data can be plotted on separate sheets of paper (keeping the\naxes scales the same) and superposed in front of a bright light. However, because small vertical\nshifts may be permitted due to the pT term in the front factor, a more rigorous approach to\ndetermine 01- is actually to shift tan 5 (recall Equation 11.2.22). As tan 5 involves the ratio of G\"\nand G\u2019, the front factor cancels, and the data can be shifted exclusively along the horizontal axis\nto obtain the best superposition. Finally, small vertical shifts can be applied based on knowledge\nof p(T). Figure 12.150 shows the unshifted values of tan 8, and Figure 12.15d the shifted values,\nusing 20\u00b0C as the reference temperature and values of aT of 12000, 130, 7, and 0.11 for 40\u00b0C,\n\u2014 20\u00b0C, 0\u00b0C and 50\u00b0C, respectively. The final master curves for G\u2019 and G\u201d are shown in Figure\n12.15e and the superposition can be seen to be excellent. Vertical shifts have not been applied in\nthis instance.\n"]], ["block_6", ["10\u201415\nand\nCE\nis\nabout\n50\u00b0C\u201460\u00b0C, for example.\nThis\napproximate\nuniversality\npermits\nreasonable estimation of the T dependence for any polymer, once Tg is known. Also, if we compare\nthe VFTH and WLF parameters directly (Equation 12.4.12 and Equation 12.4.16), we see that\n"]], ["block_7", ["~\u2014 40\u00b0C,\n\u2014 20\u00b0C, 0\u00b0C, 20\u00b0C, and 50\u00b0C.Jr The unshifted data (0' and G\u201d versus co) are shown in\nFigure 12.15a and Figure 12.15b, respectively. Generate the master curve of G\u2019 and G\u201d versus (00'1\",\nusing 20\u00b0C as the reference temperature.\n"]], ["block_8", ["Cg TE TO, i.e., Cg correSponds to the interval between the glass transition temperature and the\nVogel temperature.\nWe conclude this section with a brief discussion of when TTS is not, or may not, be applicable.\nThe single, crucial assumption is that all the relevant relaxation times have the same temperature\ndependence over the measured temperature range. Note that adherence to the WLF dependence is\nnot a requirement; the remarkably strong T dependence as Tg is approached is what makes TTS\n"]], ["block_9", ["Solution\n"]], ["block_10", ["TData from J.C. Haley, Ph.D. Thesis, University of Minnesota, 2005.\n"]], ["block_11", ["Example 12.3\n"]]], "page_501": [["block_0", [{"image_0": "501_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "501_1.png", "coords": [23, 47, 249, 258], "fig_type": "figure"}]], ["block_2", ["Figure 12.1 5\nTime\u2014temperature superposition of dynamic moduli for polyisoprene. (a) G\u2019, (b) G\u201d, and (c) tan\n8 versus frequency at five temperatures, (d) tan 6 shifted horizontally to obtain a master curve and determine the\nshift factors, (6) G\u2019 and G\u201d master curves. (Data from Haley, J.C., Ph.D. Thesis, University of Minnesota, 2005.)\n"]], ["block_3", ["490\nGlass Transition\n"]], ["block_4", ["101\nI\nE\nI\nI\n102\ni\ni\ni\n10-2\nIo\u20141\n10\u2018?l\n101\n102\n10-2\n10-1\n100\n101\n102\n(a)\na) (rad/s)\n(b)\na) (rad/s)\n"]], ["block_5", ["(c)\na) (rad/s)\n(d)\nmaT\n"]], ["block_6", ["EL\u201c\n5\nEB\nL21\n3\nin\nEIEEEEEEEEEEEEEEEE\nv104\nIS!\na\n\u20182'\nE\n-\n1-\u201c\nIS!\nEB\n21\na\n"]], ["block_7", ["107\nI\n[\nI\n'\n107\n\u2014I\u2014\nI\nI\nU _\n0\nl2! 20\u00b0C\n8\n:20:\nE 5000\nu \u201440\u00b0c\n21 20\u00b0C\n106 -\n_\n'3'\n-\n\u00b0\n\u00b0\n530\u00b0C\nUUUDUUD\n105-\n5:330\n3500\nan.\u201d\nE]\nEl\n\u00b0\nEl\nDDUUUESEEEBEEEEE\nDE]\n105.\nBEBE\nElamEEI\nIZIZI'Z1\n[3'3\n[SI\n53 \nIZI\nE\nU I:\nA\nE\na\nE\nE\nA105\n\u2019\u2014\nE\nD\na\nI:\n"]], ["block_8", ["{'3\ni\nEH\n'21\n{5\nEB\n'3\nZ\n0:4\n"]], ["block_9", ["\u2018Q\na\n'2'\nE\n\u2018O\n\ufb01g\n:\nEH\n8\nc\nEU\n'3\n\u201c5\nLE\n+\u2014I\nE\nB\nB\nE\n'H\n"]], ["block_10", [{"image_2": "501_2.png", "coords": [37, 229, 405, 458], "fig_type": "figure"}]], ["block_11", [{"image_3": "501_3.png", "coords": [38, 248, 232, 418], "fig_type": "figure"}]], ["block_12", ["10-1\nI\nI\nI\n10-1\nI\nI _I_\n_l_\nL\nl\nI\nl\n10-2\n10-1\n10\u00b0\n101\n102\n10-2\n100\n102\n104\n105\n"]], ["block_13", ["o _\nI:\nE\n0 _\nLEE\n_\n10\nE B\na\nE\nU E E U U\n10\nn\n"]], ["block_14", ["IZI\nE\n104\n\u2014\nEB\n2'\nE\nE\n-\naa\na\nIZI\n103\n_\nE\nI\na\na\nE\nE\na\nE\na\n3\n103\n102 L\nE\n-\nf\na\n"]], ["block_15", ["102\nI\n1\u2014\nI\n102\nI\nl\nl\nI\n1\nl\n1\nl\nn\na\nE\na\nEl\n8\nE\nTr=20\u00b0C\n"]], ["block_16", ["IE!\n3\nE\nVJ\n%\n101\n__\nE\n101\nL_\n4\n"]], ["block_17", ["%\nISI\n53\na\na\n"]], ["block_18", ["E\nIS]\nI: U E 33\nE\nEFF\nE B\na E\n\u201dI J\nnoun\u201d\n\u201cWEE\n1%\n"]], ["block_19", ["EB\n21\nE\nE\nE\n"]], ["block_20", [{"image_4": "501_4.png", "coords": [116, 443, 364, 624], "fig_type": "figure"}]], ["block_21", ["(9)\n(031'\n"]], ["block_22", ["ED\n103 \n\u2014\nMw=eo,ooo\n"]], ["block_23", ["105 L\nI:\n-\n\"53W [ID\nA\n..\nEl\nru\n05 L-\n~\ng,\n1\nM\n"]], ["block_24", ["E!\n4\n.0\n10 \n_\ng\nPoISOprene\n"]], ["block_25", ["a\n"]], ["block_26", ["107\nl\nl\nl\nl\nl\nI\nl\nl\n"]], ["block_27", ["101'[\nI\nl\nI\nI\nI\nI\nI\n10-2\n100\n102\n104\n105\n"]], ["block_28", ["102 \n-\n"]], ["block_29", ["Ea\nTr=20\u00b0C\n"]], ["block_30", [{"image_5": "501_5.png", "coords": [198, 231, 421, 439], "fig_type": "figure"}]], ["block_31", [{"image_6": "501_6.png", "coords": [219, 49, 426, 249], "fig_type": "figure"}]], ["block_32", [{"image_7": "501_7.png", "coords": [239, 63, 411, 118], "fig_type": "figure"}]]], "page_502": [["block_0", [{"image_0": "502_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The primary factor, of course, is monomer structure. A list of representative Tg values for common\npolymers was provided in Table 12.1. They range from \u2014130\u00b0C for poly(dimethylsiloxane) to\n240\u00b0C for poly(p-phenylene terephthalimide) (Kevlar\u00ae). As indicated at the beginning of the\nchapter, the value of Tg is the single most important parameter in selecting a polymer for a\ngiven application. Those polymers with Tg below room temperature are sometimes called elasto-\nmers; when crosslinked into a permanent network structure (see Chapter 10) they exhibit tremen-\ndous elasticity, as in rubber bands, o-rings, gaskets, and tires. Polymers with Tgs near or above\n100\u00b0C are called thermoplastics; they are processed above Tg, but then are solidified into plastic\nparts by cooling. Polymers with particularly high Tgs, approaching or above 200\u00b0C, are termed\nengineering thermoplastics. They are in high demand for more strenuous applications and as such\ntend to be rather expensive. Polymers with Tgs between room temperature and 100\u00b0C that do not\ncrystallize are rather less widely applicable as bulk materials, but are useful as adhesives.\nIt is natural to seek a general correlation between the value of 1'\"g and some other, familiar\nproperty of the polymer, but in fact no robust correlation exists. Some broad generalizations may\n"]], ["block_2", ["useful, but it is not necessary for its validity. One way to violate the necessary assumption is for the\nsample to undergo some kind of transition with temperature, such as crystallization (Chapter 13) or\ncrosslinking (Chapter 10). A second situation arises in polymer mixtures or blends, where it is\nsometimes observed that the temperature dependences of the relaxation times of the two compon-\nents differ. A more subtle failure can arise at high effective frequencies, or short effective times, in\nthe transition zone of viscoelastic response. The origin of the problem can be understood as\nfollows. In the bead\u2014spring formulation of polymer dynamics (see Section 10.4 through Section\n10.6), all of the relaxation times are proportional to an underlying segmental relaxation time,\n"]], ["block_3", ["756g \ufb01bz/kT, where g is the bead friction factor. It is this friction factor (divided by T) that follows\nthe WLF form, and therefore all the relaxation times do too. However, this model does not describe\nthe relaxation of very short pieces of chain, on the scale of a few monomers and below. On this\nvery local scale, it is reasonable to anticipate that some relaxations may have a different T\ndependence. For example, the rotation of a particular functional group might be limited by a\nparticular conformational barrier rather than free volume, and therefore have an approximately\nArrhenius dependence. Consequently, it is to be expected, and some careful measurements have\nshown, that somewhere in the transition zone of viscoelastic response TTS will break down.\n"]], ["block_4", ["In this section we consider the major factors that affect the value of Tg in a given polymer material,\nindependent of the role of kinetics and measurement technique. These experimental observations\nshould also be compared with expectations based on the Gibbs\u2014DiMarzio theory outlined in\nSection 12.2.3.\n"]], ["block_5", ["be made, however.\n"]], ["block_6", ["12.1 with the persistence lengths listed in Table 6.1. Based on \ufb02exibility alone, poly(ethylene\noxide) has an anomalously high and poly(dimethylsiloxane) an anomalously low Tg.\n2.\nThe larger the rigid sidegroup, the larger the Tg. This correlation follows the previous one, in that\nbulky sidegroups impede backbone rearrangements. Exceptions here include polyisobutylene,\n"]], ["block_7", ["Factors That Affect the Glass Transition Temperature\n491\n"]], ["block_8", ["12.6.1\nDependence on Chemical Structure\n"]], ["block_9", ["12.6\nFactors That Affect the Glass Transition Temperature\n"]], ["block_10", ["1.\nBackbone \ufb02exibility increases as Tg decreases, in general. This may be simply understood\nfrom a conformational barrier argument. Flexible polymers tend to have smaller potential\nbarriers between conformations and thus at a given temperature conformational rearrange-\nments should be more rapid (see Section 6.1). The correlation with Tg follows because the\nglass transition corresponds to the freezing out of long-range backbone rearrangements.\nHowever, the correlation is not strict, as can be seen by comparing the Tg values in Table\n"]]], "page_503": [["block_0", [{"image_0": "503_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "503_1.png", "coords": [23, 358, 175, 394], "fig_type": "molecule"}]], ["block_2", ["For each phase, we construct the entropy of the mixture as follows:\n"]], ["block_3", ["A general expression for the glass transition temperature of a mixture of two components was\ndeveloped by Couchman [8], beginning with a thermodynamic relation for the transition similar in\nspirit to those discussed in Section 12.2. Furthermore, the same relation can be applied to a\npolymer mixed with a low molecular weight compound, a blend of two polymers, or a statistical\ncopolymer. We begin by equating the (specific) entropies of the mixture in the glass (superscript g)\nand liquid (superscript 1) states at the \u201csecond-order transition\u201d temperature, Tg:\n"]], ["block_4", ["where A is an empirical parameter. This expression turns out to be a very reasonable description\nfor many polymers. For example, for polystyrenes shown in Figure 12.16, the value of A is about\n105 g/mol, which implies that Tg becomes effectively independent of M when M exceeds that\nvalue. A 10\u00b0 depression of Tg below the high molecular weight limit would be seen for a polymer\nwith M =10,000. (Recall that the polystyrene in Figure 12.11 with M = 1,100 had a Tg of 375\u00b0C,\nwhich is not in good agreement with this relation; Equation 12.6.1 is more reliable for larger values\nof M.) Problem 12.12 provides another example of the M dependence of Tg. This M dependence of\nTg is important to appreciate, but it is not a particularly useful design parameter because the range\nover which Tg may be varied while retaining other desirable properties, such as mechanical\nstrength, turns out to be rather small.\n"]], ["block_5", ["Once a polymer has been selected, how may its Tg be modified? There are two simple routes; one is\nto vary molecular weight and the other is to add some amount of a low molecular weight diluent,\ncalled a plasticizer. The molecular weight dependence follows from the molecular weight depend-\nence of density, and is thus easily understood via the free-volume concept. For a homologous series\nof compounds, such as the linear alkanes CnH2n + 2, the density increases with n. This arises\nbecause of the chain ends; they are less dense essentially because covalent bonds are shorter than\nintermolecular nearest neighbor distances. Therefore the lower the n, the higher the fraction of the\nmaterial that is made up of chain ends, and the lower the density. The chain ends may be viewed as\na kind of impurity, and as with colligative properties such as boiling point elevation, freezing point\ndepression, or osmotic pressure (see Section 7.4), the change in transition temperature should be\nlinear in the mole fraction of impurity. In this case, the concentration of chain ends varies as 11M\",\nand therefore we anticipate:\n"]], ["block_6", ["which has a significantly lower Tg than polypropylene, even though it has double the number of\nsidegroups. Also, adding \ufb02exible side chains to relatively stiff backbones will have the effect of\nlowering Tg.\n3.\nPolymers that have weak interactions, such as the purely dispersive interactions of the\npolyolefins, have lower Tgs than more strongly interacting materials, such as the more polar\npoly(vinyl chloride) or poly(vinyl alcohol). This arises from the retarding effect of stronger\nintermolecular constraints on chain relaxation.\n4.\nTo obtain the highest Tgs, high backbone stiffness is essential, which is most easily conferred by\nincorporating aromatic groups within the backbone. Poly(tetra\ufb02uoroethylene) (Teflon\u00ae) is\nanother example in this class; the bulky \ufb02uorine atoms render the all-carbon backbone\nrather stiff.\n"]], ["block_7", ["492\nGlass Transition\n"]], ["block_8", ["12.6.3\nDependence on Composition\n"]], ["block_9", ["12.6.2\nDependence on Molecular Weight\n"]], ["block_10", ["53(Tg) slag)\n(12.6.2)\n"]], ["block_11", ["5 wlSl + W252 + AS,1n\n(12.6.3)\n"]], ["block_12", ["Tg(M,,) Tg(M \u2014\u2014> oo) MA\u2014\n(12.6.1)\n"]], ["block_13", ["H\n"]]], "page_504": [["block_0", [{"image_0": "504_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "504_1.png", "coords": [33, 80, 349, 336], "fig_type": "figure"}]], ["block_2", ["inset shows a plot of TB versus l/M,, and a linear regression fit. (Data from Fox, T.G and Flory, P.J., J . Polym.\nSci. 14, 315, 1954; Santangelo, P.G. and Roland, C.M., Macromolecules, 31, 4581, 1998.)\n"]], ["block_3", ["where WI and W2 are the weight fractions of components\n1 and 2, respectively (the appropriate\ncomposition variable because we are dealing with speci\ufb01c entropies), S 1 and SQ are the component\nentropies, and ASm is the entropy of mixing. This last quantity might be given by the Flory\u2014\nHuggins theory discussed in Chapter 7, for example, but if we assume that it is independent of state\nat a given temperature, it will cancel out from both sides of Equation 12.6.2. The component\nentropies can be constructed via Equation 12.2.7 using the pure component Tg as the lower bound\nand the mixture Tg as the upper bound:\n"]], ["block_4", ["In these expressions the only quantities that depend on the state of the mixture are the heat\ncapacities (specific heats); from Equation 12.6.2 we can see that the reference state entropy of\neach component in the liquid and glassy states is the same at the pure component Tg. Armed\nwith this information we can insert Equation 12.6.4 and Equation 12.6.3 into Equation 12.6.2\nto obtain\n"]], ["block_5", ["Factors That Affect the Glass Transition Temperature\n493\n"]], ["block_6", ["Figure 12.16\nDependence of the glass transition temperature on molecular weight for polystyrene. The\n"]], ["block_7", ["\u2014\n,\n380\nFl\u2019l\u2018ljllll\nIllllllllll'lTFIIIIl:\n-\n:I \n\u2014=\ng\n:\nE\nT\n"]], ["block_8", [{"image_2": "504_2.png", "coords": [41, 505, 195, 601], "fig_type": "figure"}]], ["block_9", ["51(Tg)=SI(T,I)+\n"]], ["block_10", ["82(Tg) :82(T ,2) \u2014I\u2014 J Cp\ufb02 dlnT\n"]], ["block_11", ["400\nI\nI\nIIIIII'\n\u2014l_\nl\nllllll\u2018\nI\n[llll\u2019ll'\nI\nI\nllllll\n"]], ["block_12", ["360\n\u2018r g\n\u2014\n_\nf\n_\n"]], ["block_13", ["380\n\u2014\n\u2014\n_\n.\n.\n_\n"]], ["block_14", ["320\nL\u2014\n!\n:\u2014\n_;\n\u2014_\u2014\n9\nE\nE\n'\nI\n320:\u2014\n\u2014:\n'\n\u2018\nI\nI\n\u2018\n\"\n"]], ["block_15", ["300 \u2014\"\n;\na\n"]], ["block_16", ["280\n1\nI\nI\nI\nIIIII\nI\nI\nI\n||I||\nI\nI\nlIllll\n]\n|\nI\nlllll\n"]], ["block_17", ["103\n104\n105\n105\n107\n"]], ["block_18", ["61\n300 \nF\n\u2018I\u2019\n0100\n210-4\n410-4\n610\u20144\n\"\n"]], ["block_19", ["\u2014\n. J: _:_-\u2014\u00b0\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\n-\n;\n\u201do\n_\n"]], ["block_20", ["_\nTg=371\u20141 x105/Mn\n_\n"]], ["block_21", ["_\n.\nE\nE\n_\n"]], ["block_22", ["_\n.I\n\u2014\n_\n_\n0..\nE\nE\n\u2018\nJ\"\n340 j\u2014\n\u2014:\n\u2018\n"]], ["block_23", ["h'\nI\nI\n|\n_\n"]], ["block_24", ["T,,.\n"]], ["block_25", ["Tg,2\n"]], ["block_26", ["TE\n"]], ["block_27", ["T8\n"]], ["block_28", ["J Cm] d In T\n"]], ["block_29", [{"image_3": "504_3.png", "coords": [143, 153, 324, 312], "fig_type": "figure"}]], ["block_30", ["Mn\n"]], ["block_31", ["(12.6.4)\n"]]], "page_505": [["block_0", [{"image_0": "505_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "505_1.png", "coords": [19, 38, 389, 110], "fig_type": "figure"}]], ["block_2", [{"image_2": "505_2.png", "coords": [28, 199, 277, 247], "fig_type": "molecule"}]], ["block_3", [{"image_3": "505_3.png", "coords": [32, 121, 221, 178], "fig_type": "molecule"}]], ["block_4", ["This is a straightforward application of the Fox equation. If we take 300 K as our target T3, and 360\nK as the T3 of pure PVC (see Table 12.1):\n"]], ["block_5", ["then we need a weight fraction of DEHP wl 0.22. Of course, in an actual application the desired\nTg would be con\ufb01rmed by measurements on compositions in the neighborhood of the estimated\nvalue.\n"]], ["block_6", ["Di-n-ethylhexyl phthalate (DEHP) and related compounds are commonly used to plasticize\npoly(vinyl chloride) (PVC) to produce the pliable material generically referred to as \u201cvinyl.\u201d\nA good plasticizer is miscible with the polymer in question, does not crystallize itself, and has a\nvery low vapor pressure. What fraction of DEHP should be added to PVC to bring Tg down below\nroom temperature, given that Tg for DEHP is about \u201486\u00b0C?\n"]], ["block_7", ["These relations are compared with data for polymer blends and statistical copolymers in Figure\n12.17 and Figure\n12.18, respectively. The agreement is generally very good. The use of a\nplasticizer is illustrated in Example 12.4 below.\n"]], ["block_8", ["This expression can be rearranged to give the general expression\n"]], ["block_9", ["Simpler versions of this expression can be obtained with additional assumptions. For example, if\nthe condition ACpJTg,\u2018 zACpng holds, and the Tgs are not too different, Equation 12.6.8\nreduces to the more commonly applied Fox equation:\n"]], ["block_10", ["Solution\n"]], ["block_11", ["494\nGlass Transition\n"]], ["block_12", ["Assuming that the heat capacities do not change with temperature, the integration gives\n"]], ["block_13", ["which can be rewritten\n"]], ["block_14", ["Example 12.4\n"]], ["block_15", [{"image_4": "505_4.png", "coords": [36, 57, 365, 104], "fig_type": "molecule"}]], ["block_16", [{"image_5": "505_5.png", "coords": [43, 585, 245, 615], "fig_type": "molecule"}]], ["block_17", [{"image_6": "505_6.png", "coords": [44, 134, 220, 164], "fig_type": "molecule"}]], ["block_18", [{"image_7": "505_7.png", "coords": [44, 339, 119, 387], "fig_type": "molecule"}]], ["block_19", ["l\nw!\n1 WI\n1\n1\n__ = _\n= w]\n.___. __ .___.\n+ _\n300\n187\n360\n187\n360\n360\n"]], ["block_20", ["1_=_\"KL+12_\n(12.6.9)\nTs\nTgJ\nTg\ufb02\n"]], ["block_21", ["Ts\nTs\nw, [ ACp,1dlnT =\u2014w2 [ AsdlnT\n(12.6.6)\n"]], ["block_22", ["WlACpJ + WZACP,2\nln Tg\n(12.6.8)\n"]], ["block_23", ["T\nT\nWIACP,1 111(fg1) + WQACpQ 1n (T35) = 0\n(12.6.7)\n"]], ["block_24", ["T3\nT3\nT3\nT8\nw, [ C\u00a791dlnT+w2 [ CizdlnTzwl [ 6\u2018},,,c11nT+w2 [ CdlnT\n(12.6.5)\n"]], ["block_25", ["711,1\n733,2\n"]], ["block_26", ["T1\u00bb:\nTia\nT14\nT892\n"]], ["block_27", ["__ \n"]], ["block_28", [{"image_8": "505_8.png", "coords": [81, 206, 235, 237], "fig_type": "molecule"}]], ["block_29", ["9\n39\n"]]], "page_506": [["block_0", [{"image_0": "506_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "506_1.png", "coords": [26, 43, 313, 301], "fig_type": "figure"}]], ["block_2", [{"image_2": "506_2.png", "coords": [28, 364, 314, 609], "fig_type": "figure"}]], ["block_3", ["Figure 12.17\nGlass transition temperature versus composition for miscible polystyrene/poly(oxy-2,6\u2014\ndimethy1\u20141,4\u2014phenylene) (PPO) blends, with fits to Equation 12.6.8 (solid curve) and Equation 12.6.9 (dashed\nline). (Data described in Couchman, P.R., Macromolecules, 11, 1156, 1978.)\n"]], ["block_4", ["Figure 12.18\nGlass transition temperatures for poly(styrene-stat\u2014n\u2014butylmethacrylate) copolymers, with\nfits to Equation 12.6.8 (solid curve) and Equation 12.6.9 (dashed line). (Data described in Kahle, S., Jorus, J.,\nHempel, E., Unger, R., Horing, S., Schr\u00e9ter, K., and Donth, E., Macromolecules, 30, 7214, 1997.)\n"]], ["block_5", ["Factors That Affect the Glass Transition Temperature\n495\n"]], ["block_6", ["T9\n(K)\n"]], ["block_7", ["T9\n(K)\n"]], ["block_8", ["500\nI\n|\n|\n|\n|\nI\nl\nl\u2014\nI\nI\n[\nI\nI\nI\nI\nl\nl\n"]], ["block_9", ["480\n"]], ["block_10", ["460\n"]], ["block_11", ["440\n"]], ["block_12", ["420\n"]], ["block_13", ["400\n"]], ["block_14", ["380\n"]], ["block_15", ["360\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\nI\n|\nI\nI\nl\nI\nI\nl\n|\nI\nI\n"]], ["block_16", ["380\nl\nI\nI\nI\nI\nI\nl\nI\n|\nI\nI\nI\nI\nI\nI\nI\nI\n"]], ["block_17", ["IIIIIIIIIllIIIIIIlII\u2014IWIIIIIIIIIIIIIII\nIIIlIlIIIIlIIlIlIlLILlLlIIIIIIIIIIIIIII\n300\n"]], ["block_18", ["370\n"]], ["block_19", ["360\n"]], ["block_20", ["350\n"]], ["block_21", ["340\n"]], ["block_22", ["330\n"]], ["block_23", ["320\n"]], ["block_24", ["310\n"]], ["block_25", ["II|ITIIIIITI\u2014IIWTIIIIIIITITI\n"]], ["block_26", ["WPPO\n"]]], "page_507": [["block_0", [{"image_0": "507_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The mechanical properties of polymer solids represent an extremely rich but complicated field of\nstudy, and we must necessarily limit our focus. For those students with a background in materials\nscience and engineering, the topic of this section should be familiar, but for those trained in the\nchemical sciences an introduction may be necessary. It could also be helpful to review some of the\nmaterial in Section 10.3 and Section 10.5 on elastic deformation, elastic modulus, and the stress\u2014\nstrain behavior of elastomers. We begin with the following observations.\n"]], ["block_2", ["1.\nEvery day adjectives such as strong, hard, and tough, which might be considered to be roughly\nsynonymous, now acquire specific and distinct meanings. Table 12.3 provides a glossary of\npertinent terminology.\n2.\nThroughout our consideration of viscoelasticity in Chapter 11, we emphasized the limit of\nsmall strains and strain rates: the material response was linear. Now we are concerned with\nlarge strain, nonlinear response, and in particular we ask the question, how will a given\nmaterial break?\n3.\nThere are actually many questions that one might ask about the nonlinear response of a\nmaterial, such as: how large a strain (or stress) can the material withstand before failure?\nHow large a strain (or stress) is required to achieve a nonlinear response? How large a strain\n(or stress) is required to undergo a nonrecoverable (plastic) deformation? How many times can\na material undergo a relatively small deformation without deteriorating? The answers to these\nquestions provide \u201cfigures of merit,\u201d but they need not be simply related to one another for a\ngiven material, and a different figure of merit may be most crucial for a particular application.\n"]], ["block_3", ["Although all polymers can exhibit a glass transition, we are concerned here with thermoplastics:\nmaterials that may be processed into desired shapes or forms above Tg, and then cooled below Tg\nfor use. For semicrystalline polymers to be discussed in Chapter 13, Tm will play the same role as\nTg in setting the boundary between liquid state processing and solid state application. As most\ncommon applications occur in the vicinity of room temperature, say from\n\u2014 25\u00b0C to 50\u00b0C,\nthemroplastics must have Tg well above that. Three prevalent thermoplastics are polystyrene\n(Tg m 100\u00b0C), poly(methyl methacrylate) (Tg m 110\u00b0C), and polycarbonate (Tgw 150\u00b0C), and this\nsubset will be sufficient to illustrate the main aspects of mechanical response.\nClearly, there are many different physical attributes beyond the value of Tg (or Tm) that could\neither favor or disfavor a particular polymer for a given application, and we will make no attempt to\ntreat these comprehensively. We will emphasize mechanical strength in this section, but even in that\ncontext we will only be able to cover a few aspects of a very rich subject. In general semicrystalline\npolymers have superior mechanical strength compared to amorphous polymers, but in many cases the\nlatter class is perfectly satisfactory. Amorphous polymers do offer some processing advantages,\nbecause vitrification is essentially instantaneous upon cooling, whereas the kinetics of crystallization\ncan be highly dependent on the polymer, the flow profile, and the presence of additives. On the other\nhand, processing is usually carried out tens of degrees above Tg, because of the strong temperature\ndependence of the viscosity (recall Sections 12.4 and 12.5), whereas semicrystalline polymers can be\nprocessedjust a few degrees above Tm. One further attribute of amorphous polymers that makes them\nthe materials of choice in numerous applications is optical clarity. As the molecular packing in the\nglassy state is that of a liquid, the refractive index is spatially homogeneous and isotropic; as\ndiscussed at length in Section 8.2, this minimizes the scattering of visible light. Accordingly, the\nthree polymers identified above are familiar in everyday use: polystyrene in clear plastic cups;\npoly(methyl methacrylate) as Plexiglas\u00ae; polycarbonate in compact discs. In contrast, semicrystal-\nline polymers contain more dense crystalline regions within an amorphous matrix; the resulting\n\ufb02uctuations in refractive index often render the material opaque, or at best, hazy.\n"]], ["block_4", ["496\nGlass Transition\n"]], ["block_5", ["12.7.1\nBasic Concepts\n"]], ["block_6", ["12.7\nMechanical Properties of Glassy Polymers\n"]]], "page_508": [["block_0", [{"image_0": "508_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Table 12.3\nTerminology Relevant to the Mechanical Strength of Materials\n"]], ["block_2", ["Term\nSignificance\n"]], ["block_3", ["Brittle fracture\nCraze\nDuctility\nElastic deformation\nEngineering stress\nHardness\nPlastic deformation\nStrain hardening\nModulus\nTensile strength\nTensile stress\nToughness\n"]], ["block_4", ["True stress\nYield point\nYield strength\n"]], ["block_5", ["Source: Adapted from Callister, W.D., Materials Science and Engineering: An Introduction, 5th ed., Wiley,\nNew York, 2000.\n"]], ["block_6", ["4.\nJust as there are many questions to ask, there are also many different testing protocols. We will\nemphasize tensile testing (i.e., uniaxial extension as in Section 10.3 and Section 10.5), but\nresponse to torsion, compression, shear, bending, and sudden impact are other common\nmodes. Molecular level information about the origins of strength and mechanisms of fracture\nis often obtained from controlled crack prOpagation experiments. The re5ponse of a given\nmaterial may vary significantly from one mode of deformation to another.\n"]], ["block_7", ["Figure 12.19 illustrates three schematic stress\u2014strain curves for polymers in uniaxial extension.\nCurves A and B represent glassy materials, whereas curve C is a rubber such as that discussed in\n"]], ["block_8", ["Figure 12.19\nSchematic illustration of the stress\u2014strain curves for polymers in uniaxial extension, drawn\napproximately to scale. Curve A: brittle, curve B: ductile (plastic), curve C: elastomeric.\n"]], ["block_9", ["Mechanical Properties of Glassy Polymers\n497\n"]], ["block_10", ["\u20195\"\n0.\nE,\nw\n40\n-\u2014\u2019\n\"\u2014-\nw\n9\n.-~\n4\u2014:\nx\n\\\n(D\n'\n.\n'l\n\\\n-..,\nB\n,X\ni\n\u00b0\"'\"\n"]], ["block_11", [{"image_1": "508_1.png", "coords": [45, 418, 305, 636], "fig_type": "figure"}]], ["block_12", ["80\nl\n"]], ["block_13", ["f\nI\n.\n\u201dI\n3\nc\n\u201da\nI\n____________\n0\n\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\u2014\n1\n"]], ["block_14", ["O\n4\n8\n"]], ["block_15", ["35\n"]], ["block_16", [".'\nx\n"]], ["block_17", ["A\n"]], ["block_18", ["Failure by rapid crack propagation without much deformation\nLocalized yielding consisting of microvoids interspersed with fibrils\nAbility to undergo substantial plastic deformation before failure\nCompletely recoverable deformation (note: not synonymous with Hookean)\nForce divided by initial cross\u2014sectional area\nAbility to withstand surface abrasion or indentation\nNonrecoverable deformation\nStress increasing with strain during plastic deformation\nStress divided by strain during small elastic deformation\nTensile stress at point of fracture\nMaximum engineering stress without fracture\nAmount of energy absorbed during fracture\n(pr0portional to the area under the stress\u2014strain curve)\nForce divided by instantaneous cross\u2014sectional area\nOnset of plastic deformation\nMagnitude of stress at the yield point\n"]], ["block_19", ["Strain\n"]]], "page_509": [["block_0", [{"image_0": "509_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Chapter 10 (see Figure 10.14 for an example). Curve A illustrates a material that undergoes brittle\nfracture. The modulus E, which corresponds to the slope of the curve in the small strain limit, is\nrelatively high, and there is not much deviation from linearity in response up to the point of failure.\nFailure typically occurs at strains of less than 10%. Curve B illustrates a much richer response,\nwith the stress first rising steeply, then exhibiting a maximum followed by a decrease, which\nevolves into a broad interval of nearly constant or slightly increasing values before fracture\nultimately intervenes. The stress maximum corresponds to the yield point, the strain at which the\nmaterial begins to undergo plastic deformation. The ensuing large range of strains, during which\nthe material may exhibit some degree of both strain softening and strain hardening, is characteristic\nof a ductile material. The strain at break may correspond to extensions of up to a factor of 100%.\nAs indicated in Table 12.3, the toughness of the material is given by the area under the stress-strain\ncurve, so materials that undergo ductile deformation can be extremely tough. At room temperature,\npolystyrene and poly(methyl methacrylate) show behavior closer to curve A, whereas polycarbon-\nate follows curve B. These differences may be attributed to differences in chain stiffness and\nentanglement density, as will be discussed in Section 12.7.3. Representative values of modulus,\ntensile strength, yield strength, and elongation at break are given in Table 12.4. Curve C for the\nrubber is markedly different. First, the modulus is lower by 2\u2014\u20143 orders of magnitude; for most\npolymer glasses E is in the range 2\u20144 GPa, whereas for rubbers the values have a wider range\n(depending on crosslink density), but 0.1\u20141 MPa is typical. Second, there is no yielding; the\ndeformation is largely recoverable throughout the deformation. Third, the stress increases mono-\ntonically up to the point of failure. Fourth, the strain at break may correspond to extensions\napproaching 1000%, and generally exceeds 500%.\n"]], ["block_2", ["Table 12.4\nRepresentative Values of Mechanical Properties for Common Thermoplastics\nat Room Temperature\n"]], ["block_3", ["Figure 12.20 shows stress\u2014strain curves for poly(methyl methacrylate) at various temperatures\nbelow Tg (approximately 110\u00b0C). The most notable feature of these curves is the brittle-to\u2014ductile\ntransition that occurs between 40\u00b0C and 50\u00b0C. For lower temperatures the behavior follows Curve\nA of Figure 12.19. The stress rises almost linearly with strain up to a maximum strain of a few\npercent, at which point the specimen breaks. In contrast, for temperatures between 50\u00b0C and T3 the\nresponse follows Curve B of Figure 12.19. The yield point corresponds to a strain amplitude\nsimilar to that at the point of brittle fracture at slightly lower temperatures. As temperature\nincreases within the ductile regime, the strain\u2014to-break increases substantially. This transition\nfrom brittle failure to a ductile response as the glass transition is approached from below is\ncommon, but the temperature interval over which the (usually more desirable) ductile behavior\nis seen depends on the polymer. The phenomenon of increasing brittleness with decreasing\ntemperature is a familiar one; plastic toys, automobile parts, etc. are often noticeably more brittle,\nor at least stiffer, when left outside on a cold winter day. Similarly most of us have seen a\ndemonstration of brittle fracture for a rubber ball, \ufb02ower, or similar soft material after immersion\nin liquid nitrogen. Macroscopic failure of a polymeric material corresponds to rupture of covalent\n"]], ["block_4", ["498\nGlass Transition\n"]], ["block_5", ["Polycarbonate\n2.4\n60\u201470\n62\n1 10\u2014150\nPoly(methyl methacrylate)\n2.2\u20143.2\n48\u201472\n54\u201473\n2-6\nPolystyrene\n2.3-6.3\n36-52\n\u2014\u2014\u2014-\n1.2\u20142.5\n"]], ["block_6", ["Source: Data compiled in Callister, W.D., Materials Science and Engineering: An Introduction, 5th ed., Wiley,\n"]], ["block_7", ["Tensile strength\nYield strength\nElongation at\nPolymer\nE (GPa)\n(MPa)\n(MPa)\nbreak (%)\n"]], ["block_8", ["12.7.2\nCrazing, Yielding, and the Brittle-to-Ductile Transition\n"]]], "page_510": [["block_0", [{"image_0": "510_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "510_1.png", "coords": [13, 37, 325, 272], "fig_type": "figure"}]], ["block_2", ["bonds at the molecular level. Brittle fracture implies that there is little deformation of the material\nprior to bond rupture and as reduced temperature leads to reduced molecular mobility, it is\nreasonable that the material undergoes less deformation before fracture at lower temperatures. In\ncontrast, ductility requires substantial molecular mobility, and is thus favored at higher temperat-\nures. In a qualitative sense the process of yielding is analogous to melting; the material is able to\n\u201cflow\u201d under the externally imposed deformation. In support of this notion is the fact that the\nbrittle\u2014to\u2014ductile transition occurs at progressively higher temperatures as the rate of deformation is\nincreased. At higher rates there is less time for molecular rearrangements to relieve the stress, and\nthus bond rupture is more likely.\nBrittle fracture in polymers occurs by the formation of cracks, as in other materials such as\nmetals and ceramics. However, there are major differences between the way cracks propagate in\nglassy polymers as compared to ceramics. In the latter, cracks tend to follow grain boundaries and\nother defects, which often leads to significant variation in performance depending on the grain\nsize, defect density, etc. of the material. In glassy polymers the lack of structural regularity,\ncombined with the substantial spatial extent and interpenetration of the covalently bonded chains,\nmakes the mechanical response more uniform. Cracks in polymers, therefore, are usually initiated\nby defects on the surface, such as scratches. In polymers, the mode of crack propagation after\ninitiation is also qualitatively different than in metals and ceramics; the new phenomenon is called\ncrazing. A craze is formed by a localized cavitation process, in which microvoids are created in the\npolymer to accommodate the applied strain. The microvoids are surrounded by a fibrillar structure\nwhere the fibrils contain extended polymer chains. An electron micrograph of a craze in polystyr-\nene is shown in Figure 12.21, along with a cartoon version of the craze structure. The thickness of a\ntypical craze is on the order of 100 nm and the fibrils are typically about 10\u201420 mm thick. As the\nmacroscopic strain is increased, the crazes propagate roughly perpendicular to the strain direction\nand then give way to cracks when the individual fibrils break. The formation of crazes has two\nmajor implications for the fracture process. First, a significant amount of energy is dissipated in the\ndrawing of polymer chains into fibrils; mechanical energy expended in this manner is no longer\navailable for bond rupture, so the ability to craze greatly enhances the mechanical strength of the\n"]], ["block_3", ["Figure 12.20\nStress\u2014strain reSponse for poly(methy1 methacrylate) at various temperatures, showing the\nbrittle\u2014to\u2014ductile transition. (Reproduced from Young, R.J., and Lovell, P.A., Introduction to Polymers, 2nd\ned., Chapman & Hall, London, 1991. With permission.)\n"]], ["block_4", ["Mechanical Properties of Glassy Polymers\n499\n"]], ["block_5", ["a.\u201c\n303 K\n"]], ["block_6", ["\u201cg.-\n"]], ["block_7", ["|EZ\n40 \n-\nE\n313 K\n"]], ["block_8", ["20 \n323 K\n\u2014\n"]], ["block_9", ["60 \n293 K\n\u2018\n"]], ["block_10", ["0\n"]], ["block_11", ["I\nI.\n0\n1o\n20\n30\n"]], ["block_12", ["i\n"]], ["block_13", ["I\nI\n277 K\n"]], ["block_14", ["0/0 Strain\n"]], ["block_15", [">\n333 K\n"]]], "page_511": [["block_0", [{"image_0": "511_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "511_1.png", "coords": [24, 156, 258, 262], "fig_type": "figure"}]], ["block_2", ["material. Second, the significant deformation that takes place during crazing means that typically\nthe strain at break is several percent. This should be contrasted with ceramic materials, such as\neveryday silicon dioxide\u2014based glasses, where brittle fracture occurs at strains well below one\ntenth of 1%.\nYielding occurs when the material is able to undergo significant deformation without crazing.\nThe crucial difference between crazing and yielding is that the former is a highly localized\ndeformation, whereas the latter involves macroscopic deformation. Crazing leads to brittle failure\nbecause the imposed strain must be accommodated locally, whereas yielding allows the strain to be\ndistributed over larger volumes of material. In such a case, there is less chance of a localized stress\nbuildup that is sufficiently large to break bonds. It is often the case that the macroscopic sample\nunder tension displays a shape transformation just after yielding known as necking (see Figure\n12.22), whereby the specimen becomes visibly thinner at some point along its length. In this region\nthe individual molecules have been extended to some significant degree. By analogy to rubber\nelasticity (Chapter 10) we know that the force to extend a sample increases with extension.\nTherefore it is easier to stretch the unnecked portion than to continue to extend the necked region.\nThe consequence of this is that once necking has occurred in one location, the size of the necked\nregion tends to grow while the original neck thickness is more or less preserved. During this region\nthe engineering stress will remain roughly constant, as individual portions of the sample deform.\nEventually the neck encompasses the entire specimen, and further extension leads to a more\nuniform deformation along the sample, often accompanied by some strain hardening. Detailed\nanalysis of the material response throughout the postyield regime is complicated, especially\nbecause the strain is not homogeneous throughout the material; the distinctions between the\nengineering stress (strain) and the true stress (strain) (see Table 12.2) become very important.\nIn contrast with polystyrene and poly(methyl methacrylate), which tend to craze at any\ntemperature 50\u00b0 or more below T , polycarbonate shows yielding behavior over 200\u00b0 below T ,\nas shown in Figure 12.23. Clearly there must be some important aspects of the molecular structure\n"]], ["block_3", ["Figure 12.21\n(a) Transmission electron micrograph of a craze in polystyrene. (Reproduced from Donald,\nA.M., The Physics of Glassy Polymers, Haward, R.N., and Young, R]. (Eds), 2nd ed., Chapter 6, Chapman &\nHall, London, 1997. With permission.) (b) Cartoon of a craze, illustrating voids, fibrils, and the deformation zone.\n"]], ["block_4", ["500\nGlass Transition\n"]], ["block_5", ["----\n.,\n"]], ["block_6", ["Voids\nDeformed polymer\nFibrils\n"]], ["block_7", ["Craze tip\n"]]], "page_512": [["block_0", [{"image_0": "512_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "512_1.png", "coords": [22, 35, 238, 199], "fig_type": "figure"}]], ["block_2", [{"image_2": "512_2.png", "coords": [28, 404, 281, 590], "fig_type": "figure"}]], ["block_3", ["The molecular mechanisms of plastic deformation and failure in glassy polymers are far from\ncompletely understood, but some important basic principles have been elucidated. A crucial\nparameter turns out to be the molecular weight between entanglements, Me (recall Section 11.6)\nor equivalently the number of entanglement strands per unit volume, pe:\n"]], ["block_4", ["that accounts for this difference. The dependence of the mechanical response of glassy polymers\non molecular variables will be taken up in the next section.\n"]], ["block_5", ["Figure 12.23\nStress\u2014strain behavior of polycarbonate in uniaxial extension at various temperatures.\n(Reproduced from G\u2019Sell, C., Hiver, J.M., Dahoun, A., and Souahi, A., J. Mater. Sci., 27, 5031, 1992.\nWith permission.)\n"]], ["block_6", ["12.7.3\nRole of Chain Stiffness and Entanglements\n"]], ["block_7", ["Figure 12.22\nSchematic illustration of necking in a ductile polymer undergoing uniaxial extension.\n"]], ["block_8", ["Mechanical Properties of Glassy Polymers\n501\n"]], ["block_9", ["E\n40\n"]], ["block_10", ["E \n-\n"]], ["block_11", ["m\n135\n"]], ["block_12", ["U)\na)\n125\n.2\n,\nE 40\n\u2014\u2014\n"]], ["block_13", ["m\n80\n8h\n60 \n-\n"]], ["block_14", ["100 \n25\u00b0C\n_\n"]], ["block_15", ["120\n\u2014|\u2014\nl\nI\n"]], ["block_16", ["p \n(12.7.1)\n6\nMe\n"]], ["block_17", ["20\n1\n"]], ["block_18", ["150\n0\nI\n'\nI\nl\n0\n0.2\n0.4\n0.6\n0.8\n1.0\n"]], ["block_19", ["Effective strain\n"]]], "page_513": [["block_0", [{"image_0": "513_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["This set of possibilities is not exhaustive, and none of these possibilities are mutually exclusive.\nAt a superficial level, however, we can match these possibilities to the observed phenomena. First\n"]], ["block_2", ["1.\nBond rupture: The material could fracture by scission of enough backbone bonds, where\n\u201cenough\u201d means all the bonds traversing some fracture surface. In the case of an \u201cideal\u201d\nchain of links under tension, the tension is equal in all links, and they would all fracture\nsimultaneously when the tension in each link reached a critical value. Of course, the tension\nwill not be uniformly distributed at the level of individual carbon\u2014carbon bonds in a real\npolymer material, so some would fracture first, thereby relieving the tension on the others.\nA simple calculation to estimate the stress required to fracture all the bonds in a polymer\n(Problem 12.15) shows that this value is significantly larger than both the yield and tensile\nstrengths (see Table 12.4).\n2.\nMolecular separation: In this limit the molecules would simply move apart without bond\nrupture. The energy required to do this would be related to the intermolecular interactions,\nquantified by the cohesive energy density (see Section 7.6). The dispersive interactions that\nhold nonpolar molecules together in liquid or solid phases are orders of magnitude weaker\nthan covalent bonds, so simple estimates of the required stresses are much lower than observed\nvalues (see Problem 12.16).\n3.\nLong chain pullout: It should be apparent that as high molecular weight polymers are\nthoroughly intertwined with one another, and because mobility is very low below T , the\nsimple molecular separation mode is not feasible. Another possibility is that under tension the\nindividual molecules disentangle by a kind of forced reptation, and are pulled apart without\nbond rupture. The friction associated with disentanglement would certainly dissipate energy,\ngiving rise to a larger tensile strength than case 2. However, as we saw in Chapter 11, it takes\nmany orders of magnitude longer for chains to disentangle than to undergo local rearrange-\nments; and as the local rearrangements themselves are already very slow, such a process seems\nunlikely.\n4.\nEnd strand or short chain pullout: Chain end segments that are shorter than Me, or chains with\nM < Me, could pull out from one another without the need for forced reptation.\n5.\nChain extension: Perhaps polymers that are sufficiently \ufb02exible can be induced to undergo\nlocal conformation rearrangements (such as trans to gauche for a carbon\u2014carbon bond), and\nthereby the material can extend without either rupture or pullout. In essence, this response\nwould be equivalent to the extensional \ufb02ow of an entangled melt, but of course, requiring\nsubstantially higher stresses to bring it about.\n"]], ["block_3", ["where p is the mass density. In Chapter 11 we discussed at some length how entanglements play a\ncentral role in the dynamic properties of a polymer melt, such as the viscosity and the stress\nrelaxation modulus. This analysis, in turn, was developed by analogy to rubber elasticity, where the\nfundamental parameter was the density of elastically active strands. It was argued that the\nentanglement strand in a melt acts like an elastically active strand in a network until the reptation\nprocess finally allowed each chain to escape its entanglements and pennit \ufb02ow. The value of Me\nwas also shown to be well correlated with the chain stiffness (i.e., the characteristic ratio, C00) and\nthe average thickness of the chain, through the packing length. Now we are asserting that the same\nconcepts have relevance to deformation below the glass transition. This might seem surprising at\nfirst because after all, in both melts and rubbers local relaxations (i.e., on a length scale smaller\nthan Me) are facile, whereas the main consequence of undergoing the glass transition is to freeze\nout such motions. However, we are now dealing with large imposed strains and as a consequence\nthe molecules must do something; one possibility is that chain segments less than M, long can slip\npast one another.\nWhat are the possible responses of a polymer glass to increasing strain? Here are some to\nconsider:\n"]], ["block_4", ["502\nGlass Transition\n"]]], "page_514": [["block_0", [{"image_0": "514_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["of all, responses 2 and 3 are not important to polymers. The fact that response 3 is not relevant is\nindicated by the molecular weight independence of tensile strength and yield strength for high\nmolecular weight materials. Fracture ultimately involves some degree of response 1. Yielding and\nductility are largely encompassed by response 5. The fact that ductility involves substantial chain\nextension is demonstrated by the fact that once the plastically deformed specimen is heated above\nTg, it undergoes large scale elastic recovery just as a deformed rubber or entangled melt would.\nCrazing involves a combination of responses 1, 4, and 5, but the chain extension part is restricted\nboth in extent and localized in Space.\nIf we consider first the propagation of a craze, the material responds both by forming extended\nfibrils and by cavitation. The latter process requires the formation of free surface and therefore\ncosts an amount of energy proportional to the product of the void surface area and the surface\ntension of the material. The formation of fibrils requires local extension of the polymer chains. The\ntelling feature of these fibrils is that the local extension ratio A (recall Section 10.5) within the fibril\nremains constant, even as the craze widens. Thus the fibrils grow in length by drawing in\nmore material, not by thinning the existing fiber. Furthermore, the extension ratio varies with\nthe material, but is well approximated by the ratio of the contour length of an entanglement strand\nto the root mean square end-to-end distance of the same strand. In other words, the material within\nthe fibrils corresponds to nearly fully extended entanglement strands. The material ultimately fails\nwhen chemical bonds within the fibrils undergo rupture, presumably when fresh material cannot be\ndrawn into the fibrils at the necessary rate.\nThe preceding discussion indicates that an entanglement network is a necessary precondition for\ncrazing to occur. This is also nicely illustrated in Figure 12.24, where the elongation at break is\nplotted as a function of total molecular weight for polystyrene. For high molecular weights the\nvalue is constant, but at low molecular weights it tends to vanish. The curve extrapolates to a value\nof about 35,000, indicating that there is essentially no tensile strength for chains shorter than this.\nRecalling from Table\n11.2 that Me for polystyrene is about 13,000, we can see that 2\u20143\nentanglement strands per chain are necessary to develop appreciable tensile strength and beyond\nabout 10 strands per chain the strength becomes constant. These values are easily appreciated.\n"]], ["block_2", ["Figure 12.24\nElongation at break in polystyrene as a function of weight average molecular weight. (Data\nreported in McCormick, H.W., Brower, F.M., and Kim, L., J. Polym. Sci, 39, 87, 1959.)\n"]], ["block_3", ["Mechanical Properties of Glassy Polymers\n503\n"]], ["block_4", ["\ufb01t\n,.\nO\n_\n"]], ["block_5", ["OJ\nC\n'-\n.\n2\n\u201dJ\n|\u2014\n.\n.-\n\u201869\n"]], ["block_6", ["g\nI\n"]], ["block_7", [{"image_1": "514_1.png", "coords": [41, 398, 263, 627], "fig_type": "figure"}]], ["block_8", ["10\u201c.-\n-:\n"]], ["block_9", ["'10\nT\n'.\n"]], ["block_10", ["o\n0\n. . :9\n. \n"]], ["block_11", ["..\n.\n\u2014\n"]], ["block_12", ["|\u2014\n-\n"]], ["block_13", ["E\no\n:\n"]], ["block_14", ["l\nI\nI\nI\nI\nl\n_I\nI\n"]], ["block_15", ["I\nl\nI\nI\n\u2014I-'\nI\nI\n"]], ["block_16", ["105\n6x105\n"]], ["block_17", ["l\u2018l/IW (g/mol)\n"]]], "page_515": [["block_0", [{"image_0": "515_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Imagine a plane through the material along which we wish to drive a crack. For any chain that is\nshorter than 2\u20143 Me long that traverses this plane, at least one end will be less than Me long and will\ntherefore be able to pull out readily (case 4 above). On the other hand, for chains longer than about\n10 Me, almost all chains that traverse the plane will be well anchored by several entanglement\nstrands on each side of the interface.\nWe now turn to one last question, namely, which molecular characteristics determine whether\nthe material responds by crazing, leading to brittle fracture, or by yielding? The accumulated\nevidence indicates that increasing chain \ufb02exibility (i.e., smaller C00) and higher entanglement\ndensities,\npe,\nfavor\nyielding\n[10].\nFor\nexample,\npolystyrene\ntends\nto\ncraze\n(C00 9.5,\npc 4.8><1019 cm 3), poly(methy1 methacrylate) tends to craze but yields at higher temperature\n"]], ["block_2", ["(COO 9.0, pe 6.8x1019 cm\n\u2014 3), whereas polycarbonate yields (COO = 2.4 [10], p6 = 5 x1020\ncm\u2014B). Of course, external variables such as temperature, deformation type, and deformation\nrate can affect the answer in a particular case, but all other things being equal, this demarcation is\ngenerally correct. The question can be reformulated slightly when we recognize that for any material\nthere must be critical stress values for both crazing and yielding to occur and the observed response\nwill correspond to the process with the lower critical stress value. Smaller values of C00 are generally\nassociated with chain structures that have smaller energy differences between trans and gauche\nconformers and thus chain extension should be more facile in such cases; this would lower the\ncritical yield stress. On the other hand, crazing requires cavitation. In addition to the surface energy\npenalty noted above, the formation of a cavity requires either chain scission or pullout across\nthe interface where the cavity forms. The higher the entanglement density, the shorter will be the\ndangling entanglement strands at the ends of chains and the smaller the extension ratio of one\nentanglement strand. Both factors increase the critical stress for crazing and thus more highly\nentangled chains favor yielding. Simply put, a higher entanglement density corresponds to a material\nthat is more tightly stitched together, and therefore more resistant to localized yielding (i.e., crazing).\n"]], ["block_3", ["a given polymer may be suitable for a certain application. There is no simple way to correlate Tg\nwith a particular chemical structure, although some general rules of thumb exist.\n4.\nThe value of Tg may be modified by changing molecular weight or by blending; the molecular\nweight and composition dependences of Tg are generally straightforward.\n5.\nThe concept of free volume is a particularly useful and physically intuitive way to understand\nthe glass transition, and the profound effect that proximity to Tg has on the temperature\ndependence of any viscoelastic or transport property. The free volume approach provides a\nnatural explanation for the widely used Vogel\u2014Fulcher\u2014Tammann and Williams\u2014Landel\u2014Ferry\nequations, which describe the temperature dependence of viscoelastic pr0perties above T3.\n6.\nThe principle of time\u2014temperature superposition is an essential ingredient in the study of\npolymer viscoelasticity because small changes in temperature produce large changes in the\npolymer relaxation times. Consequently measurements over a finite range of time or frequency\nat one temperature can be superposed with measurements at other temperatures to generate\n"]], ["block_4", ["In this chapter, we have examined the transition between the liquid state and the glassy state, which\ntakes place over a range of temperatures near a characteristic glass transition temperature, Tg. The\nprincipal points are the following:\n"]], ["block_5", ["1.\nThe glass transition is a kinetic transition, but it approximates a second-order thermodynamic\ntransition. A completely satisfactory theory of the glass transition is not yet available.\n2.\nThe glass transition temperature may be located in a variety of ways, but the most common\ntools are DSC and rheology.\n3.\nThe glass transition temperature is the single most important parameter in determining whether\n"]], ["block_6", ["504\nGlass Transition\n"]], ["block_7", ["12.8\nChapter Summary\n"]]], "page_516": [["block_0", [{"image_0": "516_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1N. Hirai and H. Eyring, J. Polym. Sci. 37, 51 (1959).\nis. Matsuoka, G.E. Johnson, H.E. Bair, and E.W. Anderson, Polym. Prepr. 22(2), 280 (1981).\n"]], ["block_2", ["\u00a713.. Wunderlich, D.M. Bodily, and MH. Kaplam, J. Appi. Phys. 35, 95 (1964).\n"]], ["block_3", ["Problems\n505\n"]], ["block_4", ["Problems\n"]], ["block_5", ["1.\nHirai and EyringT assembled the following data from diverse sources (scarcely any two pieces\nof data were measured in the same laboratory, much less on the same sample):\n"]], ["block_6", ["the value of Tg. However, in one lab the increase in heat flow associated with the transition is\nabout 30% larger than in the other. What could cause this discrepancy?\nThe figure illustrates plots of heat capacity for a particular polymer sample.\u00a7 The scans were\nobtained at a constant heating rate of 5.4\u00b0C/min, but different cooling rates were used to bring\nthe sample down to the starting temperature (350 K). Explain why the shape of the curve\n"]], ["block_7", ["Use these data to evaluate Tg, assuming that the latter is a true second-order transition.\nCompare your results with the values in Table 12.1 and comment on the agreement or lack\nthereof.\nTime\u2014temperature superposition was applied to the maximum in dielectric loss factors meas-\nured on poly(viny1 acetate).i Data collected at different temperatures were shifted to match at\nTg 28\u00b0C. The shift factors for the frequency (in hertz) at the maximum were found to obey\nthe WLF equation in the following form: log (0+ 6.9 [19.6(T\u2014 28)]/[42 + (T\u2014 28)]. Esti-\nmate the fractional free volume at Tg and a: for the free volume from these data. Recalling\nfrom Chapter 11 that the loss factor for the mechanical properties occurs at cur 1, estimate\nthe relaxation time for poly(vinyl acetate) at 40\u00b0C and 285\u00b0C.\nImagine a high molecular weight polymer that is a cycle. How would its Tg differ from a linear\npolymer of the same molecular weight? Make an argument based on the spirit of the Gibbse\nDiMarzio theory, and one based on free volume ideas. Do they make the same qualitative\nprediction? Incidentally, the experimental evidence suggests that at infinite molecular weight\nlinear and cyclic polymers have the same Tg, but as the chains get shorter, Tg for cycles\nactually increases.\nA polystyrene sample is split into two, and the Tg for each sample is measured by DSC in two\ndifferent laborataries. The reported results are 98\u00b0C and 106\u00b0C. Propose three possible simple\nexplanations for this difference.\nSuppose now that the two technicians in the previous problem consult one another on their\nmeasurement techniques. After repeating the measurements, the two laboratories agree as to\n"]], ["block_8", ["master curves of dynamic response, which can extend over as many as 20 orders of magnitude\nin reduced time or frequency.\nNoncrystallizable thermoplastics are in common use, for their ease of processing, optical\nclarity, and mechanical strength. Glassy polymers under large deformation may undergo either\nbrittle failure, through a distinctive localized yielding process known as crazing, or yield\nmacroscopically, leading to very large elongations before failure. Both processes involve\nextension of individual chains; the Operative response mode is strongly in\ufb02uenced by the\n\ufb02exibility and entanglement density of the material.\n"]], ["block_9", ["Rubber\n1.1\n5><106\n4.0><10\u20184\n1>-<10\u201811\nPolystyrene\n1.0\n7.7><10\u00b0\n1.75x10\u20144\n3x10\u201c12\nPolyisobutylene\n1.1\n4.0><10\u00b0\n4.5x10\u20144\n3x10\u201c\u201d\n"]], ["block_10", ["Vsp (cm3/g)\nACP (erg/K/g)\nA0: (K\u2019 1)\nAK (cmz/dyn)\n"]]], "page_517": [["block_0", [{"image_0": "517_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1\u2018ML. Williams and JD. Ferry, J. Colloid Sci. 10, 474 (1955).\n"]], ["block_2", ["506\nGlass Transition\n"]], ["block_3", ["7.\nWilliams and Ferry)r measured the dynamic compliance of poly(methy1 acrylate) at a number\nof temperatures. Curves measured at various temperatures were shifted to construct a master\ncurve at 25\u00b0C, and the following shift factors were obtained. Assess whether these data obey\nthe WLF equation; if so, evaluate the constants C1 and C2, and also C? and C5. Note that To 79\nTg 3\u00b0C for these data.\n"]], ["block_4", ["350\n360\n370\n380\n390\n400\n"]], ["block_5", ["T (\u00b0C)\nlog aT\nT (\u2018C)\nlog aT\n"]], ["block_6", ["depends on the prior cooling rate, and identify which curve corresponds to the smallest\ncooling rate. Speculate on the identity of the polymer.\n"]], ["block_7", [{"image_1": "517_1.png", "coords": [58, 516, 214, 611], "fig_type": "figure"}]], ["block_8", ["25.00\n0\n54.90\n\u2014 3.88\n29.75\n\u2014 0.98\n59.95\n\u2014 4.26\n34.85\n\u2014 1.80\n64.70\n\u2014 4.58\n39.70\n\u2014- 2.42\n69.50\n\u2014 4.88\n44.90\n\u2014 3.00\n80.35\n\u2014 5.42\n49.95\n\u2014 3.47\n89.15\n\u2014 5.72\n"]], ["block_9", [{"image_2": "517_2.png", "coords": [66, 80, 293, 371], "fig_type": "figure"}]], ["block_10", ["(Arb. units)\n"]], ["block_11", ["1\nop\n"]], ["block_12", ["Temperature (%)\n"]]], "page_518": [["block_0", [{"image_0": "518_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["TDJ. Plazek and ma. O\u2019Rourke, J. Polym. Sci. Part A-2, 9, 209 (1971).\nI3.]. Clarson, K. Dodgson and J.A. Semlyen, Polymer, 26, 930 (1985).\n"]], ["block_2", ["Problems\n507\n"]], ["block_3", ["12.\nThe following dependence of Tg on M for poly(dimethylsiloxane) was determined by\nDSC.i Prepare a plot to compare with Figure 12.16, and determine whether the functional\nrelationship of Equation 12.6.1 is followed in this case. Then select some measure of the\ncrossover to the high molecular weight asymptotic value of Tg (e.g., perhaps M where Tg is 5\u00b0\nless than the infinite M limit), and compare these values of M for poly(dimethylsiloxane) and\npolystyrene. Criticize or defend the proposition that this crossover corresponds to a certain\nnumber of persistence lengths (see Table 6.1) independent of chain structure.\n"]], ["block_4", ["10.\nIn small molecule glass formers, a common empirical definition of Tg is the temperature\nfor which the viscosity equals 1013 P. At first glance, this definition seems wholly inappro\u2014\npriate for polymers, because the viscosity at \ufb01xed temperature varies so strongly with M\n(i.e., a3\u20184), while Tg varies weakly with M. To assess the worth of this de\ufb01nition,\nconsider polystyrene with M 105 and M 10\u00b0. What would be the difference in the two\nTgs based on 7]: 101\u00b0, and how does this compare with what you would expect from\ncalorimetry?\n11.\nIt is sometimes suggested that the temperature dependence of the viscosity should follow\na power law, with an exponent v and a divergence (viscosity becomes infinite) at some\n"]], ["block_5", ["9.\nA certain extruder has optimum performance when 11 2 x104 P. The polymer of choice had\nthis viscosity at 150\u00b0C, and its MW was 80,000. Its Tg was about 75\u00b0C. An error in\npolymerization control led to a batch of the same polymer with MW=60,000. At what\ntemperature should the extruder be run?\n"]], ["block_6", ["8.\nPlazek and O\u2019Rourker measured the viscosity of polystyrene (M: 3400) as a function of\ntemperature. Plot these data in the Arrhenius format (ln 7; versus 1/7) and find apparent\nactivation energies for the lowest and highest temperatures measured. Compare these\nvalues to those obtained from the data in Figure 12.11b, and comment on their physical\nsignificance.\n"]], ["block_7", [{"image_1": "518_1.png", "coords": [44, 108, 211, 208], "fig_type": "figure"}]], ["block_8", ["T (\u00b0C)\n\u201440\n\u201430\n\u201420\n\u2014 10\n0\n10\n20\n30\n50\naT\n12,000\n950\n130\n28\n7\n2.5\n1\n0.43\n0.11\n"]], ["block_9", ["Such a dependence is common for many experimental quantities in the vicinity of a true\nphase transition (so-called critical phenomena). Assess whether the following aT data for\npolyisoprene (in part from Example 12.3) can be modeled in this way, and compare the\nquality of the fit to that with the WLF or VFTH function. How do the values of the Vogel\ntemperature and the putative critical temperature compare?\n"]], ["block_10", ["70.0\n12.967\n94.3\n6.888\n75.0\n11.152\n100.6\n5.995\n79.8\n9.749\n109.4\n5.047\n84.3\n8.619\n130.3\n3.418\n89.9\n7.609\n144.6\n2.627\n"]], ["block_11", ["To < Tgi\n"]], ["block_12", ["T (\u20180\n10g 7?\nT (\u00b0C)\n10g 77\n"]], ["block_13", ["71 (T To)\u201d\n"]]], "page_519": [["block_0", [{"image_0": "519_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "519_1.png", "coords": [28, 50, 293, 186], "fig_type": "figure"}]], ["block_2", ["Bower, D.I., An Introduction to Poiymer Physics, Cambridge University Press, Cambridge, UK, 2002.\nDonth, E., The Glass Transition, Springer, Berlin, 2001.\nEdiger, M.D., Angell, C.A., and Nagel, S.R., Snpercooled Liquids and Glasses, J. Phys. Chem. 100, 13200,\n1996.\nFerry, J.D., Viscoeiastic Properties of Polymers, 3rd ed., Wiley, New York, 1980.\n"]], ["block_3", ["Further Readings\n"]], ["block_4", ["References\n"]], ["block_5", ["508\nGlass Transition\n"]], ["block_6", ["539E309?\"\np\u2014A\n"]], ["block_7", ["13.\nSketch what you would expect to see for the dynamic modulus, G\u2019, measured at 1 rad/s as a\nfunction of temperature, for polyisoprene (M 250,000), polystyrene (M 100,000), and\npolystyrene (M 1,000). Make the vertical axis logarithmic, and recall that the low tem-\nperature modulus for all three polymers will be roughly 1010 dyn/cmz.\n14.\nShow that the Fox equation, Equation 12.6.9, can be obtained from the Couchman equation,\nEquation 12.6.8. Hint: use the proximity of the component Tgs to eliminate the logarithms.\n15.\nEstimate the tensile strength of polystyrene by assuming that the required stress corresponds\nto breaking backbone carbon\u2014carbon bonds. The bond energy is about 80 kcal/mol, and the\ndensity of polystyrene is about 1.05 g/cm3. How does this compare to the values in Table 12.4\n(be careful to match units)? How do you account for the difference?\n16.\nIt is known that in fracturing polystyrene, by driving a crack through the material, the fracture\nenergy released is on the order of 1 kJ/m2. From Chapter 7 we recall that the solubility\nparameter of polystyrene is 9.1 (cal/cm3)\u201d2. Use this value to estimate the surface energy of\npolystyrene, for example by estimating the \u201clost\u201d monomer\u2014monomer interactions per unit\narea by creating the new surface. How do these two values compare? What is the main origin\nof the difference?\n"]], ["block_8", ["1.\nKauzmann, W., Chem. Rev. 43, 219 (1948).\n2.\nGibbs, J.H. and DiMarzio, E.A., J. Chem. Phys. 28, 373 (1958).\n3.\nKovacs, A.J., J. Polym. Sci. 30, 131 (1958).\n4.\nDoolittle, A.K., J. Appl. Phys. 22, 1031; 1471 (1951); 23, 236 (1952).\n5.\nVogel, H., Physik. Z., 22, 645 (1921); Fulcher, G.S., J. Am. Chem. Soc. 8, 339; 789 (1925); Tammann, G.\n"]], ["block_9", ["and Hesse, G.Z., Anorg. Alig. Chem. 156, 245 (1926).\nWilliams, M.L., Landel, RR, and Perry, J.D., J. Am. Chem. Soc. 77, 3701 (1955).\nCatsiff, E. and Tobolsky, A.V., J. Colioid Sci. 10, 375 (1955).\nCouchman, P.R., Macromoiecules 11, 1156 (1978).\nCallister, W.D., Materiais Science and Engineering: An Introduction, 5th ed., Wiley, New York (2000).\nWu, 3., Polym. Eng. Sci., 30, 753 (1990).\n"]], ["block_10", ["240\n123.4\n1630\n146.1\n7,720\n149.5\n310\n129.1\n2260\n147.5\n10,060\n149.5\n530\n137.4\n2460\n148.0\n12,290\n149.4\n630\n139.4\n2920\n148.3\n14,750\n149.7\n810\n141.2\n4030\n149.2\n18,250\n149.8\n990\n143.3\n4880\n148.8\n21,390\n149.8\n1290\n144.8\n6330\n149.3\n25,460\n149.7\n"]], ["block_11", ["M,\nT, (K)\nM,\nT, (K)\nM,\nT, (K)\n"]]], "page_520": [["block_0", [{"image_0": "520_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Haward RN. and Young RI. The Physics of \nSperling, L.H., Physical Polymer Science, Wiley, New Yorlf, \nStrobl, G., The Physics of Polymers, 2nd ed., Sprmger, Berhn, 199?-\nWard, I.M., Mechanical Properties of Solid PolymerS, 2Hd 6d\u201d Wlley, New YOT\na\n-\n'\n-\n'\nWunderllch, B., Thermal Analysis, Academic Press, New York, 1990.\nYoung, RJ. and Lovell, P.A., Introduction to Polymers, 2nd ed'! Chapman & H311, LOHdOH, 1991.\n"]], ["block_2", ["Further Readings\n509\n"]]], "page_521": [["block_0", [{"image_0": "521_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Which polymers crystallize and which do not?\nThe simple answer is: stereoregular polymers (polyethylene, isotactic or syndiotactic polypro-\npylene, poly(ethylene oxide), etc.) crystallize, stereoirregular polymers (atactic polystyrene and\npoly(methyl methacrylate)) or polymers of mixed microstructure (mixed cis and trans polydienes)\ndo not. In order to crystallize, it is necessary for a few monomers to pack into a regular unit cell,\nwhich can then be stacked on a lattice to fill space. It is almost impossible for an atactic polymer\nsuch as polystyrene to form a regular unit cell, because the side groups are placed on one side of the\nbackbone or the other at random. However, even this rule is not always obeyed. For example, a vinyl\npolymer with the formula \u2014(CH2-C(AB)),,\u2014, in which the groups A and B have similar sizes, may\nbe able to pack into a regular array. Or, an atactic polymer in which the side chain is strongly polar\nmay sometimes crystallize; poly(vinyl alcohol) and poly(vinyl \ufb02uoride) are examples.\n"]], ["block_2", ["What is the structure of a polymer crystal and how do we characterize it experimentally?\nPolymers crystallize with three levels of structure, as illustrated in Figure 13.1. On the first\nlevel, individual chain backbones form helices (of which an all-trans conformation is a Special\ncase), and pack with neighboring chains to form unit cells. The typical unit cell contains only a few\nmonomers, and has dimensions of 2\u201420 A on a side. The structure of unit cells will be considered\nin Section 13.2. On the second level, unit cells pack into thin sheets, called lamellae, which are\ntypically 100\u2014500 A thick and several microns wide in the other two dimensions. The chain\nbackbones lie at some fixed angle relative to the thin direction, and often fold by 180\u00b0 at the\nlamella surface in order to reenter the crystal. Chain\u2014folded lamellae, illustrated in Figure 13.1, are\n"]], ["block_3", ["Why is crystallization in polymers important?\nThe world\u2019s most popular synthetic polymer, in terms of volume produced per year, is polyethyl-\nene; polyethylene can crystallize. Other high-volume polymers such as isotactic polypropylene,\npoly(hexamethylene adipamide) (Nylon 6,6), and poly(ethylene terephthalate) crystallize, as\ndo many specialty materials, such as poly(tetra\ufb02uoroethylene) (Te\ufb02on) and poly(p-phenylene\nterephthalamide) (Kevlar). In general, crystallinity conveys enhanced mechanical strength, greater\nresistance to degradation, and better barrier properties.\n"]], ["block_4", ["In this chapter we consider several important aspects of crystallinity in polymers. It is a remarkably\nrich field, and our coverage necessarily limited, but we will raise most of the central issues and\nshow how many of them have been addressed. In the previous chapter we examined the glass\ntransition, and introduced it by contrast to crystallization and melting. The first section of Chapter\n12, therefore, serves as part of the introduction to the current topic as well. To continue this\nintroduction, we pose a series of basic questions, and the answers form the outline for the rest of\nthe chapter.\n"]], ["block_5", ["a unique morphological feature of polymers. These structural details are explored primarily\nthrough x-ray diffraction and electron microscopy, as will be discussed in Section\n13.4. In\n"]], ["block_6", ["13.1\nIntroduction and Overview\n"]], ["block_7", ["Crystalline Polymers\n"]], ["block_8", ["13\n"]], ["block_9", ["511\n"]]], "page_522": [["block_0", [{"image_0": "522_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Compared to the glass transition, crystallization and melting are thermodynamic transitions;\ndoes this mean that a thermodynamic analysis will be more successful?\nWould that it were so, but it is not. The process of crystallization is dominated by kinetics. The\nunderlying reason is simple. A molten polymer is a random jumble of intertwined chains, whereas\nthe crystal has long sections of chains fully extended and closely packed in parallel with one\nanother. Once crystallization starts, different sections of one chain may be in several different\ncrystallites, which prevents the full extension of the chain. This is illustrated schematically in\nFigure 13.2. Consequently, and recalling the characteristically sluggish dynamics of polymers (see\nChapter 11), it is not at all surprising that crystallization of bulk polymers may never achieve the\nstate of minimum free energy. Furthermore, for the same kinetic reasons it is never possible to\nachieve 100% crystallinity in a bulk polymer, and thus there is always a nonequilibrium mixture of\namorphous and crystalline regions within the material. Such polymers are more correctly termed\nsemicrystalline. Nevertheless, thermodynamics still provides a good deal of insight, particularly\nwhen the process of melting is considered; in Section 13.3 and Section 13.4 in particular we will\npursue a thermodynamic analysis.\n"]], ["block_2", ["How can we understand the kinetics of crystallization in general?\nCrystallization is a first-order phase transition (recall Section 12.2), and it proceeds by the\nprocess of nucleation and growth, that is, stable nuclei of the new (crystal) phase appear, and then\n"]], ["block_3", ["Figure 13.1\nThree levels of structure in a crystalline polymer. The packing of individual helical chains\ngives unit cells with dimensions of a few angstroms. The unit cells are packed into chain\u2014folded lamellae, with\ncharacteristic thicknesses on the order of 10 nm. The lamellae splay, bend, and branch to form spherulites,\nwhich can exceed millimeters in size. The space between the individual lamellae is filled with amorphous\nmaterial.\n"]], ["block_4", ["favorable cases, polymer single crystals can be grown from solution, which greatly facilitates\ndetailed structural analysis of the unit cell and the lamellae. Finally, in a bulk sample the lamellae\ngrow to fill space, often producing a three-dimensional structure called a spheralite, which can be\ntens or hundreds of microns across. These can be observed in an optical microscope. The structure\nof spherulites, and some other bulk morphologies, will be taken up in Section 13.6.\n"]], ["block_5", ["512\nCrystalline Polymers\n"]], ["block_6", [{"image_1": "522_1.png", "coords": [41, 95, 174, 227], "fig_type": "figure"}]], ["block_7", [{"image_2": "522_2.png", "coords": [46, 65, 263, 275], "fig_type": "figure"}]], ["block_8", [{"image_3": "522_3.png", "coords": [53, 219, 267, 273], "fig_type": "molecule"}]], ["block_9", [{"image_4": "522_4.png", "coords": [70, 96, 135, 162], "fig_type": "molecule"}]]], "page_523": [["block_0", [{"image_0": "523_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "523_1.png", "coords": [6, 52, 386, 244], "fig_type": "figure"}]], ["block_2", ["The unit cell contains the smallest number of atoms, in the appropriate spatial relationships,\nnecessary to enable prediction of the full structure of a macroscopic single crystal by repetitive\nclose stacking of unit cells. A schematic unit cell is illustrated in Figure 13.3a. The lengths of the\nthree sides are designated a, b, and c, and the correSponding angles are a, B, and 'y. A single crystal\ncan thus be generated by filling space with unit cells, so that the structure repeats exactly every a A\n"]], ["block_3", ["In this section we shall examine the smallest level of structure displayed by polymer crystals, the\nunit cell. These unit cells typically have dimensions between 2 and 20 A, which lie in the same\nrange as atomic and small-molecule crystals. Accordingly, the same experimental technique is\nemployed to determine the spacings and the symmetries of the unit cell, namely x-ray diffraction\n(XRD), or, as it is more commonly known in polymer science, wide-angle x-ray scattering\n(WAXS). We will not describe WAXS in detail, but the basic process can be understood through\ncomparison with the determination of molecular structure and size by light scattering, which was\ncovered in Chapter 8. Similarly, we will only review the basics of crystallography; more details are\navailable in a number of monographs and texts.\n"]], ["block_4", ["as one moves along the a direction, etc. However, the precise location of the various atoms within\nthe unit cell requires further information than is contained in the three lengths and three angles.\nThere are seven crystal classes (cubic, trigonal, hexagonal, tetragonal, orthorhombic, monoclinic,\nand triclinic), which are defined by different constraints on the values (a, b, c) and (a, [3, 7). These\nare given in Table 13.1. Within these classes there are further subdivisions, for example a cubic\nunit cell could be face-centered, body-centered, 0r primitive. When these possibilities are included\nit turns out that there are 14 distinct structures, called Bravais lattices, within these seven classes.\nFinally, within a certain Bravais lattice there can be many different ways in which the atoms are\n"]], ["block_5", ["Figure 13.2\nIllustration of a semicrystalline polymer melt. Individual crystal lamellae are indicated by\ndashed lines.\n"]], ["block_6", ["grow in size by incorporation of more chains or chain segments from the amorphous (liquid) phase.\nWe will examine growth kinetics on two different levels, that of individual crystals in Section 13.5,\nand that of the crystalline fraction of the material in Section 13.7.\n"]], ["block_7", ["Structure and Characterization of Unit Cells\n513\n"]], ["block_8", ["13.2.1\nClasses of Crystals\n"]], ["block_9", ["13.2\nStructure and Characterization of Unit Cells\n"]]], "page_524": [["block_0", [{"image_0": "524_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "524_1.png", "coords": [22, 576, 345, 666], "fig_type": "figure"}]], ["block_2", [{"image_2": "524_2.png", "coords": [26, 58, 215, 166], "fig_type": "figure"}]], ["block_3", ["CUbic\na=b=c\na=B=y\u201490\u00b0\nTrigonal\na=b=c\n05:13:\u201c)!\n90\u00b0\nHexagonal\na=b7\u00e9 c\n0: =6 =90\u00b0, y=120\u00b0\nTetragonal\na b # c\na B 'y 90\u00b0\nOrthorhombic\na 75 b # c\na: B 'y 90\u00b0\nMonoclinic\na 75 b 75 c\na: y=90\u00b0, B # 90\u00b0\nTriclinic\n(175 b 75 c\naa\u00e9\ufb01a\u00e9 \u201cy\n"]], ["block_4", ["Table 13.1\nConstraints on Unit Cells\n"]], ["block_5", ["arranged in detail. These arrangements are described by sets of symmetry operations that leave the\nstructure unchanged, such as rotation about an axis by an angle of 60\u00b0, 90\u00b0, or 180\u00b0, or re\ufb02ection\nthrough a plane, etc. The set of symmetry operations that applies to a particular crystal specifies a\nspace group; in total there are 230 different space groups. The determination of the space group\nand the full unit cell structure of a polymer crystal including bond angles, bond lengths, and\ninterchain distances is the first goal of polymer crystallography.\nPolymers are extended one-dimensional objects in the crystalline state, and the overall direction\nof the backbone corresponds to one axis of the unit cell. For reasons that will become apparent this\naxis is termed the\ufb01ber axis. By convention, it is assigned to the c axis in the unit cell (except for\nmonoclinic crystals, where it is b). All of the seven crystal classes in Table 13.1 are found in\npolymers, except one: cubic. This exclusion can be easily understood. In a cubic unit cell all three\naxes must be equivalent, but because of the covalent backbone it is virtually impossible for\nmonomers to pack equivalently in three orthogonal directions.\nParticular crystallographic planes or directions are commonly labeled with the assistance of\nMiller indices hkl. One vertex of the unit cell is chosen as the origin. The plane of interest\nintersects the three unit cell axes at particular coordinates x, y, and z. The Miller indices are\nobtained as h a/x, k b/y, and 1: 6/2. By choosing an equivalent plane so that it intersects all\nthree axes within a single unit cell, the Miller indices are always integers. Note that if a plane is\nparallel to a particular axis the intersection occurs at infinity, and so the corresponding Miller\nindex is 0. The rules for specifying directions are analogous. In the case of a negative Miller index,\nthe value is indicated with an overbar. By convention, a particular plane is referred to by the\nMiller indices in parentheses, that is (hid), the family of equivalent planes by {t }, a particular\ndirection as [hkl], and the family of equivalent directions by (hkl). By these conventions, the ab\nface of the unit cell is (001), and the c axis is [001]; the (110) plane and [111] axis are illustrated\nin Figure 13.3b.\n"]], ["block_6", ["Crystal class\nUnit-cell sides\nUnit-cell angles\n"]], ["block_7", ["Figure 13.3\nSchematic of a unit cell, showing (a) axes a, b, and c, and associated angles a, B, and 'y and (b)\nthe (110) plane (shaded) and the [111] direction (arrow).\n"]], ["block_8", ["514\nCrystalline Polymers\n"]], ["block_9", [{"image_3": "524_3.png", "coords": [38, 578, 202, 658], "fig_type": "figure"}]], ["block_10", [{"image_4": "524_4.png", "coords": [39, 34, 360, 175], "fig_type": "figure"}]], ["block_11", [{"image_5": "524_5.png", "coords": [156, 53, 351, 152], "fig_type": "figure"}]]], "page_525": [["block_0", [{"image_0": "525_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "525_1.png", "coords": [15, 484, 370, 669], "fig_type": "figure"}]], ["block_2", [{"image_2": "525_2.png", "coords": [33, 491, 258, 633], "fig_type": "figure"}]], ["block_3", ["and is illustrated schematically in Figure 8.4. Each of the 230 space groups has its own set\nof \u201callowed reflections,\u201d namely a set of values of q (or 6) that will satisfy Bragg\u2019s law. In\nprinciple, therefore, an experimental scattering pattern could determine the space group uniquely.\nIn practice there is a lot more to it, first because many space groups have several re\ufb02ections in\ncommon, and second because whether or not particular re\ufb02ections are actually seen will depend\non a host of factors, most important of which is the orientation of the incident x-ray beam relative\nto the crystal.\nIn an x-ray diffraction experiment, a collimated beam of monochromatic x-rays is directed on to\nthe sample and the diffracted radiation is monitored as a function of 9. The detector itself can be\none-dimensional or two-dimensional, with the latter becoming increasingly common. The condi-\ntion of the specimen itself plays a crucial role in the experiment; the limiting cases are those of a\nsingle crystal and a polycrystalline sample. In a single crystal, the entire portion of the specimen\nthat is illuminated by the x-rays has the same orientations of a, b, and c in the laboratory coordinate\nsystem, and thus the beam is incident along a single direction through the unit cell. The result is\nthat Bragg\u2019s law is satisfied not only for particular values of 6] but also for particular directions\nin space and the scattering pattern on an area detector will be a series of particular spots. This is\nillustrated in Figure 13.4a for the particular cases of a layered sample and a hexagonal crystal of\nrods. The pattern that is observed will depend critically on the particular angle between the incident\nbeam and the unit cell. This is illustrated in Figure 13.4b for the same two crystals, now rotated by\n90\u00b0 relative to Figure 13.4a. In the case of the layered sample, the beam is now incident normal to\n"]], ["block_4", ["where A is the wavelength, D is the spacing between lattice planes, 9 is the angle between the\nincident and diffracted radiation, and m is a positive integer. Recall that a reciprocal lattice vector\npoints in the direction normal to a series of lattice planes, with a magnitude given by 27r/D. The\nmagnitude of the scattering vector is defined as\n"]], ["block_5", ["Figure 13.4\nDiffraction patterns on an area detector for (a) a set of parallel sheets viewed edge\u2014on, and a\nhexagonal array of rods viewed end-on.\n(continued)\n"]], ["block_6", ["As noted above, XRD or WAXS is the standard tool for crystal structure determination. In Section\n8.2 we derived Bragg\u2019s law, and noted that the criterion for observing a diffraction peak is that the\nscattering vector, q, match a reciprocal lattice vector of the crystal (see Figure 8.3). Bragg\u2019s law\nmay be written as\n"]], ["block_7", ["Structure and Characterization of Unit Cells\n515\n"]], ["block_8", ["(a)\n"]], ["block_9", ["13.2.2\nX-Ray Diffraction\n"]], ["block_10", [{"image_3": "525_3.png", "coords": [38, 207, 121, 246], "fig_type": "molecule"}]], ["block_11", ["6\nmA 2D sin<\u20142\u2014)\n(13.2.1)\n"]], ["block_12", ["41?\n_\n9\nq 7\nSin (2)\n(13.2.2)\n"]], ["block_13", [{"image_4": "525_4.png", "coords": [55, 497, 152, 571], "fig_type": "figure"}]], ["block_14", [{"image_5": "525_5.png", "coords": [66, 494, 250, 576], "fig_type": "figure"}]], ["block_15", [{"image_6": "525_6.png", "coords": [128, 496, 253, 614], "fig_type": "figure"}]], ["block_16", [{"image_7": "525_7.png", "coords": [160, 495, 252, 564], "fig_type": "figure"}]]], "page_526": [["block_0", [{"image_0": "526_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["the layers, and no diffraction is seen, whereas when viewed side-0n, the hexagonal sample appears\nto be layers. This cartoon is overly simplified, in that we have tacitly assumed that the layers and\nrods are smooth and structureless. The arrangement of atoms within the layers could give rise to\nadditional re\ufb02ections beyond those indicated. The important conclusion, however, is that for a full\ncrystallographic analysis of a single crystal sample, the sample must be rotated systematically\nrelative to the incident beam, and the resulting scattering patterns collected and interpreted as a\n"]], ["block_2", ["Figure 13.4 (continued)\n(b) the sheets viewed through and the rods viewed edge\u2014on; and (c) a polycrys\u2014\ntalline sample.\n"]], ["block_3", ["(C)\n"]], ["block_4", ["516\nCrystalline Polymers\n"]], ["block_5", [{"image_1": "526_1.png", "coords": [37, 47, 310, 220], "fig_type": "figure"}]], ["block_6", [{"image_2": "526_2.png", "coords": [45, 45, 173, 214], "fig_type": "figure"}]], ["block_7", [{"image_3": "526_3.png", "coords": [63, 52, 166, 143], "fig_type": "figure"}]], ["block_8", [{"image_4": "526_4.png", "coords": [91, 220, 287, 388], "fig_type": "figure"}]], ["block_9", [{"image_5": "526_5.png", "coords": [114, 44, 297, 159], "fig_type": "figure"}]], ["block_10", [{"image_6": "526_6.png", "coords": [166, 51, 307, 213], "fig_type": "figure"}]], ["block_11", [{"image_7": "526_7.png", "coords": [186, 50, 296, 142], "fig_type": "figure"}]]], "page_527": [["block_0", [{"image_0": "527_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["these \u201c\ufb02uctuations\u201d are much larger, and locked in place.)\n4.\nThe main difference between light and x-rays as a structural tool lies in the value of A. We saw\nin Chapter 8 that Rg needed to be at least 100 A or larger in order to be determined by light\nscattering (A m 4000\u20147000 A). In contrast, a common x\u2014ray source involves an inner electronic\ntransition in copper, which gives off a photon at 1.542 A. For a 10 A unit cell size (i.e., D 10\nA) inserted into Bragg\u2019s law (Equation 13.6.1), we would need a diffraction angle of 8.80, a\nvery reasonable value.\n"]], ["block_2", ["whole. Furthermore, depending on the sample and the size of the detector, it may be necessary to\nmove the detector in space to collect a different angular range.\nA polycrystalline sample comprises many little crystals perhaps microns in size, with approxi-\nmately random relative orientations. Consequently, the incident beam simultaneously samples\nalmost all possible incident angles relative to the unit cell. The result is a so-called powder pattern,\nconcentric rings of scattered intensity at radial positions on the detector corresponding to particular\nvalues of q, as illustrated in Figure 1340. Such a pattern may be sufficient to narrow down the\npossible crystal structure to a few candidates, but it is unlikely to determine a space group\nuniquely. Consequently, single crystals are greatly preferred when determination of the unit cell\nstructure is the goal. However, production of a macroscopic single crystal in polymers is no easy\nfeat. It was first shown in the 19503 that single polymer crystals could be grown by careful\ncrystallization from solution, but this strategy is not always convenient or practical. A more\ncommon approach is to draw a fiber of the polymer during or prior to crystallization, by applying\na uniaxial extensional deformation (see Section 10.5). In this case the individual chains can\nbecome highly extended along the draw direction, and the crystals develop with a strong prefer\u2014\nence for the c axis to lie along this direction (hence the term fiber axis). Although the resulting\nmorphology is not exactly a single crystal (see Section 13.6), it is much more highly organized than\na polycrystalline material, and the analysis correspondingly more definitive.\nWe will not explore this analysis in any further detail, but conclude this overview with some\nadditional comments, especially in comparison to light scattering discussed in Chapter 8:\n"]], ["block_3", ["5.\nXRD or WAXS measurements can readily be made on laboratory\u2014scale instruments, but the\nadvantages of utilizing synchrotron radiation (e.g., at a National Laboratory such as Argonne\nor Brookhaven) should be noted. Synchrotron sources provide an incident \ufb02ux of x-rays that is\nseveral orders of magnitude larger than a laboratory source. Furthermore it can be much better\ncollimated, and the wavelength may be tuned to a convenient value. These features combine to\nprovide scattering patterns with much better resolution in much shorter time intervals.\n6.\nElectron diffraction measurements can also be very useful in determining the unit cell\nstructure. The experimental concept is identical to XRD, except that the incident beam consists\nof electrons. Electrons have a de Brogliewavelength of~0.03 A, and so are well suited to\nstructures on the 1\u201410 A scale. The measurement can be made on an electron microscope, to\nbe discussed in Section 13.4.4.\n"]], ["block_4", ["Structure and Characterization of Unit Cells\n517\n"]], ["block_5", ["l.\nDiffraction is the same phenomenon as coherent scattering; it is merely a convention to use the\nterm dt\u2018\ufb01raction when discussing crystalline samples, and scattering when amorphous mater\u2014\nials are studied. The former gives rise to sharp peaks at particular angles, whereas the latter\nleads to more smoothly varying intensity versus angle curves.\n2.\nThe x-ray diffraction community often employs slightly different terminology, such as k or s\nin place of q, and 9 may be defined as one half of the 6 we have employed.\n3.\nThe scattering of x-rays is caused by differences in electron density from atom to atom,\nwhereas light scattering arises from \ufb02uctuations in refractive index (i.e., polarizability). The\natoms of primary interest in polymers, C, H, O, and N, are not particularly strong scatterers of\nx-rays, but this limitation is overcome by the long-range order in the sample. (Recall that in\nlight scattering from solutions, we have to wait for a spontaneous concentration \ufb02uctuation to\nhave the correct spacing and orientation to satisfy the Bragg condition; in a crystalline sample,\n"]]], "page_528": [["block_0", [{"image_0": "528_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "528_1.png", "coords": [17, 455, 464, 642], "fig_type": "figure"}]], ["block_2", ["Note: Asterisks in column 4 denote the chain axis.\n"]], ["block_3", ["Source: From Wunderlich, B., in Macromolecular Physics, Vol. 1: Crystal Structure, Morphology, Defects, Academic Press,\nNew York, 1973.\n"]], ["block_4", ["Table 13.2\nUnit Cell Parameters for Several Polymers\n"]], ["block_5", ["Polyethylene I\nOrthorhombic\n1*2/1\n7.42, 4.95, 255*\n90, 90, 90\n4\nPolyethylene II\nMonoclinic\n1*2/1\n8.09, 253*, 4.79\n90, 107.9, 90\n4\nPoly(tetra\ufb02uoroethylene) I\nTriclinic\n1*13/6\n5.59, 5.59, 1688*\n90, 90, 119.3\n13\nPoly(tetraflu0roethylene) II\nTrigonal\n1*15/7\n5.66, 5.66, 1950*\n90, 90, 120\n15\nPolypropylene (iso)\nMonoclinic\n2*3/1\n66.6, 20.78, 6495*\n90, 99.62, 90\n12\nPolypropylene (syndio)\nOrthorhombic\n4*2/1\n14.50, 5.60, 7.40*\n90, 90, 90\n8\nPolystyrene (iso)\nTrigonal\n2*3/1\n21.9, 21.9, 6.65*\n90, 90, 120\n18\nPoly(vinyl alcohol) (atac)\nMonoclinic\n2*1/1\n7.81, 251*, 5.51\n90, 91.7, 90\n2\nPoly(vinyl fluoride) (atac)\nOrthorhombic\n2*1/1\n8.57, 4.95, 252*\n90, 90, 90\n2\n1,4\u2014Polyisoprene (cis)\nOrthorhombic\n8*1/1\n13.46, 8.86, 8.1*\n90, 90, 90\n8\n1,4\u2014Polyisoprene (trans)\nOrthorhombic\n4*1/1\n7.83, 11.87, 4.75*\n90, 90, 90\n4\nPoly(ethylene oxide)\nMonoclinic\n3*7/2\n8.02, 13.1, 19.3\n90, 126, 90\n28\nPoly(hexamethylene adipamide), a\nTriclinic\n14*1/1\n4.9, 5.4, 17.2\n48.5, 77, 63.5\n1\nPoly(hexamethylene adipamide), B\nTriclinic\n14*1/1\n4.9, 8.0, 17.2\n90, 77, 67\n2\n"]], ["block_6", ["What determines the unit cell that a particular polymer will adopt? This is sometimes difficult to\npredict a priori, but can usually be rationalized after the fact. From a thermodynamic point ofview, the\ncrystal is a low temperature state and is therefore dominated by enthalpic considerations. In particular,\nthe monomers want to maximize their favorable energetic interactions, which generally means to\npack as closely as possible; remember from Chapter 7 that the van der Waals energy of attraction\nbetween molecules falls off approximately as 1/(distance)6. (Note that certain interactions such as\nhydrogen bonding have a preferred distance.) However, there are both intramolecular and intermo\u2014\nlecular interactions to consider. In most instances, polyethylene for example, each chain first adopts\nits lowest energy conformation (all-trans in this case), and then packs as closely as it can to its\nneighbors. This prioritization corresponds to the first two of Natta and Corradiui\u2019s Rules [1] for\npolymer crystallization. On the other hand, there may be situations for which the chain conformation\nin the crystal is not the lowest energy conformation of the isolated chain. Furthermore, the optimum\npacking of chains is often quite subtle, as we shall see when we consider some examples. In general,\npolymers adopt one conformational motif for the backbone: a helix. The helix is described by three\nnumbers, such as 2* 1/1 for polyethylene. This terminology means 2 backbone atoms constitute a basic\nrepeat unit, and there is 1 repeat unit in each full turn ofthe helix. In fact, such a 1/1 helix is identical to\nthe all-trans conformation. As indicated in Table 13.2, many more interesting helices are found.\nFigure 13.5 shows the arrangement of molecules in the polyethylene unit cell. X\u2014ray measure\u2014\nments show that the dimensions are a 7.4 A, b 4.9 A, and c 2.5 A, and that it is orthorhombic.\nIn regard to this unit cell, we observe the following:\n"]], ["block_7", ["Macromolecule\nCrystal class\nHelix\na, b, c (A)\na, B, y (0)\n# of units\n"]], ["block_8", ["1.\nThe c-axis corresponds to both the short axis of the unit cell and the axis along the molecular\nchain. The observed repeat distance in the c direction is what would be expected between\nsuccessive substituents on a fully extended hydrocarbon chain with normal bond lengths and\nangles (see Section 6.1).\n2.\nThe distances between all hydrogen atoms are approximately the same in this structure, so\nthere is no problem with overcrowding.\n3.\nWhile not overcrowded, the polyethylene structure uses space with admirable ef\ufb01ciency, the\natoms \ufb01lling the available space to about 73%. For comparison, recall that close-packed\nspheres fill space with 74% efficiency, so polyethylene does about as well as is possible.\n"]], ["block_9", ["518\nCrystalline Polymers\n"]], ["block_10", ["13.2.3\nExamples of Unit Cells\n"]]], "page_529": [["block_0", [{"image_0": "529_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure\n13.5 shows that the equivalent of two ethylene units are present in each unit cell.\nAccordingly, the mass per unit cell is\n"]], ["block_2", ["Use the unit cell dimensions cited above to determine the crystal density of polyethylene. Examine\nFigure 13.5 to determine the number of repeat units per unit cell.\n"]], ["block_3", ["The density of the crystal is obtained from the ratio of these two quantities:\n"]], ["block_4", ["(In fact, x-ray diffraction can usually determine unit-cell dimensions to three or even four\nsigni\ufb01cant figures, but we have rounded off in this calculation to avoid specifying more sample\ninformation or experimental conditions.) This density may be compared with a typical value of\n0.94 g cm\u20183 for polyethylene in the molten state.\n"]], ["block_5", ["Table 13.2 provides a list of representative unit cell parameters. Several interesting observations\nmay be made on the basis of these data:\n"]], ["block_6", ["Since all angles in the cells are 90\u00b0, the volume of the unit cells is\n"]], ["block_7", ["Figure 13.5\nCrystal structure of polyethylene: (a) unit cell shown in relationship to chains and (b) view of\nunit cell perpendicular to the chain axis. (Reprinted from Bunn, C.W., F\ufb01bersfrom Synthetic Polymers, R. Hill\n(Ed.), Elsevier, Amsterdam, 1953. With permission.)\n"]], ["block_8", ["One of the things that can be done with a knowledge of the unit-cell dimensions is to calculate\nthe crystal density. This is examined in the following example.\n"]], ["block_9", ["Solution\n"]], ["block_10", ["Example 13.1\n"]], ["block_11", ["Structure and Characterization of Unit Cells\n519\n"]], ["block_12", ["1.\nPolyethylene, along with poly(tetra\ufb02uoroethylene) (Te\ufb02on) and poly(hexamethylene adipa\u2014\nmide) (Nylon 6,6) exhibits two different crystal forms. This polymorphism is not uncommon,\nand it indicates that there is a subtle balance of terms in the free energy, so that each structure\n"]], ["block_13", [{"image_1": "529_1.png", "coords": [35, 53, 217, 163], "fig_type": "molecule"}]], ["block_14", [{"image_2": "529_2.png", "coords": [43, 59, 209, 117], "fig_type": "molecule"}]], ["block_15", ["_ \n..\u2014 = 1.02\n'3\np\n9.07 x 10-23\n5 g cm\n"]], ["block_16", ["2 repeat units\n1 mol repeat units\n28.0 g\n= 9.30 x 1043 g (unit cell)_1\nunit cell\n6.02 x 1023 repeat units\n1 mol repeat units\n"]], ["block_17", ["7.4Ax4.9Ax2.5AX(1cm\n"]], ["block_18", ["3)= 9.07 x 10\u201423 cm3 (unit cell)\u20141\n103 A\nunit cell\n"]], ["block_19", [{"image_3": "529_3.png", "coords": [173, 51, 425, 195], "fig_type": "figure"}]], ["block_20", [{"image_4": "529_4.png", "coords": [220, 58, 371, 174], "fig_type": "figure"}]], ["block_21", [{"image_5": "529_5.png", "coords": [277, 99, 323, 142], "fig_type": "molecule"}]]], "page_530": [["block_0", [{"image_0": "530_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["\"\u2014 \"' O = C\nCH2\n\\\n/\n\u2018\nCH2\nCH2\n0\n/\n\\\n"]], ["block_2", ["has a temperature range where it is in the equilibrium state. These issues are often challenging\nto sort out experimentally, because of the preponderance of kinetic influences. For instance,\nthe observed form may simply be the one that crystallizes more rapidly, rather than the state of\nlowest free energy.\n2.\nAlthough poly(tetra\ufb02uoroethylene) is structurally very similar to polyethylene, the increase in\nsize and interactions in going from H to F causes a significant change in the conformation.\nBoth polymorphs are almost all\u2014trans, but not quite; form I (which is reported to be stable\nbelow 19\u00b0C) takes 13 backbone bonds to undergo six turns of the helix, and form 11 takes 15\nbonds to make seven turns.\n3.\nPoly(viny1 alcohol) and poly(vinyl \ufb02uoride) provide two examples of atactic polymers that can\ncrystallize, both in the all-trans conformation.\n4.\nThe structure of poly(hexamethylene adipamide) provides an excellent illustration of how\nspecific, strong interactions can dictate the structure. In this instance it is hydrogen bonding\nbetween the carbonyl group and the amide hydrogens that determine the packing. Figure 13.6a\nprovides a schematic illustration of the hydrogen bonding pattern in sheets, and Figure 13.6b\nshows the difference between the oz and B forms in terms of the registry of the hydrogen bonds.\n5.\nIsotactic polypropylene forms a 3/1 helix, which corresponds to a trans\u2014gauche plus\u2014trans\u2014\ngauche plus. . .sequence of backbone bond conformations. The chain packing forms a mono-\nclinic unit cell, as shown in Figure 13.7.\n"]], ["block_3", ["Figure 13.6\nThe or and [3 crystal forms of poly(hexamethylene adipamide) (Nylon 6,6), showing (a) the\nhydrogen bonding pattern and (b) the molecular registry in the two forms. (Reprinted from Young, RJ. and\nLovell, P.A., Introduction to Polymers, 2nd ed., Chapman and Hall, London, 1991. With permission.)\n"]], ["block_4", ["520\nCrystalline Polymers\n"]], ["block_5", ["/\nCH2\n/\n\\\nCH2\nC = O \n\\\n/\nC = O\n\u2014 \u2014 -\nH N\\\n\u2014 \u2014 H \u2014 N\nCH2\n\\\n/\nCH\nCH\n/\n2\n\\2\nCH\nCH\n\\2\n/\n2\ncl\nCH2\nCH\n/\n/\n\\2\n\u20193\n"]], ["block_6", [{"image_1": "530_1.png", "coords": [39, 328, 214, 633], "fig_type": "figure"}]], ["block_7", ["(a)\n(c)\n"]], ["block_8", ["CH2\nCH2\n0))\n\\\n/\nCH2\nCH2\n/\n\\\nCH2\nN H \n\\\n/\nN\u2014H\n\u2014--\nO=C\n/\n\\\n"]]], "page_531": [["block_0", [{"image_0": "531_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "531_1.png", "coords": [28, 35, 357, 251], "fig_type": "figure"}]], ["block_2", [{"image_2": "531_2.png", "coords": [33, 232, 272, 485], "fig_type": "figure"}]], ["block_3", ["Figure 13.7\n(a) Illustration of four possible 3/1 helices for an isotactic Vinyl polymer such as polypropyl-\nene. A and D are right-handed, B and C are left-handed, A and D are equivalent if inverted, as are B and C.\n(Reprinted from Wunderlich, B., Macromolecular Physics, Vol. 1: Crystal Structure, Morphology, Defects,\nAcademic Press, New York, 1973. With permission.) (b) The polyprOpylene unit cell looking down the chain\naxis. Each apex of a triangle indicates a methyl group. The thick line segments suggest the projection of the\nbond to the methyl group; there are three left-handed and three right\u2014handed helices shown.\n"]], ["block_4", ["In dealing with experimental thermodynamics, one of the criteria for a true equilibrium to have\nbeen established is to achieve the state of interest from opposite directions. Accordingly, we are in\nthe habit of thinking of the equilibrium melting point of a crystal, or the equilibrium freezing point\nof the corresponding liquid, as occurring at the same temperature. In dealing with polymer crystals,\n"]], ["block_5", ["(b)\n"]], ["block_6", ["Thermodynamics of Crystallization\n521\n"]], ["block_7", ["13.3\nThermodynamics of Crystallization: Relation of Melting Temperature\nto Molecular Structure\n"]], ["block_8", [{"image_3": "531_3.png", "coords": [44, 252, 249, 385], "fig_type": "figure"}]], ["block_9", [{"image_4": "531_4.png", "coords": [46, 45, 199, 244], "fig_type": "figure"}]], ["block_10", ["0*\n9%.\nl\n'l\n"]], ["block_11", [{"image_5": "531_5.png", "coords": [117, 49, 259, 243], "fig_type": "figure"}]], ["block_12", ["U\n"]]], "page_532": [["block_0", [{"image_0": "532_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "532_1.png", "coords": [20, 454, 284, 631], "fig_type": "figure"}]], ["block_2", ["discussion of the transition in this direction. Furthermore, because of the kinetic complications\noccurring during the formation of the crystal, the resulting transition from crystal liquid also\nbecomes more involved.\nIf polymers had infinite molecular weight and formed infinitely large crystals, the thermo-\ndynamics of the transition would be simpler, although the kinetics might very well be worse.\nAssuming, temporarily, that any kinetic complications can be overlooked, we will define the\ntemperature of equilibrium (subscript e) between crystal and liquid for a polymer meeting the\nin\ufb01nity criteria (superscript 00) stated above as T30. Subsequently, we will use the superscript 00 to\nindicate either infinite crystal dimension, infinite molecular weight, or both; it will be clear from\nthe context which is meant. In such a case, the melting point of the crystal would be T30 and the\nfreezing point of the liquid would also be T30. The facts that actual molecular weights are less than\ninfinite and that crystals have finite dimensions both tend to drive the equilibrium transition\n"]], ["block_3", ["we are not so fortunate as to observe this simple behavior. The transition from liquid \u2014> crystal is\nso overshadowed by kinetic factors that some even question the value of any thermodynamic\n"]], ["block_4", ["the lamellar thickness, 6.\n2.\nMelting occurs over a range of temperatures, as shown previously in Figure 12.1. The range\nnarrows as the crystallization temperature increases. This is probably due to a wider range of\ncrystal dimensions, and less perfect crystals, for lower temperatures of formation.\n3.\nThere is a suggestion of convergence of these lines in the upper right\u2014hand portion of Figure\n13.8. For this polymer T310 is estimated to be 28\u00b0C~\u2014\u2014not an unreasonable point of convergence\nin the lines in Figure 13.8.\n4.\nThe value of T310 is often estimated by extrapolation of the experimental Tm versus TC curve\nuntil it intersects the Tm TC line, as illustrated for poly(p-phenylene sulfide) in Figure 13.9.\n"]], ["block_5", ["temperature below Tgo. The fact that kinetic complications also interfere means, in addition, that\nthe experimental temperature of crystallization TC does not equal the temperature of melting Tm,\nand that neither equals T\u00a7\u00b0.\nFigure 13.8 illustrates some of these points for cis 1,4-polyisoprene. The temperature at which\nthe crystals are formed is shown along the abscissa, and the temperature at which they melt, along\nthe ordinate. Note the following observations:\n"]], ["block_6", ["522\nCrystalline Polymers\n"]], ["block_7", ["Figure 13.8\nMelting temperature of crystals versus temperature of crystallization for cis 1,4-polyisoprene.\nNote the temperature range over which melting occurs. (Reprinted from Wood, LA. and Bekkedahl, N., J. \nPhys, 17, 362, 1946. With permission.)\n"]], ["block_8", ["1.\nThe lower the crystallization temperature, the lower the melting point. This correlation will be\nunderstood in the next section through consideration of crystal dimensions, and particularly\n"]], ["block_9", ["(\u00b0C)\n"]], ["block_10", ["Crystallization\nTemperature\n"]], ["block_11", ["40\nI\nI\nI\nI\nI\nI\nI\nI\nI\n'I\nI\n"]], ["block_12", ["Temperature of crystallization (\u00b0C)\n"]], ["block_13", ["Melting starts\n'\n"]], ["block_14", ["\u201420\n\u2014t 0\n0\n10\n"]], ["block_15", ["I\nl\nl\nl\nl\nl\nl\n"]]], "page_533": [["block_0", [{"image_0": "533_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["We shall take up the kinetics of crystallization in more detail in Section 13.5 and Section 13.7.\nFor the present, our only interest is in examining what role kinetic factors play in complicating\nthe crystal\u2014liquid transition. The main issue is that the lamellar thickness, 6, depends on the\ncrystallization temperature, as a result of kinetic considerations. Accordingly, I? is related to Tc,\nbut may not have much to do with T50. The melting point Tm of the resulting crystal is less than it\nwould be if the crystal had infinite dimensions (7'33). This latter temperature approaches T30 as\nM\u2014>oo. In the end, Tm gives a better approximation to a valid equilibrium parameter, although it\nwill still be less than T310 owing to the finite dimensions of the crystal and the finite molecular\nweight of the polymer. We shall deal with these considerations in the next section. For now we\nassume that a value of T30 has been obtained and consider the thermodynamics of this phase\ntransition.\nWe begin our application of thermodynamics to polymer phase transitions by considering the\nfusion (subscript f) process: crystal \u2014> liquid. Figure 13.10 shows schematically how the Gibbs free\nenergy of liquid (subscript 1) and crystalline (subscript c) samples of the same material vary with\ntemperature. For constant temperature\u2014constant pressure processes the criterion for spontaneity is\na negative value for AG (just as in our consideration of phase equilibria in Chapter 7), where the A\nsignifies the difference between the final and initial states for the property under consideration.\nApplying this criterion to Figure 13.10, we conclude immediately that above T310, AGf G1 GC is\nnegative and melting is Spontaneous, whereas below T30, AGf> 0 and it is fusion that is spontan-\neous. At T310 both phases have the same value of G; at that temperature, therefore, AG: 0 and a\ncondition of equilibrium exists between the phases.\n"]], ["block_2", ["Figure 13.9\nHoffmann\u2014Weeks plot for low molecular weight (LMW) and medium molecular weight\n(MMW) poly(p-pheny1ene sulfide). (Reproduced from Lovinger, Al, Davis, DD, and Padden, F.J. Jr.,\nPolymer, 26, 1595, 1985. With permission.)\n"]], ["block_3", ["Thermodynamics of Crystallization\n523\n"]], ["block_4", ["290 \nMMW\n_\nMelting\n"]], ["block_5", ["(\u00b0C)\n"]], ["block_6", ["temperature\n"]], ["block_7", ["Such an extrapolation is known as a Ho\ufb01mann\u2014Weeks plot. However, note that the data\npresented here, and many other data sets, suggest that a linear extrapolation may not be\ncompletely accurate.\n"]], ["block_8", ["270\n.\n|\nI\n|\n|\n.\n|\nl\n|\n220\n240\n260\n280\n300\n320\n"]], ["block_9", ["230...\n\u2014\n"]], ["block_10", ["300 \n303\n_\n"]], ["block_11", ["315\n310\u2014\n~\u2014\n"]], ["block_12", ["320 \n_\n"]], ["block_13", ["330\n'\nI\u2014r\nT\n'\nI\n'\nl\n'\nI\n"]], ["block_14", [{"image_1": "533_1.png", "coords": [61, 50, 326, 296], "fig_type": "figure"}]], ["block_15", ["8\nQ\n9\nTsc\n"]], ["block_16", ["Crystallization temperature (\u00b0C)\n"]], ["block_17", ["LMW\n"]]], "page_534": [["block_0", [{"image_0": "534_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "534_1.png", "coords": [25, 556, 372, 660], "fig_type": "figure"}]], ["block_2", [{"image_2": "534_2.png", "coords": [28, 92, 240, 219], "fig_type": "figure"}]], ["block_3", [{"image_3": "534_3.png", "coords": [31, 335, 107, 371], "fig_type": "molecule"}]], ["block_4", ["Source: From Mandelkern, L, in Crystallization of Polymers, McGraw-Hill, New York, 1964.\n"]], ["block_5", ["At T1210, AGf:0, but ASf, AVf, and AHf have nonzero values. Figure 12.2a showed how V, S, and\nH (as first derivatives of G) undergo a discontinuous change at a first\u2014order transition, such as\nfusion. For any constant-temperature process,\n"]], ["block_6", ["This fundamental relationship points out that the temperature at which crystal and liquid are in\nequilibrium is determined by the balancing of entropy and enthalpy effects. Remember, it is the\ndz\ufb01erence between the crystal and liquid free energies that is pertinent; sometimes these differ-\nences are not what we might expect.\nTable 13.3 lists representative values of Tm, as well as AHf and ASf per mole of repeat units, for\nseveral polymers. A variety of experiments and methods of analysis have been used to evaluate\nthese data, and because of an assortment of experimental and theoretical limitations the values\nshould be regarded as approximate. We assume Tm g T310. In general, both AHf and ASf may be\nbroken into contributions of H0 and SO, which are independent of molecular weight, and increments\nAHm and AS\u201d for each repeat unit in the chain. Therefore, AHsO + NAHm, where N is the\ndegree\nof polymerization.\nIn\nthe\nlimit of large\nN,\nAHf\u2019\u00e9NAH\ufb02l\nand AsNASt-J,\nso\n"]], ["block_7", ["Polyethylene\n137.5\n8,220\n19.8\ncis 1,4\u2014Polyisoprene\n28\n8,700\n28.9\nPoly(ethylene oxide)\n66\n8,700\n25.1\nPoly(decamethylene sebacate)\n80\n50,200\n142.3\nPoly(decamethylene azelate)\n69\n41,840\n121.3\nPoly(decamethylene sebacamide)\n216\n34,700\n71.1\nP01y(decamethylene azelamide)\n214\n36,800\n75.3\n"]], ["block_8", ["Table 13.3\nValues of Tm, ML], and ASH for Several Polymers\n"]], ["block_9", ["Polymer\nr... (\u201dC)\nAH\u201c (J mol\u2018l)\nA5,, (J K\u201d1 moi\u20141)\n"]], ["block_10", ["and therefore at equilibrium\n"]], ["block_11", ["T3? AHm /ASf\u2018]. Some observations concerning these data sets are listed here:\n"]], ["block_12", ["Figure 13.10\nBehavior of Gibbs free energy near T3? for an idealized crystal to liquid phase transition.\n"]], ["block_13", ["524\nCrystalline Polymers\n"]], ["block_14", ["TOO\u2014AHf\n__ __\n13.3.2\nm\nASf\n(\n)\n"]], ["block_15", ["AG AH TAS\n(13.3.1)\n"]], ["block_16", ["T;\nT\n"]], ["block_17", ["Liquid\n"]]], "page_535": [["block_0", [{"image_0": "535_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "535_1.png", "coords": [11, 539, 311, 694], "fig_type": "figure"}]], ["block_2", ["Thermodynamics of Crystallization\n525\n"]], ["block_3", ["l.\nPolyethylene. The crystal structure of this polymer is essentially the same as those of linear\nalkanes containing 20\u201440 carbon atoms, and the values of Tnn and AH\u201d are what would be\nexpected on the basis of an extrapolation from data on the alkanes. Since there are no chain\nsubstituents or intermolecular forces other than dispersion (London) forces in polyethylene,\nwe shall compare other polymers to it as a reference substance.\n2.\ncis 1,4-Polyisoprene. Although AHfJ is slightly higher than that of polyethylene (on a per\nrepeat unit basis, not per gram), it is still completely reasonable for a hydrocarbon. The\nlower Tm is the result of the value of AS\u201c, which is 50% higher than that of polyethylene.\nThe low melting point of this polymer makes natural rubber a useful elastomer at ordinary\ntemperatures.\n3.\nPoly(ethylene oxide). Although AH\u201d is similar to that of polyethylene, the effect is offset\nby an increase in ASH similar to polyisoprene. The latter may be due to increased chain\n\ufb02exibility in\nthe liquid\ncaused by\nthe regular insertion of ether oxygens\nalong\nthe\nchain backbone.\n4.\nPolyesters. The next two polyesters have AH\u201d values significantly higher than polyethylene.\nOur first thought might be to attribute this to a strong interaction between the polar ester\ngroups. The repeat units of these compounds are considerably larger than in the reference\ncompound, so the AHm values should be compared on a per gram basis. When this is done,\nAH\u201c is actually less than for polyethylene. This suggests that the large value for AH\u201c is the\nresult of a greater number of methylene groups contributing London attraction for the\npolyesters, with the dipole\u2014dipole interaction of ester groups about the same in both liquid\nand crystal and therefore contributing little to AHm. When compared on the basis of the\nnumber of bonds along the backbone, AS\u201c is not exceptional either. Accordingly, is less\nfor these esters than for polyethylene.\n5.\nPolyamides. The next two compounds are the amide counterparts of the esters listed under\nitem (4). Although the values of AH\u201c are less for the amides than for the esters, the values of\n"]], ["block_4", ["Table 13.4\nValues of TH, for Poly(a-olefin) Crystals in Which the\nPolymer Has the Indicated Substituent\n"]], ["block_5", ["The melting points of a series of poly(ot-olefin) crystals were studied. All of the polymers were\nisotactic and had chain substituents of different bulkiness. Table 13.4 lists some results. Use\nEquation 13.3.2 as the basis for interpreting the trends in these data.\n"]], ["block_6", ["Source: From Billmeyer, F.W., in Textbook ofPolymer Science, 2nd ed., Wiley-\nInterscience, New York, 1971.\n"]], ["block_7", ["Tm than to predict it a priori. This state of affairs is not unique to polymers, however. The following\nexample provides another illustration of this type of reasoning.\n"]], ["block_8", ["Substituent\nTm (\u00b0C)\n"]], ["block_9", ["\u2014CH3\n165\n\u2014CH2CH3\n125\n\u2014CH2CH2CH3\n75\n\u2014CH2CH2CH2CH3\n\u201455\n\u2014\u2014CH2CH(CH3)CH2CH3\n196\n\u2014\u2014CH2C(CH3)2CH2CH3\n350\n"]], ["block_10", ["Example 13.2\n"]], ["block_11", [{"image_2": "535_2.png", "coords": [44, 560, 279, 656], "fig_type": "figure"}]], ["block_12", ["These examples show that it is often easier to rationalize an observation or trend with respect to\n"]], ["block_13", ["Tm are considerably higher. This is a consequence of the very much lower values of AS\u201d for\nthe amides. These, in turn, are attributed to the low entropies of the amide in the liquid state,\nowing to the combined effects of hydrogen bonding and chain stiffness from the contribution\nof the following resonance form: \u201d(HN+=CO_)\u2014\u2014\u2014.\n"]]], "page_536": [["block_0", [{"image_0": "536_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The bulkiness of the substituent groups increases moving down Table 13.4. Also moving down the\n"]], ["block_2", ["table, the melting points decrease, pass through a minimum, and then increase again. As is often\nthe case with reversals of trends such as this, there are (at least) two different effects working in\nopposition in these data:\n"]], ["block_3", ["All the polymers compared have similar crystal structures, but are different from polyethylene,\nwhich excludes the possibility for also including the latter in this series. Also note that the isotactic\nstructure of these molecules permits crystallinity in the first place. With less regular microstruc-\nture, crystallization would not occur at all.\n"]], ["block_4", ["In the discussion of Table 13.3, we acknowledged that there might be some uncertainty in the\nvalues of the quantities tabulated, but we sidestepped the origin of the uncertainty. In the next\nsection we shall consider the most important of these areas: the effect of crystal dimensions on the\nvalue of Tm.\n"]], ["block_5", ["2.\nAs the bulkiness of the chain substituents increases, the energy barriers to the\nchain backbone increase. As seen in Chapter 6, this decreases chain flexibility\n"]], ["block_6", ["Whenever a phase is characterized by at least one linear dimension that is small (a few microns or\nless), the properties of the surface begin to make significant contributions to the observed behavior.\nIn contrast, most thermodynamic analyses are conducted on the assumption of bulk (i.e., effect-\nively infinite) phases. As illustrated in Figure 13.1, lamellae tend to have thicknesses on the order\nof 100 A, and thus surface effects can be substantial. The following summary of generalizations\nabout these crystals will be helpful:\n"]], ["block_7", ["Solution\n"]], ["block_8", ["526\nCrystalline Polymers\n"]], ["block_9", ["state. It is this flexibility that permits the molecules to experience large con-\nformations and therefore have high entropies. If the \ufb02exibility is reduced, the entropy change\non melting is less than it would otherwise be. Accordingly, the entropy of fusion decreases\nmoving down the table.\n3.\nSince Tm AHf/ASf, the observed behavior of this series of polymers may be understood as a\ncompetition between these effects. For the smaller substituents, the effect on AHf dominates\nand Tm decreases with bulk. For larger substituents, the effect on A5} dominates and Tm\nincreases with bulk.\n"]], ["block_10", ["1.\nThe dimensions of the lamellae perpendicular to the smallest dimension depend on the\nconditions of the crystallization, but are many times larger than the thickness of a well-\ndeveloped crystal.\n2.\nThe chain direction within the crystal tends to be along the short dimension of the crystal,\nindicating that the molecule folds back and forth, fire-hose fashion, with successive layers of\nfolded molecules accounting for the lateral growth of the platelets. The section of chain that\ntraverses a lamella once is called a stem.\n3.\nA crystal lamella does not consist of a single molecule, nor does a molecule need to reside\nexclusively in a single lamella.\n"]], ["block_11", ["4.\nThe loop formed by the chain as it emerges from the crystal, turns around, and reenters the\ncrystal may be regarded approximately as amorphous polymer, but is insufficient to account\nfor the total amorphous content of most crystalline polymers.\n"]], ["block_12", ["13.4.1\nSurface Contributions to Phase Transitions\n"]], ["block_13", ["1.\nAs the bulkiness of the substituents increases, the chains are prevented from coming into\nintimate contact in the crystal. The intermolecular forces that hold these crystals together are\nall London forces, and these become weaker as the crystals loosen up owing to substituent\nbulkiness. Accordingly, the value for the heat of fusion decreases moving down Table 13.4.\n"]], ["block_14", ["13.4\nStructure and Melting of Lamellae\n"]]], "page_537": [["block_0", [{"image_0": "537_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["To develop this model into a quantitative relationship between Tm and the thickness of the\ncrystal, we begin by realizing that for the transition crystal\u2014>1iquid, AGf is the sum of two\ncontributions. One of these is AGOO, which applies to the case of a crystal of infinite size\n"]], ["block_2", ["To develop a more quantitative relationship between particle size and Tm, suppose we consider the\nmelting behavior of the cylindrical crystal sketched in Figure 13.11. Of particular interest in this\nmodel is the role played by surface effects. The illustration is used to define a model and should not\nbe taken too literally, especially with respect to the following points:\n"]], ["block_3", ["As noted above, since the polymer crystal habit is characterized by plates whose thickness\nis small, surface phenomena are important. During the early development of the crystal, the\nlateral dimensions are also small and this effect is even more pronounced. The key to understand-\ning this fact lies in the realization that all phase boundaries possess surface tension, and that\nthis surface tension re\ufb02ects the Gibbs free energy stored per unit area of the phase boundary. As\na qualitative illustration,\nconsider cutting\na piece of polymer into two\nalong\na\nselected\nplane. Before cleavage, there were cohesive (attractive) interactions across the plane, which are\nnow lost. This energy per unit area becomes the surface energy of the newly exposed material.\nNow place two different materials in contact across a plane; unless their surface energies happen to\nbe identical, there would be an interfacial energy, or surface tension, y. Now suppose we consider a\nspherical phase of radius r, density p, and surface tension y. The total surface free energy\nassociated with such a particle is given by the product of y and the area of the sphere, or\ny(4qrr2). The total mass of material in the sphere is given by the product of the density and the\nvolume of the sphere, or p(47rr3/3). The ratio of the former to the latter gives the Gibbs free\nenergy arising from surface considerations, expressed per unit mass; that is, the surface Gibbs\nfree energy per unit mass is 3y/pr. Since 7 is small compared to most other chemical and physical\ncontributions to the free energy, surface effects are not generally considered when, say, the AG\u201d\nof formation is quoted for a substance. The above argument shows that this becomes progressively\nharder to justify\nas the particle\nsize\ndecreases.\nThe emergence of a new phase implies\nstarting from an r value of zero in the argument above, and the surface contribution to the\nenergy becomes important indeed. (Since two phases with their separating surface must already\nexist\nfor\ny\nto\nhave\nany\nmeaning,\nwe\nare\nspared\nthe\nembarrassment\nof\nthe\nsurface\nfreeenergy becoming infinite at r: 0.) Nevertheless, it is apparent that the effect of the surface\nfree-energy contribution is to increase the total G. Inspection of Figure 13.10 shows that an\nincrease in the G value for the crystalline phase arising from its small particle size has the\neffect of shifting Tm to lower temperatures. The smaller the particle size, the bigger the effect.\nThis is the origin of all superheating, supercooling, and supersaturation phenomena: an equilibrium\ntransition is sometimes overshot because of the kinetic difficulty associated with the initiation\n(nucleation) of a new phase. Likewise, all nucleation practices\u2014cloud seeding, bubble chambers,\nand the use of boiling chips\u2014are based on providing a site on which the emerging phase can\ngrow readily.\n"]], ["block_4", ["1.\nThe geometry of the cylinder is a matter of convenience. Except for numerical coefficients, the\nresults we shall obtain will apply to plates of any cross\u2014sectional shape.\n2.\nThe thickness of the plate, although small, is greater than the few repeat units shown.\n3.\nThe specific nature of the reentry loops is not the point of this illustration. The sketch shows\nboth hairpin turns and longer loops. Problem 6 at the end of the chapter examines the actual\nnature of the reentry loop.\n"]], ["block_5", ["13.4.2\nDependence of Tm on Lamellar Thickness\n"]], ["block_6", ["5.\nPolymer chain ends disrupt the orderly fold pattern of the crystal, and tend to be excluded from\nthe crystal and relegated to the amorphous portion of the sample. The same is true of\nstereochemical or microstructural defects, or comonomers.\n"]], ["block_7", ["Structure and Melting of Lamellae\n527\n"]]], "page_538": [["block_0", [{"image_0": "538_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "538_1.png", "coords": [20, 47, 278, 288], "fig_type": "figure"}]], ["block_2", [{"image_2": "538_2.png", "coords": [31, 616, 151, 663], "fig_type": "molecule"}]], ["block_3", ["When this infinite phase is in equilibrium with the melt, AG?)O \u2014> 0 and T \u2014> T3,\u201c. Accordingly,\nwe can solve Equation 13.4.4 for A500 =All-11$;j HT;0 and substitute this back into Equation\n13.4.4:\nr\nAG? :AH$(1\"TTO)\n(13.4.5)\n"]], ["block_4", ["For the bulk effect, we proceed on the basis of a unit volume (subscript V) and immediately write\n"]], ["block_5", ["Now each of these can be developed independently.\nAs in the qualitative discussion above, let y be the Gibbs free energy per unit area of the\ninterface between the crystal and the surrounding liquid. This is undoubtedly different for the\nedges of the plate than for its faces, but we shall not worry about this distinction. The area of each\nof the circular faces of the cylinder is 7W2, and the area of the edge is 27rr\u20ac, where r is the radius of\nthe face and E is the length of the side as shown Figure 13.11. Since surface is destroyed by the\nmelting process, the net contribution to AGf is\n"]], ["block_6", ["(superscript 00); the other, AG, arises specifically from surface effects (superscript s), which\nre\ufb02ect the finite size of the crystal:\n"]], ["block_7", ["and\n"]], ["block_8", ["Figure 13.11\nIdealized representation of a polymer crystal as a cylinder of radius r and thickness E.\n"]], ["block_9", ["528\nCrystalline Polymers\n"]], ["block_10", [{"image_3": "538_3.png", "coords": [37, 471, 240, 496], "fig_type": "molecule"}]], ["block_11", ["A02; :AH? TASS?\n(13.4.4)\n"]], ["block_12", ["AGf :AG\u00b0\u00b0 + AGS\n(13.4.1)\n"]], ["block_13", ["AGOO mzeacif\n(13.4.3)\n"]], ["block_14", ["AG5 [2777'2 + 2771*617/ \u201427rr2y(1 + g)\n(13.4.2)\n"]]], "page_539": [["block_0", [{"image_0": "539_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Note that this equation is dimensionally correct, as 7 has units of energy area\u20141 and AH\ufb01}O has units\nenergy volume\u20141. Therefore the units of EAHQE and y cancel, as do the units of E and r, leaving\nonly temperature units on both sides of the equation. All of the quantities on the right-hand side of\nthe equation are positive (A1130 is the heat of fusion), which means that T3,? > Tm, as anticipated.\nThe difference AT between a thermodynamic boundary and the temperature of interest is often\nreferred to as the undercooling; as derived in Equation 13.4.8 it reflects the melting point\ndepression due to particle size. Several limiting cases of this equation are of note:\n"]], ["block_2", ["Equation 13.4.9 shows that a direct proportionality relationship should exist between crystal\nthickness 1? and the ratio YES/AT; a plot of E versus T310/AT should result in a straight line of zero\nintercept with a slope proportional to y/AHSF. Figure 13.12 shows such a plot for polyethylene in\nwhich T? was taken to be 137.5\u00b0C and the E values were determined by x-ray scattering. While\nthere is considerable and systematic divergence from the predicted form at large undercoolings, the\ndata show a linear relationship for the higher\u2014temperature region. In the following example we\nanalyze the linear portion of Figure 13.12 in terms of Equation 13.4.9.\n"]], ["block_3", ["This gives the value of AG; at any temperature in terms of the two parameters AHS? and T33.\nCombining Equation 13.4.1 through Equation 13.4.3 and Equation 13.4.5 enables us to write\n"]], ["block_4", ["or\n"]], ["block_5", ["Use Equation 13.4.9, the results in Figure 13.12, and the data in Table 13.3 to estimate a value for 'y\nfor polyethylene. Figure 13.5 shows the unit cell of polyethylene; the equivalent to two chains\nemerges from an area 0.740 by 0.493 nm2. On the basis of the calculated value of y and the\ncharacteristics of the unit cell, estimate the free energy of the fold surface per mole of repeat units.\n"]], ["block_6", ["When the value of this AG is zero, we have the actual melting point of the crystal of finite\ndimension Tm. That is,\n"]], ["block_7", ["1.\nIf y: 0, AT: 0, regardless of particle size. This is not likely to apply, however, since chains\nemerging from a crystal face either make a highly constrained about-face and reenter the\ncrystal or meander off into the liquid from a highly constrained attachment to the solid. In\neither case, a surface free-energy contribution is inescapable.\n2.\nAs r a 0, AT6 00, showing that the lateral dimensions of the plate are critical for very small\ncrystals. This makes the crystal nucleation event especially crucial.\n3.\nAs 1\u2018 \u2014> 00, which describes well\u2014developed crystals, Equation 13.4.8 becomes\n"]], ["block_8", ["Structure and Melting of Lamellae\n529\n"]], ["block_9", ["Example 13.3\n"]], ["block_10", [{"image_1": "539_1.png", "coords": [36, 189, 211, 243], "fig_type": "molecule"}]], ["block_11", [{"image_2": "539_2.png", "coords": [46, 146, 231, 181], "fig_type": "molecule"}]], ["block_12", ["2\n7\nt?\n___ TOO _Tm :_\n.2\nT00\n.\n.\nAT\nm\n6: $0049)\u201d, \n(1348)\n"]], ["block_13", ["TOO\nI'I'l\nAGf (nr2\u20ac)AH{'? (1 L) 27rr2y(1 + 9\n(13.4.6)\n"]], ["block_14", ["which shows that an undercooling is still important because of the platelike crystal habit of\npolymers with limited crystal dimensions along the chain direction. Equation 13.4.9 is often\nreferred to as the Thompson\u2014Gibbs equation.\n"]], ["block_15", ["3/\n1\nA1\":2TDO\n\u2014-\nm AHQE3 E\n(13.4.9)\n"]], ["block_16", [{"image_3": "539_3.png", "coords": [95, 87, 256, 116], "fig_type": "molecule"}]], ["block_17", ["Too \nm\ng\n"]]], "page_540": [["block_0", [{"image_0": "540_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "540_1.png", "coords": [30, 46, 344, 270], "fig_type": "figure"}]], ["block_2", ["Therefore y 1/2(13.7 x 10\"10 m) (2.8 x 108 J m\u20143) 0.192 J m\u2018z. From the data on the unit cell\n"]], ["block_3", ["Figure 13.12\nCrystal thickness versus Tm/AT for polyethylene. (Reprinted from Mandelkem, L., Crystal-\nlization of Polymers, McGraw\u2014Hill, New York, 1964. With permission.)\n"]], ["block_4", ["Equation 13.4.9 predicts a straight line of zero intercept and slope of 27/AH8? when E is plotted\nversus Trof/AT. The solid line in Figure 13.12 has a slope of 650 194/51: 13.7 .4. Therefore\n2y/AH8? :13.7 X 10\u201810 m. The value of AH\u201d given in Table 13.3 is used for AH\ufb01\u2019,O after the\nfollowing change of units:\n"]], ["block_5", ["Before concluding this section, there is one additional thermodynamic factor to be mentioned,\nwhich also has the effect of lowering Tm. The specific effect we consider is that of chain ends (and\ntherefore the number-average molecular weight), but the role they play is that of an \u201cimpurity\u201d\nfrom the viewpoint of crystallization. As such the treatment is similar to the effect of a solute on\nany colligative property, such as the osmotic pressure considered in Section 7.4. Furthermore,\nother \nimpurities\u201d such as comonomers or low molecular-weight species can be treated in a\n"]], ["block_6", ["Although it applies to a totally different kind of interface, the value of y calculated in the example\nis on the same order of magnitude as the y value for the surface between air and liquids of low\nmolecular weight.\n"]], ["block_7", ["Solution\n"]], ["block_8", ["530\nCrystalline Polymers\n"]], ["block_9", ["13.4.3\nDependence of Tm on Molecular Weight\n"]], ["block_10", ["(nm)\n"]], ["block_11", ["Thickness\n"]], ["block_12", ["22J\n1\n1\n1\n16\n3\nM3028\n0\n\"we\ng\nO\u2014CT22.8x108Jm\u20193\nmole)<\n28g\nXEEFX\n1m\n"]], ["block_13", ["40\u2014\n08\nA\no.\n3\nA\n30- .0\n\u2018\na\n\u2018A\nA\n20-\nA\n10-\n"]], ["block_14", ["80\u2014\n"]], ["block_15", ["70-\n"]], ["block_16", ["60-\n"]], ["block_17", ["01O \n"]], ["block_18", ["0.192 J x\n(0.074\n"]], ["block_19", ["m2\nunit cell\nX\n1013 nm2\nX\n2 molecules\n"]], ["block_20", ["l\n|\n|\n|\nl\nl\nl\n|\n|\n|\nl\nl\n|\n16111621263136414651566166\nr:\nAT\n"]], ["block_21", [".02\n1023\nl\nl\nx \nx\nmo ecu es = 2.1 kJ/mol\nmol\n"]], ["block_22", ["X 0.493 nm2\n1 m2\n"]], ["block_23", [">\n"]], ["block_24", ["1 unit cell\n"]]], "page_541": [["block_0", [{"image_0": "541_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "541_1.png", "coords": [33, 360, 188, 398], "fig_type": "molecule"}]], ["block_2", ["where Tm is the melting point of the polymer under consideration. Equation 13.4.13 indicates that a\nfreezing point depression is to be expected from an increased concentration of chain ends.\nQualitatively, at least, the presence of other types of defects is also expected to lower Tm.\nRemember that in the present discussion TI? is the melting point of a polymer of infinite molecular\nweight without regard to the crystal size, whereas in Equation 13.4.8 it was the melting point for a\ncrystal of infinite dimension without regard to molecular weight. The two effects are therefore\ncomplementary, and both are Operative if both particle dimension and molecular weight are small\nenough to lower the freezing point appreciably.\nThroughout this section we have focused attention on thermodynamic melting points. The\nsame thermodynamic arguments can be applied to the raising and sharpening of this transition\ntemperature through annealing. When a crystal is maintained at a temperature between the\ncrystallization temperature and the equilibrium melting point, an increase in Tm is observed.\nThis may be understood in terms of the melting of smaller, less perfect crystals and the redisposi\u2014\ntion of the polymers into larger, more stable crystals. This is analogous to the procedure of\ndigesting a precipitate before filtration. There is more to the story than this, however. The digestion\nanalogy would suggest that those crystals that are enlarged simply add more folded chains around\ntheir perimeter. In fact, x-ray diffraction studies reveal progressive thickening of lamellae with\nannealing, that is, as T1,, increases. This requires large-scale molecular reorganization throughout\nthe crystal. Such rearrangements apparently require the molecule to snake along the chain axis,\nwith segments being reeled in and out across the crystal surface. The process of annealing,\ntherefore, not only involves crystal thickening, but also provides the opportunity to work out\nkinks and defects.\n"]], ["block_3", ["This equation follows from the definition of the \u201cactivity\u201d of middle segments, am (see Equation\n7.1.13) and the approximation of ideality (a1,, xm) as the impurity concentration vanishes (xe << 1).\nEquation 13.4.11 is the analog of Equation 7.4.4 for the osmotic pressure. We can now insert this\nexpression into Equation 13.4.5, but now applied to melting a crystal of chains of molecular weight\nM at temperature Tm compared to a crystal of infinite chains:\n"]], ["block_4", ["The proportion of chain ends increases with decreasing molecular weight, and hence for a linear\nchain (two ends) x(3 2- ZMO/Mn, where M0 is the molecular weight of the repeat unit. Now we take\nthe molar Gibbs free energy of a finite chain crystal compared to that of an infinite chain crystal\n(xmz 1):\n"]], ["block_5", ["or, after some rearranging,\n"]], ["block_6", ["similar fashion. In this context the repeat units in a polymer may be divided into two classes: those\nat the ends of the chain (subscript e) and the others, which we view as being in the middle\n(subscript m) of the chain. The mole fraction of each category in a sample is xe and xm,\nrespectively. Since all segments are of one type or the other,\n"]], ["block_7", ["Structure and Melting of Lamellae\n531\n"]], ["block_8", [{"image_2": "541_2.png", "coords": [37, 304, 155, 341], "fig_type": "molecule"}]], ["block_9", [{"image_3": "541_3.png", "coords": [39, 183, 252, 237], "fig_type": "molecule"}]], ["block_10", ["xm :1\u2014 La\n(13.4.10)\n"]], ["block_11", ["Tm\nAHf(1\u2014\u201cj:;> :Rme\n(13.412)\n"]], ["block_12", ["G(xm) GO:m 1) \u2014RTln am :\u2014RT 1n xm\n: \u2014RTln(1\u2014xe)mRTxe\n(13.4.11)\n"]], ["block_13", ["m\n"]], ["block_14", ["film? 2M0\n"]]], "page_542": [["block_0", [{"image_0": "542_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Current understanding of the structure of chain-folded lamellae has been greatly facilitated by\nstudy of solution-grown single crystals. There are many parameters of a single crystal that are\nof interest, in addition to the structure of the unit cell itself and the lamella thickness. These include\nthe shape and size of the crystal in the other two dimensions, the orientation of the unit cell relative\nto the thickness direction, and the nature of the fold surface. Valuable information about all of\nthese features has been obtained with electron microscopy (EM) techniques, and so we will\nprovide a brief introduction to this important characterization tool.\nThe electron microscope uses the de Broglie waves associated with accelerated electrons to\nproduce an image in much the same way as visible light produces an image in an optical microscope.\nElectromagnets and imposed electrical fields function as lenses for the electron beam and the image\nis formed on a phosphorescent screen, photographic plate, or charge-coupled device (CCD) camera.\nThe wavelength of an electron under typical operating conditions in an electron microscope is on the\norder of A as 0.03 A; it depends primarily on the energy of the electrons, which in turn is dictated by\nthe accelerating voltage. In all types of microscopy it is the resolving power rather than the\nmagnification per se, which is the limiting factor. Waves are diffracted from the edges of illuminated\nbodies and this diffraction blurs their boundaries. The resolving power measures the minimum\nseparation between objects that will produce discernibly different images in a microscope. In a well-\ndesigned instrument this separation is on the order of the wavelength of the illuminating radiation.\nTherefore, the resolving power of an electron microscope is potentially smaller by some five orders\nof magnitude than that achieved by optical microscopes. In reality, imperfections in electron optics\nlimit the actual resolution to something closer to 100 A, but nevertheless this resolution matches\natomic dimensions, and so EM is extremely useful.\nAlthough the concepts of wave optics apply equally well to visible light, x-rays, and electrons,\nthere are important differences that affect the information that can be obtained,\nand the\nexperimental design. First, the incident electrons carry charge, and so interact relatively strongly\nwith matter. Consequently, the scattering cross-section is high, which means that an electron does\nnot have to go very far through a material before being scattered. The primary consequence of this\nfact is that samples must be very thin (perhaps 100 nm) in order for much of the electron beam to\nbe transmitted. Second, in one way or another the image obtained is based on the ability of\nthe material to scatter or diffract the electron beam. In Chapter 8 we discussed how a perfectly\nuniform material would not scatter light. However, in that case a typical wavelength is 5000 A, and\na small\u2014molecule liquid or glass could be almost completely transparent; in contrast, the electron\nwavelength is so small that no material made up of nuclei and electrons can appear transparent.\nThird, the short wavelength of electrons is determined by their high energy. One consequence is\nthat the electron beam is likely to damage the sample. Part of the art of EM is to limit the sample\nexposure while maintaining sufficient \ufb02ux to obtain a good image.\nIn microscopy, one obtains some kind of direct picture of the structure in question, a real-space\nimage. This should be contrasted with scattering or diffraction, in which the structural information\nis contained in the angle-dependent intensity, providing a so-called reciprocal-space image.\nDiffraction experiments suffer from the inversion problem, namely that there is no unique way\nto take the diffraction pattern and invert it into the actual structure. This problem becomes more\nacute as the structure under examination becomes more complicated. All other things being equal,\ntherefore, a real-space image is preferable. Of course, all other things are often not equal. There are\nseveral potential limitations to BM in its application to polymers. Among these are:\n"]], ["block_2", ["l.\nThe need for very thin samples, with its attendant challenges in sample preparation.\n2.\nThe generally low contrast between different monomers (in a mixture or copolymers) or\ndifferent structures in a single polymer. This arises because the chemical compositions\n(C, H, 0, maybe N...) and material densities (typically 0.9\u20141.1 g cc\u201c) of most organic\npolymers are similar.\n"]], ["block_3", ["532\nCrystalline Polymers\n"]], ["block_4", ["13.4.4\nExperimental Characterization of Lamellar Structure\n"]]], "page_543": [["block_0", [{"image_0": "543_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Fortunately, most of these issues are of less importance when examining polymer single crystals\ncollected from a solution crystallization process; the crystals are inherently thin, \ufb01xed in structure,\nand small in lateral dimensions.\nThere are a variety of different ways in which an image can be generated in EM, and we brie\ufb02y\nidentify some of the terminology that is encountered. Two procedures of electron microscopy\n\ufb01nd particular applicability in the study of polymer crystals: shadow casting and dark-field\noperation. Shadow casting is used to improve the contrast between a sample and its background\nand between various details of the sample surface. Because polymer crystals are so thin and mostly\nconsist of atoms of low atomic number, some sort of contrast enhancement is important. In the\nshadowing method the sample is placed in an evacuated chamber and a heavy metal is allowed to\nevaporate in the same chamber. The position of the metal source is such that the metal vapor strikes\nthe sample at an oblique angle and condenses on the cool surface. The thin metal film thus formed\nliterally casts shadows, which enhance the image of the sample. If the angle of incidence of the\nheavy metal beam is known, the thickness of a crystal or the height of surface protuberances can be\ndetermined from the length of the shadow by simple trigonometry.\nIn dark-\ufb01eld electron microscopy it is not the transmitted beam that is used to construct an\nimage but, rather, a beam diffracted from one facet of the object under investigation. One method\nfor doing this is to shift the aperture of the microscope so that most of the beam is blocked and only\nthose electrons scattered into the chosen portion of the diffraction pattern contribute to the image.\nThis decreases the intensity of the illumination used to produce the dark-\ufb01eld image and therefore\nrequires longer exposure times, with the attendant modification or even degradation of the\npolymer. Nevertheless, dark-field operation distinguishes between portions of the sample with\ndifferent orientations, as the diffracted beams will appear in different directions, and therefore\nproduces a more three-dimensional representation of the sample. Bright-field operation, on the\nother hand, utilizes the transmitted main beam to form the image.\nFigure 13.13a through Figure 13.13c provide examples of shadow-contrast, bright-field, and\ndark-field transmission electron micrographs of solution-grown polyethylene crystals, respect-\nively. The effect of shadowing is evident in Figure 13.13a, where those edges of the crystal that\ncast the shadows display sharper contrast. The roughly diamond shape of the crystal lamella is also\nclearly evident. The cracks evident in the larger crystals result from the fact that in solution the\ncrystals actually form hollow pyramids, whereas they are \ufb02attened onto a viewing grid for the EM.\nThere is also evidence of multilamellar growth in the lower-right comer of the image. Figure 13.3b\nshows the clear outlines of a \u201ctruncated\u201d diamond, and the same kind of crack as in Figure 13.3a.\nThe dark-\ufb01eld image of the same crystal in Figure 13.3c is particularly revealing, because now the\nsingle crystal is clearly shown to be made up of four sectors, two of which are dark because the\nunit-cell orientation is such as to diffract the electrons away from the detection aperture. This\nsectorization and the hollow pyramid form are illustrated schematically in Figure 13.14.\nThe electron micrographs of Figure 13.13 are more than mere examples of EM technique. They\nare the \ufb01rst occasion we have had to actually look at single crystals of polymers. Although there is\n"]], ["block_2", ["3.\nThe need for selective staining techniques (using heavy atoms such as Os and Ru) to generate\ncontrast.\n4.\nThe inability to view the samples directly under different conditions of interest, such as\ntemperature, pressure, or \ufb02ow.\n5.\nThe need for sample fixation to prevent changes in structure occurring between the original\nsample and the sample that is imaged in the microsc0pe.\n6.\nThe likelihood of electron\u2014beam damage.\n7.\nAn EM image represents a projection completely through the sample, so care must be taken in\ninterpreting features that may represent objects that are separated along the beam direction.\n8.\nA tightly focussed electron beam will only interrogate a small portion of the sample (less than\n"]], ["block_3", ["Structure and Melting of Lamellae\n533\n"]], ["block_4", ["1 um3), so care must be taken to establish how representative a given image is.\n"]]], "page_544": [["block_0", [{"image_0": "544_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "544_1.png", "coords": [19, 523, 206, 624], "fig_type": "figure"}]], ["block_2", [{"image_2": "544_2.png", "coords": [29, 328, 259, 436], "fig_type": "figure"}]], ["block_3", ["534\nCrystalline Polymers\n"]], ["block_4", ["casting. (Reprinted from\n"]], ["block_5", ["(b)\n"]], ["block_6", ["Figure 13.13\nElectron micrographs \n"]], ["block_7", ["With permission.) \n"]], ["block_8", ["(b) Bright-field \n"]], ["block_9", ["1906, 1960. With \n"]], ["block_10", ["different fold directions.\n"]], ["block_11", ["Figure 13.14\nSchematic side \n"]], ["block_12", ["_\n.\n"]], ["block_13", ["(C)\n"]], ["block_14", [{"image_3": "544_3.png", "coords": [204, 524, 401, 631], "fig_type": "figure"}]]], "page_545": [["block_0", [{"image_0": "545_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "545_1.png", "coords": [16, 509, 379, 651], "fig_type": "figure"}]], ["block_2", ["The foregoing is by no means a comprehensive list of the remarkable structures formed by the\ncrystallization of polymers from solution. The primary objective of this brief summary is the\n"]], ["block_3", ["2.\nCrystallization conditions such as temperature, solvent, and concentration can in\ufb02uence\ncrystal form. One such modification is the truncation of the points at either end of the long\ndiagonal of the diamond-shaped crystals seen in Figure 13.13b and Figure 13. 13c. The facets\nof the diamond-shaped crystal correspond to { 110} planes of the polyethylene unit cell (see\nFigure 13.5), whereas the truncated sections in Figure 13.13b and Figure 13.13c are {100}\nplanes. If the rate of crystallization onto { 100} planes is much greater than onto { 110] planes,\nthen the { 100] facets will disappear and a diamond will result. Twinning and dendritic growth\nare other examples of such changes of habit, and these features can usually be attributed to the\nrelative growth rates for different crystallographic faces.\n3.\nHollow pyramids are thoroughly documented and fairly well understood. The underlying\nfactor is the nature of the fold as a chain exits and reenters the lamella. If we consider two\nstems plus one fold to form a \u201cU,\u201d defining a fold plane, it turns out that the fold does\nnot remain in this plane, but is tilted. This can be understood from Figure 13.5. The fold\nplane is a (110) plane, and consequently the all-trans orientation for each adjacent stem is\nrotated by 90\u00b0. Each successive fold plane is displaced vertically by one CH2 unit, which leads\nto the pyramidal form; the chain axis is perpendicular to the base of the pyramid, not to the\nfold surface.\n4.\nThe nature of the chain folding at the fold surface has been a subject of great interest (and no\nlittle controversy). The implication of the cartoon in Figure 13.14 is that the folding is very\nregular, and immediate in the sense that an emerging chain folds back directly into the crystal\n"]], ["block_4", ["a great deal to be learned from studies of single crystals by EM, we shall limit ourselves to just a\nfew observations:\n"]], ["block_5", ["Structure and Melting of Lamellae\n535\n"]], ["block_6", ["Figure 13.15\nSchematic illustration of adjacent reentry (left side of crystal) and random reentry (right side).\n"]], ["block_7", ["1.\nSingle crystals such as those shown in Figure 13.13 are not observed in crystallization from the\nbulk. Crystallization from dilute solutions is required to produce single crystals with this kind\nof macroscopic perfection. Polymers are not intrinsically different from low molecular-weight\ncompounds in this regard.\n"]], ["block_8", ["as the adjacent stem in the fold plane. This limiting behavior is known as adjacent reentry, and\nmay be contrasted with the opposite extreme of random reentry, or the switchboard model, as\nshown in Figure\n13.15. For solution-grown single crystals adjacent reentry is certainly\nprevalent. However, for crystals grown from the melt, the evidence favors a much more\nrandom folding process; in particular, the radius of gyration of a single chain in the melt\n(measured by neutron scattering) changes little on crystallization, which suggests that it enters\nand departs from several lamellae (see Figure 13.2).\n"]]], "page_546": [["block_0", [{"image_0": "546_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "546_1.png", "coords": [12, 428, 377, 639], "fig_type": "figure"}]], ["block_2", ["In this section our objective is to introduce the basic factors that govern the rate of growth of\npolymer crystals, once a sample has been cooled to a temperature below T310. Some of these factors\npertain to many phase transitions, whereas others are particular to polymer crystallization. We will\nemphasize the rate of growth of individual lamellae, whether in bulk or in solution, as the\nfundamental process of importance. In Section 13.7 we will return to crystallization kinetics, but\nfrom a broader perspective; there we will consider the rate at which a macroscopic sample\nbecomes (semi)crystalline.\nBefore we begin a more systematic treatment, consider the following illustrative example. If we\ntake a polymer sample that has been allowed to crystallize at some temperature TC, we will \ufb01nd a\npredominant average lamellar thickness. What determines this thickness? The answer is a com\u2014\nbination of thermodynamics and kinetics, but mostly the latter. Thermodynamics tells us that only\nlamellae that have a lower free energy than the liquid state can grow spontaneously. In the previous\nsection, we saw that the melting temperature of lamellae increased with thickness, so we may\nconclude that as TC decreases below T310, we will progressively increase the range of smaller\nlamellar thicknesses that may grow. The role of kinetics is to dictate which of the thermodynam-\nically allowed values of E is observed, namely, the one that grows most rapidly. As a rule, thinner\nlamellae tend to grow more rapidly, so as TC is decreased, the observed I? will decrease. The\nresulting lamellae must therefore be viewed as metastable, because a crystal with larger 1? should\nhave a lower free energy.\nThis general principle is beautifully illustrated by the isothermal crystallization data for modest\nmolecular-weight poly(ethylene oxide) shown in Figure 13.16. These single crystals were grown in\nsolution, and the dominant morphology is indicated on the plot. As TC decreases, the rate of\ncrystallization increases, but with abrupt jumps. Each jump in rate is associated with a discrete\nchange in E, from straight\u2014chain lamellae to once-folded, then from once-folded lamellae to twice-\nfolded, etc. The particular temperatures at which the jumps in crystallization rate occur can be\n"]], ["block_3", ["Figure 13.16\nCrystal growth rate for low molecular\u2014weight poly(ethylene oxide) crystallized from solution.\n(Reproduced from Strobl, G., The Physics of Polymers, Springer, Berlin, 1996. With permission.)\n"]], ["block_4", ["verification that single crystals can be formed, and characterized in detail, not only by x-ray\ndiffraction, but also by electron microscopy.\n"]], ["block_5", ["536\nCrystalline Polymers\n"]], ["block_6", ["13.5\nKinetics of Nucleation and Growth\n"]], ["block_7", ["10*? \nN\nv\n(cm\n"]], ["block_8", ["-\n"]], ["block_9", ["109L\n0\n10\u201410\nm\nI\nI\nn\nI\nI\nI\nI\nI\n46\n48\n50\n52\n54\n56\n58\n6O\n62\n64\n"]], ["block_10", ["10\u20183\n[\u2014\n10\u20144 \n[U\n"]], ["block_11", ["10*5 \n\\\n10\"5 \n"]], ["block_12", ["10\u20143\u2014\n"]], ["block_13", ["T(\u00b0C)\n"]]], "page_547": [["block_0", [{"image_0": "547_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "547_1.png", "coords": [32, 409, 350, 640], "fig_type": "figure"}]], ["block_2", ["AG\n"]], ["block_3", ["Figure 13.17\nDependence of free energy on drop size according to Equation 13.5 .1, illustrating the radius\nof the critical nucleus.\n"]], ["block_4", ["Crystallization, like many other first\u2014order phase transitions, proceeds by the process known as\nnucleation and growth. Nucleation refers to the appearance of domains of the new phase that are\nsufficiently large to become stable; recalling the discussion in Section 13.4.1, this amounts to the\ndomains becoming large enough for the bulk energy gain of the new phase to outweigh the\nunfavorable surface energy. In many transitions, nucleation is the rate\u2014limiting step. The subsequent\ngrowth of the particles will be considered in the following section. Nucleation is commonly\nclassified as either heterogeneous or homogeneous. The former denotes the situation where a foreign\nparticle, impurity, or surface provides a site for facile nucleation, whereas the latter indicates the\nspontaneous formation of nuclei by random \ufb02uctuations. In practical situations heterogeneous\nnucleation is almost always more important, and often dominant; indeed, in many polymer processes\nnucleating agents such as talc powder are added to accelerate crystallization. However, in the\nlaboratory the homogenous case is also important and furthermore a simple treatment of homogen-\nous nucleation will allow us to understand a good deal about the subsequent growth processes as well.\nWe begin by adapting our treatment of surface effects in Section 13.4.2 to nucleation. The free\nenergy of a spherical droplet with radius R of the new phase, AG, can be written as the sum of a\nsurface term and a volume term (see Equation 13.4.1):\n"]], ["block_5", ["where y is the surface energy and AGV is the free energy change per unit volume; note that when\nT < Tm, AGV is negative. Figure 13.17 illustrates the functional form of Equation 13.5 .1, with the\n"]], ["block_6", ["easily interpreted based on the foregoing discussion. Thus 635\u00b0C is the temperature below which\nstraight-chain lamellar crystals have a lower free energy than the liquid, and so they grow. Then\n595\u00b0C is the point at which once\u2014folded lamellae become possible, and they grow much more\nrapidly than the straightmchain lamellae. Similarly, below about 565\u00b0C twice-folded lamellae\nbecome possible, and because they grow significantly more rapidly than straight\u2014chain or once-\nfolded lamellae, they become the predominant form.\n"]], ["block_7", ["13.5.1\nPrimary Nucleation\n"]], ["block_8", ["Kinetics of Nucleation and Growth\n537\n"]], ["block_9", ["4\nAG(R) = 477R27+\u00a7R3AGV\n(13.5.1)\n"]], ["block_10", ["Critical nucleus, Ff\"\n"]], ["block_11", ["F?\n"]]], "page_548": [["block_0", [{"image_0": "548_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "548_1.png", "coords": [31, 256, 210, 307], "fig_type": "molecule"}]], ["block_2", ["where K denotes a collection of temperature-insensitive quantities. The temperature dependence is\ncomplicated in detail, but the qualitative consequences of the two terms are clear. The transport\nterm indicates that at low temperatures (T\u2014\u2014>T0) the rate will go to zero, as nothing can move.\nHowever, we recall from Chapter 12 that the glass\u2014transition temperature is usually significantly\nbelow the melting temperature, so this effect can be avoided. The barrier term indicates that as\ntemperature decreases, the rate will increase, because TQM\")2 will increase. This is a simple\nconsequence of an increased thermodynamic driving force to crystallize as the undercooling\nincreases. From Equation 13.5.2 we can see that the critical nucleus shrinks as the crystallization\ntemperature decreases, so nuclei are easier to form.\n"]], ["block_3", ["essential feature that there is a special size, R*, where AG has a positive maximum. For the new\nphase to form, droplets must somehow grow larger than R*, so a droplet of this size is called a\ncritical nucleus. Once a drop exceeds this size, addition of further molecules will only decrease\nAG, and therefore growth is spontaneous. In contrast, it is thermodynamically \u201cuphill\u201d for a\ndroplet smaller than R* to grow. By differentiating Equation 13.5.1 with respect to R, and setting\nthe result equal to zero, we arrive at an expression for R*:\n"]], ["block_4", ["It is difficult to predict with certainty the rate at which critical nuclei form, but the two most\nimportant factors can be identified. The probability of a critical nucleus should be proportional to a\nBoltzmann factor, exp(\u2014AG*/kT), which would determine the \u201cequilibrium\u201d concentration of\nsuch nuclei. Secondly, the rate of formation will be proportional to the rate of arrival of new\nmolecules at the droplet surface, or, in the case of polymers in the bulk, the rate of segmental\nrearrangements at the surface in order to fit into the lattice. In either case we know from the\ndiscussion in Section 12.4, and Equation 12.412 in particular, that the temperature dependence of\npolymer transport follows the Vogel\u2014Fulcher or Williams\u2014Landel\u2014Ferry form:\n"]], ["block_5", ["This, in turn, can be substituted into Equation 13.5.1 to \ufb01nd the free-energy barrier associated with\nachieving the critical nucleus:\n"]], ["block_6", ["where A, B, and T0 are parameters that can be related to free volume. (However, in practice the\nvalues of A, B, and T0 will be different for crystallization rate than for, say, the macroscopic\nviscosity.) Combining Equation\n13.5.4 and Equation 13.5.5, we can extract the temperature\ndependence of the nucleation rate:\n"]], ["block_7", ["We can also use Equation 13.4.5 to replace AGV with a term involving the enthalpy of fusion and\nthe undercooling:\n"]], ["block_8", ["533\nCrystalline Polymers\n"]], ["block_9", ["27\nR* \n13.5.2\nAGV\n(\n)\n"]], ["block_10", ["6\n3\n16\n3\nT00 \nAG,,=1_v 72\n____7T r2\n(\nrn)2\n3\nAGV\n3\nAHV (AT)\n(13.5.4)\n"]], ["block_11", ["Polymer or segment mobility A exp(\u2014\n)\n(13.5.5)\nT To\n"]], ["block_12", ["4\nAG* = 477R*2y + %R*3AGV = \n(13.5.3)\n"]], ["block_13", ["3\nAG*\nln(rate) oc 1n\nexp\n_T\u2014\u2014T0\nexp\n\u2014\nk7\"\n+ constant\n"]], ["block_14", [{"image_2": "548_2.png", "coords": [51, 485, 292, 547], "fig_type": "molecule"}]], ["block_15", ["B\n= \n\u2014\n+ constant\nT \u2014 To\nrmrf\n(13.5.6)\n"]], ["block_16", [{"image_3": "548_3.png", "coords": [147, 489, 294, 524], "fig_type": "molecule"}]]], "page_549": [["block_0", [{"image_0": "549_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "549_1.png", "coords": [30, 466, 143, 507], "fig_type": "molecule"}]], ["block_2", ["The next step is to repeat this process, but differentiating Equation 13.5.7 (with p set equal to 19*)\nwith respect to E and setting the result equal to zero. This process locates that value of E, E*, which\nminimizes the nucleation barrier AG*( p*, E). In other words, AG( p, E) is a surface with respect to\np and E; we have found the curve that represents the maximum with respect to p (p p*), and now\nwe seek the minimum along that curve in terms of E (E E*). The resulting point is called a saddle\npoint; just as a hiker seeks a low altitude pass through a mountain range, the crystallization process\nwill seek the easiest route across the nucleation barrier. When this is done (after some algebra,\nProblem 8), the resulting value of the critical stern length E* turns out to be simply\n"]], ["block_3", ["We now take a closer look at the nucleation of a polymer crystal. Experimental evidence, much\nof it indirect, suggests that the critical nucleus size is on the order of 10 nm. Furthermore, even in\nthe early stages the nuclei can be faceted to re\ufb02ect the underlying unit cell. The important new\nissue is: what sets the thickness of the crystal nucleus, E? We already know that lamellae tend to\ngrow with a (nearly) single value of E that depends primary on AT, so presumably the critical\nnucleus has to set the stage for this choice of E. A straightforward extension of the critical nucleus\nanalysis can give some insight into how this might come about. Assume a cylindrical nucleus just\nas in Figure 13.11, with height E and radius r. Assume the nucleus contains p stems (which might\ncome from several chains, but not necessarily p different chains), and that each stem occupies a\ncross-sectional area d2. The volume of this nucleus can be written as arrzE, or as pdzE, so\nr :di/p/rr. We can now write the analogous expression to Equation 13.5.1, Equation 13.4.2,\nand Equation 13.4.3 for the free energy of the nucleus, but now in terms of p and E:\n"]], ["block_4", ["Note that we are making the simplifying assumption that both top and side faces have the same\nsurface energy. The critical nucleus size in terms of the number of stems can be found as before, by\ndifferentiating Equation 13.5.7 with respect to p and setting the result equal to zero. The answer is\n(see Problem 8)\n"]], ["block_5", ["Kinetics of Nucleation and Growth\n539\n"]], ["block_6", ["This relation has exactly the same form as Equation 13.5.4; all that has changed is the numerical\nprefactor. In particular, the nucleation barrier height still depends on the inverse square of the\nundercooling, and so nucleation should be more facile at lower temperatures. What is new from\nthis analysis is the existence of a preferred lamellar thickness for nucleation, E*. It depends linearly\non the inverse undercooling, so we expect thicker nuclei (and crystals) as TC is lowered; The\ndependence of AG from Equation 13.5.7 on p and E is illustrated in Figure 13.18, for a!2 20 A2 per\nchain, and E* 100 A. For each value of E, AG shows a maximum in p, and the lowest value of\nthis maximum occurs for E* 100 A and p* =400. These curves can be used to estimate the\nrelative rates of nucleation of different sized nuclei, as illustrated in Problem 9.\n"]], ["block_7", ["Substituting Equation 13.5.9 and Equation 13.5.8 back into Equation 13.5.7 gives the critical\nbarrier height:\n"]], ["block_8", ["Although the processes of lamellar growth are still far from fully understood on the molecular\nscale, a reasonable understanding of the principal factors can be extracted by extension of the\n"]], ["block_9", ["13.5.2\nCrystal Growth\n"]], ["block_10", [{"image_2": "549_2.png", "coords": [42, 272, 140, 323], "fig_type": "molecule"}]], ["block_11", ["_\nWEZ')!2\n7\nd2(EAGv + 2302\n19*\n(13.5.8)\n"]], ["block_12", ["AG( p, 2) pdzEAGV + 2pd2\u2018y + 2dE'y. /\u20147Tp\n(13.5.7)\n"]], ["block_13", ["4?\nE* \nAGV\n"]], ["block_14", ["AG(p*, E*) =\n(13.5.10)\n"]], ["block_15", ["877' 73\nAG?)\n"]], ["block_16", ["(13.5.9)\n"]]], "page_550": [["block_0", [{"image_0": "550_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "550_1.png", "coords": [13, 536, 369, 663], "fig_type": "figure"}]], ["block_2", [{"image_2": "550_2.png", "coords": [27, 548, 266, 641], "fig_type": "figure"}]], ["block_3", ["concepts developed in the previous section. It is well established that under conditions of\nisothermal crystallization the growth velocity v of a lamellar face is a constant, and that the\nlamellar thickness also remains constant. The temperature dependence of v can be quite interest-\ning, however, as we shall see.\nWe begin by assuming we have a perfect crystal face of height 6 and width W, as shown in\nFigure 13.19. To start a new layer of chains (whether folded immediately or not), a single stern\nmust attach to the crystal. There is a barrier to this process, because the new stern has increased the\n"]], ["block_4", ["AG]?\n\"'-'-\n"]], ["block_5", ["surface area of the crystal. Indeed, this process is termed secondary nucleation, because it\nnucleates the growth of a single new layer, in contrast to the primary nucleation process considered\n"]], ["block_6", ["Figure 13.18\nThe free energy of forming a nucleus, divided by the surface energy, as a function of the\nnumber of stems, p, for different values of the lamella thickness, according to Equation 13.5.7.\n"]], ["block_7", ["a perfectly \ufb02at crystal face.\n"]], ["block_8", ["540\nCrystalline Polymers\n"]], ["block_9", ["Figure 13.19\nSchematic of the secondary nucleation process, whereby a single complete new stem adds to\n"]], ["block_10", ["200\n400\n600\n800\n1 000\nO\n1\n"]], ["block_11", [{"image_3": "550_3.png", "coords": [62, 198, 233, 330], "fig_type": "figure"}]], ["block_12", ["o\n\u2018\n\\\nI\n80A\n"]], ["block_13", [".\nE\" values\n.\n\\\n"]], ["block_14", [{"image_4": "550_4.png", "coords": [82, 54, 361, 360], "fig_type": "figure"}]], ["block_15", [{"image_5": "550_5.png", "coords": [122, 209, 213, 321], "fig_type": "figure"}]], ["block_16", ["100A\n~\n\\\n----- 120A\n,\n\\_\n-----140A\n~\n*\n"]], ["block_17", ["---60A\n.\n"]], ["block_18", ["---.-\n"]]], "page_551": [["block_0", [{"image_0": "551_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "551_1.png", "coords": [31, 212, 201, 264], "fig_type": "molecule"}]], ["block_2", ["and the temperature dependence of v, once corrected for the transport term, will be linear in l/TAT.\nIn fact this is often observed; an example from the solution crystallization of polyethylene is shown\nin Figure 13.20a (where the transport term is not so important).\nA strong indication that the real process is much more complicated than the above description\ncan be seen in Figure 13.20b. Here, logarithmic growth velocities, corrected for the transport term,\nare plotted as a function of 1/TAT, for four different polyethylenes. Each curve exhibits two\ndifferent linear portions, separated by rather sharp changes; for each curve the higher-temperature\nregime has the larger slope. Several polymers display this general behavior, whereas in others the\n"]], ["block_3", ["However, under normal conditions we expect AGg to be negative, or crystallization is not really\ngoing to proceed at all, and so there is no barrier and therefore g does not depend on the\nundercooling, AT.\nIn the above scenario, it is not unreasonable to expect that the rate of growth, g, will exceed the\nrate of secondary nucleation, r,, so much so that each new layer will fill in completely before the\nnext layer starts. If true, the overall growth velocity (units of length time\u2014l) will simply be given\nby the product of the layer thickness, d, the width, W and r5 (units of length\u20181 time\u20141):\n"]], ["block_4", ["The rate of growth of a single face, g, should follow an expression analogous to Equation 13.5.12:\n"]], ["block_5", ["where AG;k is the barrier to secondary nucleation. Once again the relevant process is governed by\nthe competition between the gain in bulk free energy, proportional to AGV, and the surface energy\npenalty, involving yf and yg. Note that we cannot differentiate Equation 13.5.11 to find a \u201ccritical\u201d\nnucleus, because we have already assumed that it is a single stem of length 1?. A much more\ndetailed treatment, due to Hoffmann and Lauritzen [2], gives AG;k \u2014d'yfyg/AGV, and therefore\nln r5 will be proportional to a term in 1/TAT (see Equation 13.5.6).\nOnce the secondary nucleus is in place, the rest of the layer could fill in by adding adjacent\nstems. Each new stem increases the crystal volume by the same amount, d326, but also increases the\nfold surface area by 2d2; it does not increase the exposed growth-surface area. Thus for the growth\nprocess we can write\n"]], ["block_6", ["The overall rate of secondary nucleation per unit width of the growth surface, which we will call rs,\nshould be proportional to the product of a dynamics term and the apprOpriate Boltzmann factor,\njust as in Equation 13.5.6:\n"]], ["block_7", [{"image_2": "551_2.png", "coords": [34, 421, 202, 456], "fig_type": "molecule"}]], ["block_8", ["in the previous section. For simplicity we assume that the stern has a square cross-section with side\nd. The barrier to the addition of a single stem, AGS, contains three terms just as in Equation 13.5.7.\nThere is a favorable contribution from the increase of the crystal bulk, given by EdzAGv (recall\nthat AGV is negative). There are two unfavorable terms for the added surface, one from the top and\nbottom of the lamella, and the other from the new faces of the stem. The former is given by 2d2'yf\nand the latter by 2d\u20acyg, where we distinguish the two surface energies with subscripts f for \u201cfold\nsurface\u201d and g for \u201cgrowth surface,\u201d respectively. Thus the free-energy change associated with\nsecondary nucleation can be written as\n"]], ["block_9", ["Kinetics of Nucleation and Growth\n541\n"]], ["block_10", ["v =dWr,\n(13.5.15)\n"]], ["block_11", ["B\nA\ng oc exp (\u2014 T _ To) exp <\u2014 %)\n(13.5.14)\n"]], ["block_12", ["AGg :dQEAGV + My,\n(13.5.13)\n"]], ["block_13", ["B\nAG*\nrs OC EXP (\u2014 T _ To) exp (-\u2014 HS)\n(13.5.12)\n"]], ["block_14", ["AG, dQEAGV + My, + 2cm,\n(13.5.11)\n"]]], "page_552": [["block_0", [{"image_0": "552_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["542\n"]], ["block_2", ["Figure 13.20\nCrystallization rates for polyethylenes. (a) Single crystals grown from solution. (Reprinted\nfrom Toda, A. and Kiho, H., J. Polym. Sci, Polym. Phys. Ed., 27, 53, 1989. With permission.) (b) Spherulites\ngrown from the melt. (Reprinted from Lambert, W.S. and Phillips, P.J., Macromolecules, 27, 3537, 1994.\nWith permission.)\n"]], ["block_3", [",...0.5-\nIf\u201d\nIi.\nmv\n\u2014\nEl:\n"]], ["block_4", ["\u2018I:\n_\n"]], ["block_5", ["5\u201d:0:\u2014\nO?\nI\n_O\n_.\n"]], ["block_6", [">(\u20183.>\n+N\u2014o5\u2014\n>\nU)\n_o\n"]], ["block_7", ["(a)\n10WTAT\n"]], ["block_8", ["\u20141_\u2014\n"]], ["block_9", ["(b)\n"]], ["block_10", [{"image_1": "552_1.png", "coords": [54, 39, 286, 313], "fig_type": "figure"}]], ["block_11", ["E]\n1_\nDD\n\u2014_\n:\nDE\n3\n"]], ["block_12", ["\u20141.5\u2014\n"]], ["block_13", ["\u20142.5IIII1II[IIIIIrI[IIII\n"]], ["block_14", ["1.5\u2014\n"]], ["block_15", ["67.0\n71.0\n75.0\n79.0\n_I\nI\nI\nl\nI\nl\nI\n"]], ["block_16", [{"image_2": "552_2.png", "coords": [70, 331, 320, 589], "fig_type": "figure"}]], ["block_17", ["Ill\nllllllll\nl\n"]], ["block_18", ["9\n10\n11\n12\n"]], ["block_19", ["|\nI\nI\nI\nI\nl\n|\n"]], ["block_20", ["14.0\n180\n1/TAT(x10%\n"]], ["block_21", ["T00)\n"]], ["block_22", ["Crystalline Polymers\n"]]], "page_553": [["block_0", [{"image_0": "553_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["If this assumption is correct, the dependence of v on temperature will still come largely from rs, as\nnoted above. Because of the square root, the argument of the Boltzmann factor will now be\nmultiplied by 1/2, and thus the slope will decrease by a factor of 2, as in Figure 13.20b. In the\ncontext of the Hoffmann and Lauritzen theory, this transition in mechanism corresponds to passage\nfrom so-called Regime I crystallization, where rs << g, to Regime H, where r8 2:. g. With still deeper\nundercooling, the system can undergo another transition into Regime HI, where rS >> g. In this case,\nv rs again because there is essentially no lateral growth of a crystal layer; as a consequence the\nslope of In v versus l/TAT will increase by a factor of 2 from that in Regime II.\nWe have deliberately avoided any more detailed examination of the secondary nucleation and\nlayer growth processes, in part because a full molecular level description would be very compli-\ncated, and in part because these issues are far from fully resolved. Details we have not considered\ninclude the following: Does a secondary nucleus have a stem of length 6, where E is the thickness of\nthe primary nucleus, or is it different? Does growth proceed by one chain at a time folding\nregularly, like a fire hose (as the Hoffmann\u2014Lauritzen theory proposes), or do new stems from\nother chains participate? Does a new layer immediately grow with thickness 6, or does it armeal to\nfull thickness after first attaching to the surface? Does a new stem begin by one repeat unit sticking\nto the face of a unit cell, or do longer helical sections form first? Is the melt in the immediate\nvicinity of the growth surface completely disordered, or is there some intermediate level of order\nthat precedes attachment of new stems? Is the lamella growth surface actually \ufb02at, or is it rough,\nthereby removing the need for secondary nucleation?\nThese considerations aside, there is one other general feature of the growth process outlined\nabove that should be brought out. In Regime I, the thermodynamic drive to grow is relatively\nsmall, and the system can be thought of as being close to a local equilibrium between stems\nattaching and detaching. In other words, a new stem might be formed and then melt off the surface\nseveral times before actually being locked in place. Under such quasi-equilibrium growth condi-\ntions, very smooth and regular crystals can be grown. This is consistent with the familiar\nexperience of growing small-molecule crystals, where modest undercoolings and long times are\nnecessary to obtain large single crystals. In contrast, for deeper undercoolings in Regime III, stems\ncan be envisioned as sticking virtually irreversibly on the \ufb01rst attempt. This kind of a disorganized\nprocess leads to rapid growth, but more defect-laden, irregular structures.\nOne final issue to consider: what is the role of molecular weight in all of this? We have so far\ncompletely ignored the dependence of nucleation or growth on this most important polymer\nvariable. Two \ufb01gures serve to illustrate the main points. Figure 13.2] shows the temperature\ndependence of crystal growth for various molecular weights of poly(tetramethyl-p-phenylene\nsiloxane). There is a strong peak in growth velocity near 65\u00b0C, with the rate tending to zero\n"]], ["block_2", ["lower-temperature regime has the larger slope. Still other polymers show three or even four\nregimes, but in most cases the slopes differ by factors of about 2 (or 1/2). These abrupt changes\nsuggest that the growth mechanism is changing in some distinct way. A good possibility to\nconsider is that the assumption g > rS may not always apply. In particular, the Boltzmann factor\nfor r, depends on l/TAT, whereas for g it depends only on 1/2\". A small change in T will have no\nappreciable effect on g, but it can have a profound effect on rs. Consequently as the undercooling is\nincreased, perhaps rS and g become competitive, or even rS >> g.\nThese possibilities are also considered in the theory of Hoffmann and Lauritzen. The main\nconclusion of interest to us can be anticipated in a rather straightforward way. Suppose that the\nundercooling has increased to the point where rS and g are similar. In that case a new layer will fill\nin by both random attachment and adjacent attachment of stems. The growth velocity v will depend\non both rs and g. As r5 has units of length\u20181 time\u201c, and g has units of length time\u20141, dimensional\nanalysis suggests that\n"]], ["block_3", ["Kinetics of Nucleation and Growth\n543\n"]], ["block_4", ["v oc ark/rsg\n(13.5.16)\n"]]], "page_554": [["block_0", [{"image_0": "554_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "554_1.png", "coords": [24, 38, 350, 386], "fig_type": "figure"}]], ["block_2", ["Figure 13.21\nGrowth rates of poly(tetramethyl-p-phenylene siloxane) crystals as a function of temperature,\nfor the indicated molecular weights. (Reprinted from Magill, J.H., J. Appl. Phys, 35, 3249, 1964. With\npermission.)\n"]], ["block_3", ["near 120\u00b0C and 0\u00b0C. This feature is exactly what we would anticipate based on the discussion\nabove, namely that at relatively small undercoolings the rate increases with lower temperatures,\nbut eventually the transport term takes over and the rate goes to zero. For comparison purposes, the\nglass\u2014transition temperature for this polymer is about \u201420\u00b0C, and so the rate becomes negligible a\nfew degrees above Tg. The important new information in this \ufb01gure is that the peak position is\nindependent of molecular weight, but the peak growth velocity decreases significantly with\nincreasing molecular weight. The peak position is controlled by a balance between the thermo\u2014\ndynamics associated with adding stems (whether by secondary nucleation or by layer growth), and\nthe dynamics of molecular rearrangements. Neither AGV nor the various surface energies depend\nappreciably on molecular weight, and neither do the Vogel\u2014Fulcher or WLF temperature depend\u2014\nence of chain dynamics (at least for reasonably long chains, see Section 12.4); thus the position of\nthe peak is also insensitive to chain length.\nThe molecular weight dependence of the growth velocity can be better seen when plotted\ndirectly against inverse molecular weight, as shown in Figure 13.22. At high molecular weights,\nwhen the chains are well entangled (see Chapter 11) the rate is lowest, but independent of\nmolecular weight. At lower molecular weights the rate increases as the chains become shorter.\nThe molecular weight independence for long chains argues for a rate~detennining step that\ninvolves rearrangements of a portion of the chain, perhaps a few times the length of a stern, rather\nthan diffusion of the whole polymer. Lower molecular weight chains can accommodate the\nnecessary conformational rearrangements more rapidly, as entire molecules.\n"]], ["block_4", ["544\nCrystalline Polymers\n"]], ["block_5", ["E\n10,000\n.3:\n60 \n_\n>\n15,800\n.20.\n6:\u201c\n50 \n25,000\n-\n"]], ["block_6", ["80 \n-\nl\nM,\n70 \n\u2014\nE\u201c\n8,700\n"]], ["block_7", ["100 \n-\n"]], ["block_8", ["1 10\nI\n|\nI\nl\nl\n|\nl\nl\nI\n"]], ["block_9", ["27,000\n40 \n\u2018\n37,500\n'\n"]], ["block_10", ["90 \n_\n"]], ["block_11", ["30 \n\\\n\u2014\n"]], ["block_12", ["20 \n.\n\\\\u2018fl\n143,000\n_\n"]], ["block_13", ["10 \n'\n+9\n\u2014\nI\n"]], ["block_14", ["I'T\nI\nI\nI\n:\u2018E.\nI\n, ..-\n\u201440\n\u201420\n0\n20\n40\n60\n80\n100\n120\n140\n160\n"]], ["block_15", [{"image_2": "554_2.png", "coords": [130, 66, 328, 344], "fig_type": "figure"}]], ["block_16", ["Temperature (\u00b0C)\n"]], ["block_17", [".\n"]], ["block_18", ["\u2018\n,\n56,000\n"]]], "page_555": [["block_0", [{"image_0": "555_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Figure 13.22\nSame data as in Figure 13.21, now plotted versus inverse molecular weight at the indicated\ntemperatures.\n"]], ["block_2", ["Suppose a bulk-crystallized polymer sample is observed in a polarizing optical microscope, with\nthe sample placed between two polarizers oriented at right angles to each other. In the absence of\nany sample, no light would be transmitted owing to the 90\u00b0 angle between the vectors describing\nthe light transmitted by the two polarizers (see Section 8.1 for a discussion of polarized light). With\na crystalline sample of polymer in place, however, an image such as that shown in Figure 13.23 is\ngenerally observed. The field of view becomes at least partially filled with domains called\nspherulites, which grow in time, impinge upon one another, and eventually fill space. They\ngenerally exhibit the following features:\n"]], ["block_3", ["At this point we have a good picture of the organization of polymer molecules at the unit cell level\n(~1 nm) and within a lamella (~10 nm). To complete the picture we need to consider how the\nlamellae and the intervening regions of amorphous material arrange themselves to fill up the bulk\nof the material. By far the most commonly observed morphology is that of the spherulite, to be\nconsidered \ufb01rst, but other structures such as hedrites, dendrites, and shish kebabs are also found\nunder certain crystallization conditions.\n"]], ["block_4", ["13.6.1\nSpherulites\n"]], ["block_5", ["13.6\nMorphology of Semicrystalline Polymers\n"]], ["block_6", ["Morphology of Semicrystalline Polymers\n545\n"]], ["block_7", [{"image_1": "555_1.png", "coords": [35, 64, 354, 328], "fig_type": "figure"}]], ["block_8", ["1.\nThey possess spherical symmetry around a single center of nucleation. This symmetry projects\n"]], ["block_9", ["_\n60\u00b0\nc\no\n'E\n80\nE\n(5\no\n10\n40\n"]], ["block_10", ["100\n"]], ["block_11", ["2..\n20\u00b0 lllllllll\n|II|||l||\nOJ\n1.0\n100\n"]], ["block_12", ["a perfectly circular cross-section if the development of the spherulite is not stopped by contact\nwith another expanding spherulite. The spherical structure indicates a single growth rate in\n"]], ["block_13", ["100\u00b0\n"]], ["block_14", ["1 03/MW\n"]]], "page_556": [["block_0", [{"image_0": "556_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "556_1.png", "coords": [0, 31, 303, 285], "fig_type": "figure"}]], ["block_2", ["On the basis of a variety of experimental observations, including an analysis of the ubiquitous\nMaltese cross, a number of aspects of the structure of spherulites have been elucidated. The\nspherulites are aggregates of lamellar crystals radiating from a single nucleation site. The latter\ncan be either a spontaneously formed single crystal or a foreign body. The spherical symmetry is not\npresent at the outset, but develops with time. Fibrous or lathlike lamellar crystals grow away from\nthe nucleus, and begin branching and fanning out. As the lamellae spread out radially and three\ndimensionally, branching of the crystallites continues to sustain the spherical morphology. Figure\n13.24 represents schematically the leading edge of some of these fibrils, one of which has just split.\nThe molecular alignment within these radiating fibers is tangential, that is, perpendicular to\nthe radius of the individual spherulite. The individual lamellae are similar in organization to single\ncrystals: they consist of ribbons on the order of 10\u2014100 nm in thickness, built up from successive\n"]], ["block_3", ["three dimensions, which we will need to reconcile with the one\u2014 or two\u2014dimensional growth\nmode of individual lamellae.\n2.\nEach spherulite is revealed by the characteristic Maltese cross optical pattern under crossed\npolarizers, although the Maltese cross is truncated in the event of impinging spherulites.\n3.\nSuperimposed on the Maltese cross may be such additional optical features as banding,\nillustrated in Figure 13.26.\n4.\nA system of mutually impinging spherulites ultimately develop into an array of irregular\npolyhedra, the dimensions of which can be as large as a millimeter or more. The size of the\ndomains will obviously increase as the number of nuclei decreases, and so information about\nnucleation density may be inferred even after crystallization is complete.\n5.\nA larger number of smaller spherulites are produced at larger undercoolings, as the barrier to\nnucleation is reduced. Various details of the Maltese cross pattern, such as the presence or\nabsence of banding, may also depend on the temperature of crystallization.\n6.\nSpherulites have been commonly observed in organic and inorganic systems of synthetic,\nbiological, and geological origin, including moon rocks, and are therefore not unique to\npolymers.\n"]], ["block_4", ["Figure 13.23\nSpherulites of poly(L-lactide) growing from the melt. Courtesy of R. Taribagil.\n"]], ["block_5", ["546\nCrystalline Polymers\n"]]], "page_557": [["block_0", [{"image_0": "557_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "557_1.png", "coords": [27, 41, 232, 244], "fig_type": "figure"}]], ["block_2", ["\u201dl l:l \\\n\u2018. ___,_\n(4/? __ _\n'4? \\\n.\n\\\n"]], ["block_3", ["1.\nThe ordered polymer chains, that is the stems, are consistently oriented perpendicular to the\nradius of the spherulite.\n2.\nThe index of refraction of all polymers differs for light polarized parallel to the chain axis\nversus normal to it. Recalling the discussion in Section 8.1, the refractive index of a material\nre\ufb02ects the polarizability of the constituent molecules. Individual polymer molecules have a\npolarizability anisotropy, which leads to anisotropy in refractive index when the molecules are\naligned. Substances showing this anisotropy of refractive index are said to be birefringent.\n3.\nItems (1) and (2) indicate that the refractive index in the tangential direction of the spherulite\ndiffers from that along the radius. It actually does not matter in this context which refractive\nindex is greater; for polyethylene and poly(ethylene oxide), it is greater along the chain\nbackbone, but for polystyrene and poly(vinyl chloride) it is greater normal to it.\n4.\nThe electric vector of the polarized light emerging from the first polarizer (let us say the beam\nis traveling along lab direction 2 and is polarized along )2) may be resolved into components\nalong the radial and tangential directions of the spherulite. These two components propagate\nat different speeds (recall Equation 8.1.6), and thus will not recombine to recover perfectly\ny-polarized light. The resulting x\u2014component of light is transmitted by the second polarizer,\nleading to a bright image.\n5.\nAs we proceed around the spherulite, at 90\u00b0 intervals the polarization axis of the light will\ncoincide with either the radial or tangential direction in the crystal. At that point, the electric\nvector has no component along the orthogonal axis, and so there is no component to sense the\ndifferent refractive index. Consequently the light emerges with polarization preserved, and is\nextinguished by the second polarizer. Thus there are four dark sectors in the image, producing\nthe Maltese cross.\n"]], ["block_4", ["layers of folded chains. Growth is accomplished primarily by the addition of successive layers of\nchains to the ends of the radiating laths. This feature is also indicated schematically in Figure\n13.24. Portions of an individual long polymer chain can be incorporated into different lamellae,\nand thus link them in a three-dimensional network. These interlamellar links are not possible in\nspherulites of low molecular-weight compounds, which show much poorer mechanical strength as\na consequence.\nThe molecular chain folding is the origin of the Maltese cross. The Maltese cross pattern arises\nfrom a spherical array of birefringent particles through the following considerations (see also\nFigure 13.25):\n"]], ["block_5", ["Morphology of Semicrystalline Polymers\n547\n"]], ["block_6", ["Figure 13.24\nSchematic illustration of the leading edge of a lathlike crystal within a spherulite.\n"]]], "page_558": [["block_0", [{"image_0": "558_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In many cases, the microscope image reveals another feature, namely banded spherulites, as\nillustrated in Figure 13.26. In addition to the Maltese cross, there are concentric dark rings at\nregular intervals moving out from the center. This is due to twisting of the individual lamellar\nribbons along the radial direction; from the spacing of the bands, the period of the twist can be\ncalculated and is found to depend on crystallization conditions. The fact that the dark rings are\nmore or less continuous around a circle implies that the material within the ring is optically\nisotropic, rather than oriented with its optical axes parallel to the polarizers. As an individual\nlamella twists, the fold plane containing the chain stems rotates about the radial direction. The\nrefractive index is also different parallel and perpendicular to the stem within the fold plane, and so\nthe value of the mean refractive index along the tangential direction varies continuously between\nthese limits with the twist angle. At certain twist angles, this projected tangential value of\nrefractive index can be identical to the radial value, leading to optical isotropy and light extinction.\nThe band spacing re\ufb02ects the distance over which a lamella completes a helical twist; the origin of\nthe twist is thought to lie in the particular chain conformations of the fold surface.\n"]], ["block_2", ["5.\nYou should convince yourself that if the polarizers are held fixed and the sample rotated\nbetween them, the Maltese cross remains \ufb01xed because of the symmetry of the spherulite.\n"]], ["block_3", ["Figure 13.25\ny-Polarized light is incident on a spherulite. The small rectangles and arrows illustrate the\nlocal orientation of the refractive indices of the crystal. Light is transmitted through an x\u2014polarizer only when\nthe local orientation is not parallel to x or y.\n"]], ["block_4", ["Spherulitic growth is the natural consequence of a crystallization process that proceeds steadily\nfrom a single nucleus, with a constant growth rate (although that is not strictly required), and is\nallowed to proceed with equal probability in three dimensions. A more subtle consideration is that\nthe growth rate has to be sufficiently slow so that the relevant lamellar facets extend outward with\nno significant change in structure. Such a growth process is a natural consequence of a situation\nwhere secondary nucleation is the rate-limiting step, as for example in the theory of Hoffmann and\nLauritzen described in Section 13.5.2. In this process, it takes a while for a growth face to add the\nfirst stern of the next layer, but once it does, the layer fills in completely and relatively rapidly.\n"]], ["block_5", ["548\nCrystalline Polymers\n"]], ["block_6", ["13.6.2\nNonspherulitic Morphologies\n"]], ["block_7", [{"image_1": "558_1.png", "coords": [71, 46, 299, 262], "fig_type": "figure"}]], ["block_8", ["X\n"]], ["block_9", [{"image_2": "558_2.png", "coords": [94, 98, 234, 241], "fig_type": "figure"}]], ["block_10", [{"image_3": "558_3.png", "coords": [140, 39, 385, 232], "fig_type": "figure"}]], ["block_11", [{"image_4": "558_4.png", "coords": [220, 45, 357, 187], "fig_type": "figure"}]]], "page_559": [["block_0", [{"image_0": "559_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["In this manner the structure of the lamellae is preserved. In another limit, addition of a new stem\nto a smooth facet could be more rapid than filling out the new layer. This is reminiscent of a\ngeneral growth process known as diffusion limited aggregation, in which new particles stick to\nthe first part of a growing cluster that they encounter. Under these conditions, a crystal or\naggregate would grow much more haphazardly, leading to dendritic structures (which are often\nakin to the form of a fir tree). An example of dendritic growth of polyethylene crystals in\nsolution is shown in Figure 13.27. This growth mode was accessed by the simple expedient of a\ndeeper undercooling; the increased thermodynamic drive to form crystals reduces the barrier to\nsecondary nucleation.\nHedrites, 0r axialites as they are sometimes termed, represent an alternative morphology that is\nsometimes encountered. A hedrite may be defined as a crystallite that looks like a polygon when\nviewed from at least one direction [3]; an example is shown in Figure 13.28a. A cartoon version of\none possible hedrite structure is shown in Figure 13.28b. In this case several lamellae are stacked\non top of one another, but then splay out when growing further from the center. This is somewhat\nanalogous to splaying the pages in a book or sheets in a stack of paper. Qualitatively, one can\nimagine such a structure emerging when the individual lamellae grow at comparable rates in two\ndimensions; this growth mode inhibits structures with spherical symmetry. In contrast, the spher-\nulite results from primarily one\u2014dimensional growth of individual lamellae, that fan out in three\ndimensions over time.\nWe conclude this section on crystal morph010gy by brie\ufb02y describing crystallization under\napplied stress. This is a fascinating topic in its own right, and of great importance in processing of\nsome semicrystalline polymers. We noted this possibility in Chapter 10 when discussing deviations\n"]], ["block_2", ["Figure 13.26\nSpherulites of poly(1\u2014pr0py1ene oxide) observed through crossed polarizers by optical\nmicroscopy. (Reproduced from MaGill, J.H., Treatise on Materials Science and Technology, Vol. 10A,\nSchultz, J.M. (Ed), Academic, New York, 1977. With permission.)\n"]], ["block_3", ["Morphology of Semicrystalline Polymers\n549\n"]]], "page_560": [["block_0", [{"image_0": "560_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "560_1.png", "coords": [13, 20, 256, 319], "fig_type": "figure"}]], ["block_2", ["from the simple model for rubber elasticity under large deformations. Stress-induced crystallinity\nis important in film and fiber technology. For example, when dilute solutions of polymers are\nstirred rapidly, or when fibers are spun from relatively dilute solutions, characteristic structures\ndeve10p, which are described as having a shim\u2014kebab morphology. A beautiful example is shown\nin Figure\n13.2%, and a cartoon of the underlying structure is provided in Figure\n13.2%.\n"]], ["block_3", ["Figure 13.28\n(a) A polyethylene hedrite. (b) Cartoon of a hedrite viewed end\u2014on. (Reproduced from\nBassett, D.C., Keller, A., and Mitsuhashi, S., J. Polym. Sci, 1A, 763, 1963. With permission.)\n"]], ["block_4", ["Figure 13.27\nPolyethylene dendrite grown from dilute toluene solution and observed by interference\n(optical) microscopy; the long dimension is approximately 100 um. (Reproduced from Wunderlich, B. and\nSullivan, P., J. Polym. Sal, 61, 195, 1962. With permission.)\n"]], ["block_5", ["550\nCrystalline Polymers\n"]], ["block_6", ["(b)\n"]], ["block_7", ["\u00a7\u2014\u2014\u2019C/f:\u2014~d\u2014j\n"]]], "page_561": [["block_0", [{"image_0": "561_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "561_1.png", "coords": [13, 182, 290, 368], "fig_type": "figure"}]], ["block_2", ["These consist of chunks of folded chain crystals (kebabs) strung out along a fibrous central column\n(shish). In both portions the polymer chain axes are parallel to the overall axis of the structure.\nThe essence of this process is to extend the individual chains to a substantial degree, before\ncrystallization. In this way the extended chain shish provides nucleation sites for the chain-folded\nkebabs. Relatively dilute solutions are favored, because in a highly entangled state it is very difficult\nto extend individual chains before crystallization. Similarly, relatively high molecular weight chains\nare favored because it is difficult to apply enough stress in solution to extend a short chain.\nExtremely strong fibers can be fabricated from such shish kebabs, as the high degree of crystallinity\ncombined with the uniform orientation of the chain axis imparts remarkable tensile strength. An\nexample of such a material is gel-spun polyethylene, which is prepared in two stages. In the first\nstage, a hot solution is extruded to partly align the chains and cooled into a gel (crystallinity provides\nthe cross-link sites). In the second stage, the material is drawn into fibers while the remaining solvent\nis removed. The resulting fibers have far superior mechanical properties to standard fibers, and are\nused in demanding specialty applications such as bullet-resistant garments and in racing yachts.\n"]], ["block_3", ["Figure 13.29\nTransmission electron micrograph of polyethylene shish kebabs crystallized from xylene\nsolution during flow (a), and (b) schematic of the underlying chain structure. (Reproduced from Pennings,\nA.J., van der Mark, J.M.A.A., and Kiel, A.M., Kolloia\u2019 Z.Z. Polym, 237, 336, 1970. With permission.)\n"]], ["block_4", ["In the previous sections we have discussed the thermodynamic factors that influence crystalliza-\ntion, and we have considered the role of kinetics in the growth of individual crystals or lamellae.\nIn the preceding section we have examined the diverse morphologies that can result from polymer\n"]], ["block_5", ["13.7\nKinetics of Bulk Crystallization\n"]], ["block_6", ["Kinetics of Bulk Crystallization\n551\n"]], ["block_7", ["(a)\n"]], ["block_8", ["I\u201c.\nr\n'.'\n"]], ["block_9", ["\u2019.\nU.\n"]], ["block_10", ["I'r\u2018I:i\n"]], ["block_11", ["'\n\u2018\n"]], ["block_12", ["v\n"]], ["block_13", ["H.\n.'\\\n.3.\"\n"]], ["block_14", [".g'\ni.\nJ..-\n"]], ["block_15", ["\u2018\n"]], ["block_16", ["\\Yr,\n"]], ["block_17", ["___.\u2018,\n"]], ["block_18", ["C\u2014TT.\n"]], ["block_19", [".\nH,-\n"]], ["block_20", [".\n"]], ["block_21", ["n.\n.\n"]], ["block_22", ["_\u2018|'\n"]], ["block_23", [",\n.\n"]], ["block_24", [",\n\"Jui- \n"]], ["block_25", ["*9.-\n"]], ["block_26", ["l\n"]], ["block_27", ["..\n"]], ["block_28", ["m\u201c.\n"]], ["block_29", [".\n"]]], "page_562": [["block_0", [{"image_0": "562_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "562_1.png", "coords": [29, 508, 203, 607], "fig_type": "figure"}]], ["block_2", ["Figure 13.30\nThe growth of disk-shaped crystals. (3) All crystals have been nucleated simultaneously.\nAll crystals have the same radius 1\u2018: after an elapsed time t. (b) Nucleation is Sporadic. Crystal A has\nhad enough time to reach point x, while B has not, although both originate in the same ring a distance r\nfrom x.\n"]], ["block_3", ["Figure 13.30a represents the top view of an array of these disks after the crystals have been\nallowed to grow for a time t after nucleation. The three disks on the left are separated widely\nenough to still have room for further growth; the three disks on the right have impinged upon one\nanother and can grow no more. We saw in the previous section that this latter situation can be\nobserved microscopically.\nSuppose we de\ufb01ne the rate of radial growth of the crystalline disks as f. Then disks originating\nfrom all nuclei within a distance ft of an arbitrary point, say, point x in Figure 13.30a, will reach\nthat point in an elapsed time r. If the average concentration of nuclei in the plane is N (per unit\narea), then the average number of fronts [\u20147, which converge on x in this time interval is\n"]], ["block_4", ["If a second growth front were to impinge on a point like this, its growth would terminate at x.\nSuppose we imagine point x to be \u201ccharmed\u201d in some way such that any number of growth fronts\n"]], ["block_5", ["(a)\n(b)\n"]], ["block_6", ["1.\nThe crystals are initially assumed to be circular disks. This geometry is consistent with\nprevious thermodynamic derivations. It has the advantage of easy mathematical description.\n2.\nThe disks are assumed to lie in the same plane. Although this picture is implausible for bulk\ncrystallization, it makes sense for crystals grown in ultrathin \ufb01lms, adjacent to surfaces, and in\nstretched samples. A similar mathematical formalism will be deve10ped for spherical growth\nand the disk can be regarded as a cross-section of this.\n3.\nNucleation is assumed to begin simultaneously from centers positioned at random throughout\nthe liquid. This is more descriptive of heterogeneous nucleation by foreign bodies introduced\nat a given moment than of random nucleation. We shall subsequently disPense with the\nrequirement of simultaneity.\n4.\nGrowth in the radial direction is assumed to occur at a constant velocity. There is ample\nexperimental justi\ufb01cation for this in the case of three\u2014dimensional spherical growth.\n"]], ["block_7", ["The following derivation will illustrate how the rates of nucleation and growth combine to give the\nnet rate of crystallization [4,5]. The theory we shall develop at first assumes a speci\ufb01c picture of\nthe crystallization process, but then we can generalize the result. The assumptions of the model and\nsome comments on their applicability follow:\n"]], ["block_8", ["crystallization under various situations. In this section we turn our attention back to kinetics, and in\nparticular we consider the following questions: How long will it take a macroscopic sample to\ncrystallize? If we have information about the crystalline fraction as a function of time, what, if\nanything, can we infer about the crystal growth mechanism?\n"]], ["block_9", ["13.7.1\nAvrami Equation\n"]], ["block_10", ["552\nCrystalline Polymers\n"]], ["block_11", [{"image_2": "562_2.png", "coords": [44, 503, 351, 622], "fig_type": "figure"}]], ["block_12", ["T? = 77(ft)2N\n(13.7.1)\n"]], ["block_13", [{"image_3": "562_3.png", "coords": [64, 510, 171, 590], "fig_type": "figure"}]], ["block_14", [{"image_4": "562_4.png", "coords": [170, 514, 338, 618], "fig_type": "figure"}]]], "page_563": [["block_0", [{"image_0": "563_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["can pass through it without interference. If we were to monitor the number of (noninterfering)\nfronts that cross x in a series of observations, we would expect a distribution of values because of\nthe random placement of the nuclei. Furthermore, the distribution of F values is expected to pass\nthrough a maximum. Fronts arising from nuclei very close to x can easily cross x in the allotted\ntime, but the area of melt under consideration in this case is small, so the number of fronts is small.\nAs the area around the charmed point x is enlarged, a larger number of nuclei will be encompassed\nso the number of fronts crossing x will increase. This increase \u2018is offset by the fact that fronts\noriginating from more distant nuclei will require more time to reach x. Therefore the number of\nfronts that cross x (remember that these are free from interference by hypothesis) will increase,\npass through a maximum, and decrease as we allow them to originate from all parts of the sample.\nThis distribution of values of F is our next interest.\nWe propose to describe the distribution of the number of fronts crossing x by the Poisson\ndistribution function, discussed in the context of living polymerization in Chapter 4. This prob-\nability distribution function describes the random partitioning of a set of objects into a fixed\nnumber of boxes. In this case, the probability P(F) describes the likelihood of a speci\ufb01c number of\nfronts, F, arriving per unit time in terms of F and the average number F, as follows (see Equation\n4.2.19):\n"]], ["block_2", ["Next we apply this distribution to the case where F 0, that is, to the case where no fronts have\ncrossed point x. There are several aspects to note about this situation:\n"]], ["block_3", ["for F O.\n2.\nThe condition of no fronts crossing x is automatically a condition of noninterference, so the\nspecial magic previously postulated for point x poses no problem.\n3.\nSince point x is nonspeci\ufb01c, Equation 13.7.3 describes the fraction of observations in which no\nfronts cross any arbitrary point, or the fraction of the area in any one experiment that is crossed\nby no fronts.\n4.\nThis last interpretation makes P(0) the same as the fraction of a sample in the amorphous state.\nIt is conventional to focus on the fraction crystallized, (be; therefore the amorphous fraction is\n"]], ["block_4", ["Kinetics of Bulk Crystallization\n553\n"]], ["block_5", ["or\n"]], ["block_6", ["5.\nInverting and taking the logarithm of both sides of Equation 13.7.4, we obtain\n"]], ["block_7", ["Equation 13.7.1 and Equation 13.7.5 both describe the same situation, and can be equated to give\n"]], ["block_8", ["1.\nSince F0 1 and 0! 1, Equation 13.7.2 becomes\n"]], ["block_9", [{"image_1": "563_1.png", "coords": [42, 265, 116, 302], "fig_type": "molecule"}]], ["block_10", ["P(F) \nFl\n(13.7.2)\n"]], ["block_11", ["1\n_\n.2\n111(1 _ 75\u00a2) arr NR\n(13.7.6)\n"]], ["block_12", ["4),, 1 exp(\u20147Tf2Nt2)\n(13.7.7)\n"]], ["block_13", ["1\n_\nln(l __ (be) F\n(13.7.5)\n"]], ["block_14", ["P(0) e\u2014F\n(13.7.3)\n"]], ["block_15", ["1\u2014 e, P(0) e'F\n(13.7.4)\n"]], ["block_16", ["1 (be and\n"]]], "page_564": [["block_0", [{"image_0": "564_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["where in the previous case the so-called Avrami exponent m 2, and the associated rate constant,\nK, is w2. Suppose rather than growing in two dimensions, the crystal fronts grew uniformly in\nthree dimensions. An analysis similar to the one we just conducted would give m = 3 and\nK (4/3)71-Nf3 in Equation 13.7.8. In this case N would be the number of nuclei per unit volume\nat t=0, and the volume swept out per nucleus in time I would be (4/3)7r(ft)3. Similarly, if the\ncrystals tends to grow in one dimension, as in a growing rod or \ufb01bril, we would find m = 1.\nIn terms of spontaneous crystallization, the assumption that N nuclei begin to grow simultan-\neously at r: 0 is unrealistic. It corresponds most closely to the case of heterogeneous nucleation,\nwhere a fixed number of nucleation sites are in place at t: 0, but that by itself does not guarantee\nsimultaneous onset of growth. We can modify the model to allow for random, spontaneous\nnucleation, a description more appropriate for homogenous nucleation, by the following argument.\nWe draw a set of concentric rings in the plane of the disks around point x as shown in Figure\n13.30b. If the radii are r and r + dr for the rings, then the area enclosed between them is 211-rdr. We\npostulate that spontaneous random nucleation occurs with a frequency of N, having units area-1\ntime'l. The rate of formation of nuclei within the ring is therefore N211? dr.\nWe continue to assume that the crystals so nucleated display a constant rate of radial growth 1\u2018.\nThis means that it takes a crystal originating in a ring of radius r around point x a time given by 17?\nto cross x. The crystal labeled A in Figure 13.30b has had just enough growth time to reach x. On\nthe other hand, a crystal nucleated in this ring after t :7? will not have had time to grow to x. The\ncrystal labeled B in Figure 13.30b is an example of the latter case. It is only nucleation events that\noccur up to t\u2014 179, which have time to grow from the ring of radius r and cross point x by their\ngrowth front. The increment in this number of fronts for the ring of radial thickness dr is\n"]], ["block_2", ["Remember the units involved here: for r' they are length time\u20141; for N, length\u20142; and for r,\ntime. Therefore the exponent is dimensionless, as required. The form of Equation 13.7.7 is such\nthat at small times the exponential equals unity and c=0; at long times the exponential\napproaches zero and (150 =1. In between, an S-shaped curve is predicted for the development of\ncrystallinity with time. Experimentally, curves of this shape are indeed observed. However, we\nshall see presently that this shape is also consistent with other mechanisms in addition to the\none considered so far.\nEquation 13.7.7 may be written in the following general form, known as the Avrami equation\n"]], ["block_3", ["The average number of fronts crossing point x at a time of observation I is the sum of\ncontributions from all rings, which are within reach of x in time t. The most distant ring included\nby this criterion is a distance it from x. The average number of fronts, therefore, is given by\nintegrating Equation 13.7.9 for all rings between r 2: 0 and r 2 ft:\n"]], ["block_4", ["554\nCrystalline Polymers\n"]], ["block_5", ["As before, this quantity in relation to the degree of crystallinity is given by Equation 13.7.5, so\nequating the latter to Equation 13.7.11 gives\n"]], ["block_6", ["As far as this integration is concerned,\n1\u2018 and r are constants, so Equation 13.7.10 is readily\nevaluated to give\n"]], ["block_7", ["r\":\nF = 217A? Jr__\n1\n.\nF 377ml?\n(13.7.11)\n"]], ["block_11", [{"image_1": "564_1.png", "coords": [52, 531, 144, 579], "fig_type": "molecule"}]], ["block_12", ["0\n"]]], "page_565": [["block_0", [{"image_0": "565_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "565_1.png", "coords": [10, 327, 368, 499], "fig_type": "figure"}]], ["block_2", [{"image_2": "565_2.png", "coords": [28, 97, 160, 133], "fig_type": "molecule"}]], ["block_3", [{"image_3": "565_3.png", "coords": [28, 367, 307, 476], "fig_type": "figure"}]], ["block_4", ["0.1\n3.89 x 102\n5.33 x 102\n6.24 x 102\n0.2\n5.67 x 102\n6.85 x 102\n7.53 x 102\n0.3\n7.18 x 102\n8.02 x 102\n8.47 x 102\n0.4\n8.59 x 102\n9.03 x 102\n927\u2014102\n0.5\n1\u2014103\n1\u2014103\n1 x 103\n0.6\n1.15 x 103\n1.10 x 103\n1.07 x 103\n0.7\n1.32 x 103\n1.20 x 103\n1.15 x 103\n0.8\n1.52 X 103\n1.32 x 103\n1.23 x 103\n0.9\n1.82 x 103\n1.49 x 103\n1.35 x 10-\"\n"]], ["block_5", ["Equation 13.7.7 and Equation 13.7.13 are analogous, except that the former assumes instantaneous\nnucleation at N sites per unit area while the latter assumes a nucleation rate of N per unit area per\nunit time. It is the presence of this latter rate that requires the power of t to be increased from 2 to 3\nin this case. Again, Equation 13.7.13 is a particular case of the Avrami equation, Equation 13.7.8;\nthe effect of switching from instantaneous nucleation to a constant nucleation rate is to increase the\nvalue of the Avrami exponent, m, by 1. For instantaneous nucleation, m: 1, 2, or 3, and for a\nconstant nucleation rate, m =2, 3, or 4, depending on the dimensionality of the growth process.\nTo acquire some numerical familiarity with the Avrami function, consider the following\nexample.\n"]], ["block_6", ["Three different crystallization systems show m values of 2, 3, or 4. Calculate the value required for\nK in each of these systems so that all will show 650 0.5 after 103 s. Use these m and K values to\ncompare the development of crystallinity with time for these three systems.\n"]], ["block_7", ["01\u2018\n"]], ["block_8", ["(15,, 0.4 and 0.6). For m =4, the range is narrowest (726 s) and the maximum slope is steepest\n(1.4 x 10\u20143).\nA further extension to the Avrami equation concerns the rate-determining step of the crystal-\nlization process. Equation 13.7.1 and those following it imply that contact between the growing\ndisk and the surrounding melt for time t is sufficient for crystallization. Another possibility is\nthat allowance must be made for the diffusion of crystallizable molecules to (or noncrystallizable\n"]], ["block_9", ["Solve Equation 13.7.8 for K and evaluate at t: 103 s for each of the m values: K [\u2014ln(1 \u00a2C)]/\nt\u2019\". For m 2, K (1n 0.5)/(103)2 6.93 X 10\u20147 s\u2014Z; for m = 3, K 6.93 x 10\u2014 3\u20143; for m 4,\nK 6.93 x 10\u2014 3\u20144. Note that the units of K depend on the value of m. Solve Equation 13.7.8 for\nt and evaluate at different 6150\u2019s for the m and K values involved.\nThese three systems describe a set of crystallization curves that cross at 650 0.5 and t: 103 s,\n"]], ["block_10", ["m=2\nm=3\nm=4\n6,\n(K=6.93 x 10\u20147)\n(K=6.93 x 10\u201410)\n(K=6.93 x 10\u201413)\n"]], ["block_11", ["as shown in Figure 13.31. For the case where m 2, the time interval over which the change occurs\nis widest (1430 s from 65, 0.1\u20140.9) and the maximum slope is smallest (7.8 x 10\u20144 s\u20141 between\n"]], ["block_12", ["Kinetics of Bulk Crystallization\n555\n"]], ["block_13", ["Solution\n"]], ["block_14", ["Example 13.4\n"]], ["block_15", ["956 1\u2014 941%- ;Nr\u2018gtz\u2019)\n(13.7.13)\n"]], ["block_16", ["1\n.\n111(1 _\u00a2) =Z3ZN1\u20182t3\n(13.7.12)\n"]], ["block_17", [{"image_4": "565_4.png", "coords": [47, 45, 144, 95], "fig_type": "molecule"}]], ["block_18", ["t in seconds\n"]]], "page_566": [["block_0", [{"image_0": "566_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "566_1.png", "coords": [34, 432, 107, 476], "fig_type": "molecule"}]], ["block_2", ["While there are several instances of redundancy among the Avrami exponents arising from\ndifferent pictures of the crystallization process, there is also enough variety to make the experi-\nmental value of this exponent a valuable way of characterizing the crystallization process. In the\nnext section we shall examine the experimental side of crystallization kinetics.\n"]], ["block_3", ["If this result is substituted into the previous expressions containing r, the effect is to replace f with\n(6D)\u201d2 and to multiply those t\u2019s that accompany r by Fm.\nThis rather complex array of possibilities is summarized in Table 13.5. Table 13.5 lists the\npredicted values for the Avrami exponent for the following cases:\n"]], ["block_4", ["molecules away from) the growth site. For example, it may be that amorphous molecules\nmust diffuse out of the crystal domain to allow space for the crystallizing molecules. For a\ncrystal of radius r, the time required for molecules to diffuse out of this domain can be\ndetermined from Equation 9.5.1 as r: (6Dt)\u201d2. In Equation 13.7.1 this radius is written r fl.\nThus, if the growth rate is diffusion controlled, these two expressions for r can be equated and\nsolved for r:\n"]], ["block_5", ["1.\nGrowth geometry: 1D (e.g., fibrillar rod), 2D (e.g., disk or sheet), and 3D (e.g., sphere)\n2.\nNucleation mode: simultaneous (heterogeneous) and sporadic (homogeneous)\n3.\nRate determination: contact and diffusion\n"]], ["block_6", ["In order to carry out an experimental study of the kinetics of crystallization, it is \ufb01rst necessary to\nbe able to measure the fraction of polymer crystallized, (1%- While this is necessary, it is not\n"]], ["block_7", ["Figure 13.31\nCrystallized fraction versus time for the indicated Avrami parameters, as discussed in\nExample 13.4.\n"]], ["block_8", ["\u00a2c\n"]], ["block_9", ["13.7.2\nKinetics of Crystallization: Experimental Aspects\n"]], ["block_10", ["556\nCrystalline Polymers\n"]], ["block_11", [{"image_2": "566_2.png", "coords": [43, 55, 307, 321], "fig_type": "figure"}]], ["block_12", ["1\": (7)\n(13.7.14)\n"]], ["block_13", ["0.8 \n"]], ["block_14", ["0.6 \n"]], ["block_15", ["0.4\n\u2014\n"]], ["block_16", ["0.2 \n"]], ["block_17", ["0\n---_-'5-3\n_|_\nl\nL\n"]], ["block_18", [{"image_3": "566_3.png", "coords": [61, 82, 153, 160], "fig_type": "figure"}]], ["block_19", ["0\n500\n1 000\n1 500\n2000\n"]], ["block_20", ["Time (s)\n"]]], "page_567": [["block_0", [{"image_0": "567_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "567_1.png", "coords": [30, 59, 434, 250], "fig_type": "figure"}]], ["block_2", ["(be, and that Pf; must be available, at least by extrapolation. The heat of fusion is an example of a\nproperty of the crystalline phase that could be used this way. However, it might be difficult to show\nthat the value of AH? is constant per unit mass at all percentages of crystallinity, and to obtain a\nvalue for AH? for a crystal free from defects. Therefore, while conceptually simple, the actual\nutilization of Equation 13.7.15 in precise work may not be straightforward.\nFigure 13.32b shows a variation in which a property of the sample (no subscript) is found to\nvary linearly with qbc, having a value Pa when (35., 0 and a value PC when (1)., 1. The slope of this\nline is simply PC \u2014Pa, since the difference of 9150 values is unity for this difference in P. The\nequation for the line in Figure 13.32b is\n"]], ["block_3", ["Specific volume (or density) is an example of a property that has been extensively used in this way\nto evaluate the. If the amorphous component contributes nothing to the measured property (as with\nthe heat of fusion), then Equation 13.7.17 reduces to Equation 13.7.15.\nFigure 13.320 illustrates how x\u2014ray diffraction techniques can be applied to the problem\nof evaluating (be. If the intensity of scattered x-rays is monitored as a function of the angle of\ndiffraction, a result like that shown in Figure 13.320 is obtained. The sharp peak is associated with\n"]], ["block_4", ["sufficient; we must also be able to follow changes in the fraction of crystallinity with time. So far\nin this chapter we have said nothing about the experimental aspects of determining (be. We shall\nnow brie\ufb02y rectify this situation by citing some of the methods for determining (be. It must be\nremembered that not all of these techniques will be suitable for kinetic studies.\nSince the fractions of crystalline (subscript c) and amorphous (subscript a) polymer account for\nthe entire sample, it follows that we may measure whichever of the two is easiest to determine, and\nobtain the other by difference. Generally, it is some property PC of the crystalline phase that we are\nable to monitor. If this property can be measured for a sample that is 100% crystalline (superscript \u00b0),\nwe can compare the value of PC measured on an actual sample (no superscript) to evaluate gbc:\n"]], ["block_5", ["Avrami exponent\nCrystal geometry\nNucleation mode\nRate determination\n"]], ["block_6", ["Table 13.5\nSummary of Exponents in the Avrami Equation (Equation 13.7.8) for Different\nCrystallization Mechanisms\n"]], ["block_7", ["0.5\nRod\nSimultaneous\nDiffusion\n"]], ["block_8", ["1\nDisk\nSimultaneous\nDiffusion\n1.5\nSphere\nSimultaneous\nDiffusion\nl .5\nRod\nSporadic\nDiffusion\n2\nDisk\nSimultaneous\nContact\n2\nDisk\nSporadic\nDiffusion\n2\nRod\nSporadic\nContact\n2.5\nSphere\nSporadic\nDiffusion\n"]], ["block_9", ["This relationship is sketched in Figure 13.32a, which emphasizes that Pc must vary linearly with\n"]], ["block_10", ["which can easily be solved for (be as a function of P, Pa, and PC:\n"]], ["block_11", ["3\nSphere\nSimultaneous\nContact\n"]], ["block_12", ["3\nDisk\nSporadic\nContact\n4\nSphere\nSporadic\nContact\n"]], ["block_13", ["Kinetics of Bulk Crystallization\n557\n"]], ["block_14", ["1\nRod\nSimultaneous\nContact\n"]], ["block_15", [{"image_2": "567_2.png", "coords": [35, 558, 114, 599], "fig_type": "molecule"}]], ["block_16", ["P Pa + \u00a2C(pc pa)\n(13.7.16)\n"]], ["block_17", ["P Pa\n= _E\n13.7.17\na.\nP. P.\n(\n>\n"]], ["block_18", ["PC\n= __\n13.7.15\n"]], ["block_19", [{"image_3": "567_3.png", "coords": [234, 81, 420, 238], "fig_type": "figure"}]]], "page_568": [["block_0", [{"image_0": "568_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "568_1.png", "coords": [25, 67, 292, 181], "fig_type": "figure"}]], ["block_2", [{"image_2": "568_2.png", "coords": [27, 79, 154, 171], "fig_type": "figure"}]], ["block_3", ["the crystalline diffraction, and the broad peak with the amorphous contribution. If the area A under\neach of the peaks is measured, then\n"]], ["block_4", ["An obvious difficulty here arises in deciding the location of the broken-line portions of the peaks in\nthe region of overlap. Some features of the infrared absorption spectrum may also be analyzed by\nthe same procedure to yield values for (350.\nAs noted above, not all techniques that provide information regarding crystallinity are useful to\nfollow the rate of crystallization. In addition to possessing sufficient sensitivity to monitor small\nchanges, the method must be rapid and suitable for isothermal regulation, quite possibly over a\nrange of different temperatures. The Spectrosc0pic techniques of infrared absorption, Raman\nscattering, and NMR have all been used successfully for this purpose, as has WAXS with a\nsynchrotron source. Specific volume measurements are also convenient, and we shall continue\nthis discussion using specific volume as the experimental method.\nAlthough the extent of crystallinity is the desired quantity, time is the experimental variable.\nAccordingly, what is done is to identify the specific volume of a sample at t 0 (subscript 0) with Va,\nthe volume at r 00 (subscript 00) with VC, and the volume at any intermediate time (subscript r)\nwith the composite volume. On this basis, Equation 13.7.17 becomes\n"]], ["block_5", ["Figure 13.33a shows how this quantity varies with time for polyethylene crystallized at a series of\ndifferent temperatures. Several aspects of these curves are typical of all polymer crystallizations\nand deserve comment:\n"]], ["block_6", ["and the amorphous fraction becomes\n"]], ["block_7", ["Figure 13.32\nVarious representations of the properties of mixture of crystalline and amorphous polymer.\n(a) The monitored property is characteristic of the crystal and varies linearly with (be. (b) The monitored\nproperty is characteristic of the mixture and varies linearly with (be between Pa and PC. (c) X-ray intensity is\nmeasured with the sharp and broad peaks being PC and Pa, respectively.\n"]], ["block_8", ["1.\nThe decrease in amorphous content follows an S-shaped curve. The corresponding curve for\nthe growth of crystallinity would show a complementary but increasing plot. This aspect of the\nAvrami equation was noted in connection with the discussion of Equation 13.7.8.\n2.\nThe greater the undercooling, the more rapidly the polymer crystallizes, as discussed in\nSection 13.5. Although the data in Figure 13.33 are not extensive enough to show it, this\ntrend does not continue without limit. As the crystallization temperature is lowered still\n"]], ["block_9", ["(C)\nb\n( )\n0\n<90\n1\nAngle\na( )\n0\n\u201890\n1\n"]], ["block_10", ["558\nCrystalline Polymers\n"]], ["block_11", [{"image_3": "568_3.png", "coords": [37, 470, 121, 509], "fig_type": "molecule"}]], ["block_12", ["Ac\n(be \n13.7.\nAc+Aa\n(\n18)\n"]], ["block_13", ["V00 V,\n1 \n: -\u2014\u2014\n13.7.20\n"]], ["block_14", ["Vt V0\n: __\n13.7.19\n"]], ["block_15", [{"image_4": "568_4.png", "coords": [138, 79, 269, 173], "fig_type": "figure"}]], ["block_16", [{"image_5": "568_5.png", "coords": [161, 42, 433, 177], "fig_type": "figure"}]], ["block_17", ["Intensity\n"]], ["block_18", [{"image_6": "568_6.png", "coords": [282, 47, 421, 161], "fig_type": "figure"}]]], "page_569": [["block_0", [{"image_0": "569_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The preceding example of superpositioning is an illustration of the principle of time\u2014temperature\nequivalence, as was discussed extensively in Section 12.5 in connection with the viscoelastic\nbehavior of polymer samples. The current application differs in that the reason for the strong change\nin rate with temperature is the thermodynamic driving force, as well as partly to the reduction in\nfree volume as Tg is approached, but the basic idea of time\u2014temperature equivalence is the same.\nNow let\nus examine an experimental test of the Avrami equation and the assortment\nof predictions from its various forms as summarized in Table 13.5. Figure 13.34 is a plot of\nIn [ln(1 (350)\u20141] versus lnt for poly(ethylene terephthalate) at three different temperatures. This\nformat is suggested by rearranging Equation 13.7.8, and then taking the natural logarithm twice:\n"]], ["block_2", ["further, the rate of crystallization passes through a maximum and then drops off as Tg is\napproached, as illustrated in Figure 13.21.\n3.\nBecause bulk samples never become 100% crystalline, there is a potential ambiguity to\nquantities like V00; does it refer to 100% crystallinity, a state that is not achievable, or to the\nfinal value that is actually attained? Clearly the value of qbc determined via Equation 13.7.19\nwill depend on the meaning employed. Depending on circumstances, either interpretation can\nbe useful, but it is important to be clear as to which one is being used.\n4.\nReplotting the data on a logarithmic timescale has an interesting effect: Figure 13.33b shows\nthat this modi\ufb01cation produces a far more uniform set of S curves. As a matter of fact, if the\nvarious curves are shifted along the horizontal axis, they may be superimposed to produce a\nreasonable master curve. By comparing the times corresponding to 50% crystallinity at 120\u00b0C\nand 130\u00b0C, there is a shift of over a factor of 1000. In other words, increasing the undercooling\nby 10 degrees increases the rate of crystallization by more than three orders of magnitude.\n"]], ["block_3", [{"image_1": "569_1.png", "coords": [34, 328, 154, 392], "fig_type": "molecule"}]], ["block_4", ["Figure 13.33\nFractions of amorphous polyethylene as a function of time for crystallization at the indicated\ntemperatures, plotted on a (a) linear scale\n(con\ufb02nued)\n"]], ["block_5", ["Kinetics of Bulk Crystallization\n559\n"]], ["block_6", ["1\u2014 (be exp( Ktm)\n"]], ["block_7", [">0 0.60\n"]], ["block_8", [">3\n0.50 \n"]], ["block_9", [{"image_2": "569_2.png", "coords": [48, 364, 296, 621], "fig_type": "figure"}]], ["block_10", ["1\nln[ln(\n)] mlnt+ a\n(13.7.21)\n1\u2014 qbc\n"]], ["block_11", ["ln(1 qbc) \u2014Kt\u2019\"\n"]], ["block_12", ["(a)\n0\n100\n200\n300\n400\n500\n600\n700\n800\n900\nTime (min)\n"]], ["block_13", ["020\n125\u00b0\n0.10 \n120\u00b0\n-\n"]], ["block_14", ["0.90\n"]], ["block_15", ["0.30\n"]], ["block_16", ["0.70\n"]], ["block_17", ["(l40--\n"]], ["block_18", ["0.30\n"]], ["block_19", ["1.00\n"]], ["block_20", ["l\nI_\n|\nl\nl\nl\n_|_\nl\n"]]], "page_570": [["block_0", [{"image_0": "570_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "570_1.png", "coords": [26, 1, 277, 330], "fig_type": "figure"}]], ["block_2", ["According to Equation 13.7.21, this representation should yield a straight line, the slope of which\ncorresponds to the Avrami exponent m.\nThe data in Figure 13.34 show that linearity is indeed obtained and that the slope equals 2 when\nthe crystallization is carried out at 110\u00b0C and changes to 4 at higher temperatures. The melting\npoint of poly(ethy1ene terephthalate) is 267\u00b0C, so that when the undercooling is about 25\u00b0C, three-\ndimensional growth with sporadic nucleation is suggested. With an undercooling of 150\u00b0C, the\nmechanism of crystallization is clearly different, although it is not possible to identify the specific\ncombination of factors responsible for the exponent 2. The values of K in Equation 13.7.21 are best\nobtained analytically, once the exponent has been determined graphically. The two K values for the\ncase where m=4 in Figure 13.34 are 2.94 x 10\"7 min\"4 at 236\u00b0C and 3.13 X 10\u20183 min\u20184 at\n240\u00b0C. The mechanism is apparently the same in these two cases, but the rate is more than nine\ntimes faster when the temperature is lowered by only 4\u00b0C. Note that it is not possible to resolve K,\nwhich is a cluster of nucleation and growth parameters, into its constituent factors, even when the\nvalue of the exponent identifies the mechanism unambiguously. At both 110\u00b0C and 120\u00b0C (not\nshown), H122 and the values of K are 7.93 x 10\"4 and 7.45 x 10\"3 min\u20142, respectively. In this\nregion the rate is about 10 times slower when the temperature is lowered by 10\u00b0C. Thus, both the\nvalue of m and the effect on K of changing temperature are different for these two regimes of\nbehavior.\nThe testing of the Avrami equation reveals several additional considerations of note:\n"]], ["block_3", ["1.\nThe multiple use of logarithms in the analysis presented by Figure 13.34 can obscure much of\nthe deviation between theory and experiment. More stringent tests can be performed by other\nnumerical methods.\n2.\nDeviations from the Avrami equation are frequently encountered in the long time limit of\nthe data.\n"]], ["block_4", ["Figure 13.33 (continued)\n(b) logarithmic scale. (Reprinted from Mandelkern, L., Growth and Perfection\n0fCrystals, Doremus, R.H., Roberts, B.W., and Tumbull, D. (Eds.), Wiley, New York, 1958. With permission.)\n"]], ["block_5", ["(b)\nlog TU, min)\n"]], ["block_6", ["560\nCrystalline Polymers\n"]], ["block_7", ["0\n0.20\nI\n\u00b0\n"]], ["block_8", ["0.90\n"]], ["block_9", ["0.80\n"]], ["block_10", ["0.70\n"]], ["block_11", ["1.00\n"]], ["block_12", ["1\n1o\n102\n1o3\n104\n"]], ["block_13", ["I\nI\n"]]], "page_571": [["block_0", [{"image_0": "571_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nExponents other than integral multiples of one\u2014half are observed. In fact, a method for\ndetermining the Avrami exponent, which is based on graphical differentiation rather than\nlogarithmic analysis yields instantaneous m values at particular values of (be rather than a\nsingle value averaged over the entire transition. When this method is used, it is found that m\nincreases initially, before eventually leveling off.\n4.\nThese unpredicted Avrami exponents may be indications that multiple mechanisms are\noperative or that f or N is itself a function of gbc.\n5.\nIn general, one must exercise caution in inferring too much about the crystallization process\nfrom the Avrami analysis alone. This situation is analogous to that touched on in considering\npolymerization kinetics in Chapter 2 through Chapter 4, namely that it is dangerous to infer\na polymerization mechanism from the resulting molecular weight distribution alone. Among\nthe difficulties to bear in mind are the following: different mechanisms can lead to the same\nexponent; the nucleation may be due to a combination of heterogeneous and homogenous\nprocesses; each spherulite contains both crystalline and amorphous material; and the relative\nproportion may change with time.\n"]], ["block_2", ["Figure 13.34\nLog\u2014log plot of ln(1-q\u00a7c)_1 versus time for poly(ethy1ene terephthalate) at three different\ntemperatures. (Reprinted from Morgan, L.B., Philos. Trans. R. Soc. London, 247A, 13, 1954. With permission.)\n"]], ["block_3", ["In this chapter we have examined many aspects of the fascinating field of polymer crystallization.\nThe main topics emphasized were the complex structural features of crystalline polymers, the\ninterplay of thermodynamic and kinetic factors that dictate the structural details, and an introduction\nto techniques for the experimental characterization of both structure and crystallization kinetics:\n"]], ["block_4", ["Chapter Summary\n561\n"]], ["block_5", ["13.8\nChapter Summary\n"]], ["block_6", [{"image_1": "571_1.png", "coords": [36, 37, 233, 317], "fig_type": "figure"}]], ["block_7", ["1.\nAt the smallest structural length scale, the unit cell, individual chains form helices to minimize\nintramolecular energetic constraints, and the helices pack together to maximize intermolecular\n"]], ["block_8", ["I \n\u2018\no\nE\n\u20142.0\n-\n'\n.\n"]], ["block_9", ["6\u2019\nx!\no\n1-\n|\n_\n.\n"]], ["block_10", [{"image_2": "571_2.png", "coords": [37, 41, 97, 314], "fig_type": "figure"}]], ["block_11", ["+2.0 \n"]], ["block_12", ["\u20144.0 \n-'\n\u00b0\n"]], ["block_13", ["_5_0\n1\nI\n[\nI\n1\n0\n2\n4\n6\n"]], ["block_14", ["0\n'\u2014\nI\"\na.\n"]], ["block_15", ["\u2014\n110\u00b0\n.\n'\n"]], ["block_16", ["__\n236\u00b0\n240\u00b0C\n"]], ["block_17", [{"image_3": "571_3.png", "coords": [119, 52, 215, 285], "fig_type": "figure"}]], ["block_18", ["ln t(t, min)\n"]]], "page_572": [["block_0", [{"image_0": "572_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["\"KH. Illers and H. Hendus, Makromot. Chem, 113, 1 (1968).\n"]], ["block_2", ["interactions. Polymer unit cells represent many of the 230 possible space groups, except those\nwith cubic symmetry, and it is not unusual for a given polymer to exhibit two or more different\npolymorphs under different conditions.\n2.\nOn intermediate length scales the unit cells are organized into chain-folded lamellae, such that\nportions of the individual chain backbones, or stems, lie parallel to each other and approxi-\nmately parallel to the thin axis of the lamella. Within an isothermally crystallized sample, the\nlamellae are of roughly constant thickness, but the thickness varies from a few tens to a few\nhundreds of angstroms, depending on crystallization conditions. Upon exiting the lamellar\nsurface, a particular chain may execute a tight fold back into the crystal to become an adjacent\nstem, or it may wander off to reenter the same crystal at a different site, or even a different\nlamella. The prevalence of adjacent reentry is much greater in solution-grown single crystals\nthan in melt-crystallized materials, and in crystals grown at smaller undercoolings.\n3.\n0n larger length scales the lamellae grow into spherulites, which may be viewed and\ncharacterized with a polarizing optical microscope. Other morphologies such as dendrites,\nhedrites, and shish kebabs can also be observed under particular conditions. A bulk sample\nnever becomes 100% crystalline, and the lamellae are interspersed with amorphous regions\nthat can often comprise the majority of the material.\n4.\nPolymer crystals emerge by a process of nucleation and growth. The nucleation may\nbe heterogeneous, homogeneous, or a combination of both. The barrier to homogeneous\nnucleation is dependent on the competition between bulk and surface contributions to the\nfree energy; in general nucleation is more rapid, and the critical nuclei size smaller, the greater\nthe undercooling. The growth process at the level of an individual lamella is still not fully\nunderstood, but in many cases the temperature dependence of the growth rate can be inter-\npreted via the competition between the rate of addition of a single stem to a growth face and\nthe rate of adding stems at neighboring sites.\n5.\nThe overall evolution of crystallinity often follows the Avrami equation, in which the type of\nnucleation, the spatial dimensionality of growth, and the presence or absence of diffusion\nlimitations interact to yield a particular Avrami exponent. At relatively small undercoolings\nthe rate of crystallization increases as temperature decreases, but eventually the rate decreases\nand vanishes as the approach to the glass transition inhibits any kind of chain or segmental\nmotion. The overall rate of crystallization is typically independent of molecular weight\nfor high molecular-weight polymers, but increases with decreasing molecular weight for\nshorter chains.\n6.\nA wide variety of experimental tools are useful in the study of polymer crystallization. We\nhave highlighted the use of electron microscopy to visualize single crystals in exquisite detail,\nand the use of x-ray and electron diffraction methods to characterize the unit cell structure.\n"]], ["block_3", ["Problems\n"]], ["block_4", ["562\nCrystalline Polymers\n"]], ["block_5", ["1.\nIllers and Hendus measured the melting points of polyethylene crystals whose thickness was\nvaried by controlling the conditions of crystallization, and which was measured by x-ray\ndiffraction. The following results were obtained:T\n"]], ["block_6", ["Tmo(\u00b0C)\n139.4\n137.5\n136.0\n134.9\n131.9\n127.9\n117.9\nE (A)\n1750\n758\n481\n392\n258\n177\n100\n"]], ["block_7", ["Prepare a plot of Tm versus 1?\", and from the slope and intercept, respectively, evaluate T3?\nand y from these data. Compare the values obtained with quantities given in Table 13.3 and\nExample 13.3.\n"]]], "page_573": [["block_0", [{"image_0": "573_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["10.\n"]], ["block_2", ["Problems\n"]], ["block_3", ["be valid, and why might they not apply?\n_\nFollow through the analysis of Equation 13.5.7 to \n"]], ["block_4", ["13.5.9.\nIn the context of Figure\n13.18 and associated discussion,\nnucleation rate would change for 20% change in E away \n.\n.\nIsotactic polypropylene crystallizes in 3/1 helices in a monoclinic \n"]], ["block_5", ["Figure 13.7. Individual helices can either be right\u2014handed \n"]], ["block_6", ["getically degenerate. In this particular polymorph, \n"]], ["block_7", ["polymer can also be induced to crystallize in a hexagonal \n"]], ["block_8", ["crystal, emerging and folding \n"]], ["block_9", ["Molecular weights for the two shortest chains observed \n"]], ["block_10", ["of the unit cell along the chain axis, 2.53 131, as the distance \n"]], ["block_11", ["latter with the crystal thickness determined by x\u2014ray diffraction, \n"]], ["block_12", ["of chain lengths for the first and second peaks suggest \n"]], ["block_13", ["In discussing Figure 13.8 and Figure 13.9, the issue was linear\nHoffmann\u2014Weeks extrapolation. What conditions should \n"]], ["block_14", ["The resulting chain\nfragments\nare\nseparated chromatograhically\nand their \n"]], ["block_15", ["According to this picture, the shortest chain showing \n"]], ["block_16", ["some amount, which measures the length of the loop made \n"]], ["block_17", ["subtracted from each of these molecular weights to \n"]], ["block_18", ["Calculate the degree of polymerization of each molecule \n"]], ["block_19", ["equal the crystal thickness in length. The second shortest chain tWice \n"]], ["block_20", ["polymer crystals with fuming nitric acid, which cleaves fold \n"]], ["block_21", ["could have a lower free energy than a straight-chain crystal, 'of\nchain end placement and to the avoidance of the need for the \n"]], ["block_22", ["degree of polymerization.\n.\n.\nChemical evidence for chain folding in polyethylene \n"]], ["block_23", ["to bring about a 1\u00b0 change in Tm, all other things being \n.\nFor n-alkanes, there is little doubt that the lowest free energy \n"]], ["block_24", ["proposition: for a typical polydisperse sample of linear polyethylene, \n"]], ["block_25", ["weights determined by osmometry. The folded chain \n"]], ["block_26", ["straight-chain lamella. For high molecular\u2014weight polyethylenes, same\nargument, if the sample were perfectly monodisperse. Criticize \n"]], ["block_27", ["Polystyrene\n104.1\n\u2014\n\u2014\n6.63\n18\n1.126\n"]], ["block_28", ["Polyisobutene\n56,1\n6.94\n11.96\n_\n16\n0,937\n"]], ["block_29", ["Poly(vinyl chloride)\n62.5\n10.11\n5.27\n5.12\n4\n\u2014\n"]], ["block_30", ["Nylon 8\n\u2014\n4.9\n4.9\n22\n2\n1-033\nPoly(methy1\n100.1\n21.08\n12.17\n10.55\n\u2014\u2014\n1-23\n"]], ["block_31", ["For polyethylene, make an estimate of how much the molecular \n"]], ["block_32", ["this? What would you predict for 4/1 helices?\nThe polymers listed below are all known to form unit cells in \nUse this fact plus the data given to complete the following table:\n"]], ["block_33", ["cell. For example, 3/1 and 6/1 helices often lead to hexagonal \n"]], ["block_34", ["Polymer\nMD\na\nb\nC M\n"]], ["block_35", ["There is a general correlation between the nature of the helix \n"]], ["block_36", ["1260 and 2530. Since the cleaved chains end in nitro \n"]], ["block_37", ["methacrylate)\n"]], ["block_38", ["Number \nDensity\n"]], ["block_39", ["estimate \n"]], ["block_40", ["all of the \n"]], ["block_41", ["563\n"]]], "page_574": [["block_0", [{"image_0": "574_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1LED. Hartley, F.W. Lord, and LB. Morgan, Phil. Trans. R. Soc. London, 247A, 23 (1954).\n1A. Keller, G.R. Lester, and LB. Morgan, Phil. Trans. R. Soc. London, 247A, 1 (1954).\n\u00a7P. Parrini and G. Corrieri, Makromol. Chem, 62, 83 (1963).\n"]], ["block_2", ["564\n"]], ["block_3", ["11.\n"]], ["block_4", ["12.\n"]], ["block_5", ["13.\n"]], ["block_6", ["14.\n"]], ["block_7", [{"image_1": "574_1.png", "coords": [56, 438, 242, 554], "fig_type": "figure"}]], ["block_8", ["Calculate the time required for (,bc to reach values of 0.1, 0.2, ..., 0.9 for each of these\nsituations. Graph qbc versus I using the results calculated at 110\u00b0C and 240\u00b0C, plotting both in\nthe same figure. Because of the much larger K at 180\u00b0C, the crystallization occurs much more\nrapidly at this temperature than at either 110\u00b0C or 240\u00b0C. Multiply each of the times\ncalculated at 180\u00b0C by the arbitrary constant 60 and plot the data thus shifted on the same\ncoordinates as the other curves. What generalization appears concerning the relative slopes at\n"]], ["block_9", ["the helices pack? Some possibilities to consider are (a) Three right-handed and three left-\nhanded helices occupy particular relative positions within the unit cell. (b) The crystalline\npacking does not care; right and left are mixed randomly. (c) Only right-handed or left-\nhanded forms crystallize\ntogether;\neach lamella\nis pure right or left.\nExplain your\nreasoning.\nConsider a polymer that tends to crystallize in needle-shaped crystals, rather than in lamellae.\nAssuming the needles have length L and radius R, with R/L << 1, derive the dependence of the\nmelting temperature Tm on R. Why is the answer the same/different from the dependence of\nTm on the lamella thickness?\nThe crystallization of poly(ethylene terephthalate) at different temperatures after prior fusion\nat 294\u00b0C had been observed to follow the Avrami equation with the following parameters\napplying at the indicated temperatures?r\n"]], ["block_10", ["(pa 0.5?\nPoly(ethy1ene terephthalate) was crystallized at 110\u00b0C and the densities were measured after\nthe indicated time of crystallization.i Using density as the property measured to determine\ncrystallinity, evaluate (1'), as a function of time for these data. By an appropriate graphical\nanalysis, determine the Avrami exponent (in doing this, ignore values of (150 < 0.15, since\nerrors get out of hand in this region). Calculate (rather than graphically evaluate) the value of\nK consistent with your analysis.\n"]], ["block_11", ["0\n1.3395\n35\n1.3578\n"]], ["block_12", ["15\n1.3438\n50\n1.3655\n20\n1.3443\n60\n1.3675\n25\n1.3489\n70\n1.3685\n30\n1.3548\n80\n1.3693\n"]], ["block_13", ["The crystallization rate of isotactic polypropylene (MW 181,000, Tm =172\u00b0C) was studied\nunder various patterns of temperature change.\u00a7 Solids were melted at Tf, held at Tf for 1 h,\nand then crystallized at TC. The following Avrami exponents were observed:\n"]], ["block_14", ["110\n2\n3.49 x 10\u20144\n180\n3\n1.35\n240\n4\n5.05 x 10*8\n"]], ["block_15", ["5\n1.3400\n40\n1.3608\n10\n1.3428\n45\n1.3625\n"]], ["block_16", ["t (min)\np (g cm_\u00b0)\nt (min)\np (g cm\u201c\u00b0)\n"]], ["block_17", ["T (\u00b0C)\nm\nK (min)\n"]], ["block_18", ["Crystalline Polymers\n"]]], "page_575": [["block_0", [{"image_0": "575_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["3.\nGeil, P.H., Polymer Single Crystals, Wiley, New York, 1963.\nAvrami, M. J. Chem. Phys., 7, 1103, (1939); 8, 212, (1940); 9, 177 (1941).\n5.\nEvans, U.R. Trans. Faraday $06., 41, 365 (1945).\nP\n"]], ["block_2", ["Bassett, D.C., Principles of Polymer Morphology, Cambridge University Press, Cambridge, UK, 1981.\nGeil, P.H., Polymer Single Crystals, Wiley, New York, 1963.\nMandelkem, L., Crystallization ofPolymers, McGraw-Hill, New York, 1964.\nSchultz, J.M., Polymer Crystallization, Oxford University Press, New York, 2001.\nTadokoro, H., Structure of Crystalline Polymers, Wiley-Interscience, New York, 1979.\nWunderlich, B., Macromolecular Physics, Vol I: Crystal Structure, Morphology, Defects, Academic Press,\n"]], ["block_3", ["l\u2014 I\nNatta, G. and Corradini, P. J. Polym. Sci., 39, 29 (1959).\n2.\nLauritzen, J.l. and Hoffmann, J.D., J. Res. Natl. Bur. Stds., A64, 73 (1960); Hoffmann, JD. and Lauritzen,\nJ.I., J. Res. Natl. Bur. Stds., A65, 297 (1961).\n"]], ["block_4", ["Further Readings\n565\n"]], ["block_5", ["References\n"]], ["block_6", ["Further Readings\n"]], ["block_7", ["15.\n"]], ["block_8", ["16.\n"]], ["block_9", ["17.\n"]], ["block_10", [{"image_1": "575_1.png", "coords": [44, 77, 309, 156], "fig_type": "figure"}]], ["block_11", ["Suppose a polymer spherulite grew by wrapping each chain tightly around and around the\nsurface like a ball of string. What would you expect to see in a polarizing microscope? What\nwould you expect to see if the chains stretched straight out along the radial direction from the\ncenter?\nIn understanding the mechanical properties of metals and alloys, crystal defects such as\ndislocations play a key role. Although polymer crystals certainly exhibit many analogous\nstructural defects, such defects play almost no role in discussions of the mechanical proper-\nties of polymer materials. Why is this?\nFor a crystal that grows in an n\u2014dimensional space, with dimensions R1, ..., Rn along its\nvarious facets, show that the melting temperature Tm always varies as l/Rj, where R}- is the\nsmallest dimension.\n"]], ["block_12", ["Avrami exponent\nTr (\u20180\nTC :150\u00b0C\nT, = 155\u00b0C\nr, 160\u00b0C\n"]], ["block_13", ["190\n\u2014\n3.1\n3.5\n210\n2.9\n3.3\n4.1\n220\n3.1\n3.8\n\u2014\n230\n3.1\n4.0\n\u2014\n"]], ["block_14", ["On the basis of these observations, criticize or defend the following propositions:\n"]], ["block_15", ["3.\nChanges in m are consistent with the idea that under some conditions, nuclei from the\noriginal solid survive the period in the melt and nucleate the recrystallization.\n"]], ["block_16", ["1.\nWhen both T1: and TC were low, the Avrami exponents are consistent with three-\ndimensional growth on contact with sporadic nucleation.\n2.\nThe change in m can be interpreted as arising from a change in either the growth\ngeometry or nucleation situation. That is, the change in m for [Tf and TC low]\u2014> [Tf and\nTc high] could arise from either the change spherical ~+ disk geometry or the change\nsporadic \u2014> simultaneous nucleation.\n"]], ["block_17", ["New York, 1973.\n"]]], "page_576": [["block_0", [{"image_0": "576_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "576_1.png", "coords": [32, 181, 151, 223], "fig_type": "molecule"}]], ["block_2", ["Then, when f(x) (1 +x)\" we have\n"]], ["block_3", ["The natural logarithm of (1 +x) where |x| < 1 also arises often. Recall that d(ln x)/dx= 1/15, and\nthat d(x_ \u2018)/dx ix\n\u2014 0 + 1):\n"]], ["block_4", ["Many common functions (such as sin 1:, cos x, ex, 1n (1 + x) \n- -) can be represented by power series,\ni.e., a sum of terms with increasing powers of the relevant argument x. Such series are useful in\nallowing the function itself to be replaced by an algebraically simple approximation appropriate in\nsome limit (e.g., x\u2014> 0, x\u2014a 1, x\u2014> 00). These series approximations can be looked up in many\nhandbooks, but they can also often be derived from the McLaurin series. A functionf(x) is said to\n"]], ["block_5", ["A.1\nSeries Expansions\n"]], ["block_6", ["be analytic if all derivatives (first, second, third,...) exist over the relevant range of x. The\nMcLaurin series representation of an analytic function f(x) is given by\n"]], ["block_7", ["where\nthe\nith\nderivative of f(x)\nis\nto\nbe\nevaluated\nat i: 0,\nand\nwhere\n2' factorial\nis\nilzix (i\u20141)><(i\u2014-2) >< --->< 1.Bydefinition,0!=1.\nAs an example, consider ex, and recall that d(e\u201c)/dx ex. Therefore from Equation A.l.1\n"]], ["block_8", ["Series expansions for trigonometric functions can also be readily obtained, recalling that\nd(sin x)/dx :cos x, d(cos x)/dx sin x, sin 0 0, and cos 0 1:\n"]], ["block_9", [{"image_2": "576_2.png", "coords": [39, 402, 270, 455], "fig_type": "molecule"}]], ["block_10", [{"image_3": "576_3.png", "coords": [41, 261, 215, 319], "fig_type": "molecule"}]], ["block_11", ["f(X)=:.\u2014,1(dDg0xi\n(A11)\n"]], ["block_12", ["1\n1\n1\ncosx\n\u2014\n6\nCOS(0)JCO \n1\u2014 sin(0)x1\u2014 El cos (0)}:2 +\n"]], ["block_13", ["1\n1\n1\nex_\n0_! H(0)0_|_1_\n\u201c(0)1737?\n"]], ["block_14", ["1\n1\n1\n,\n2\nsinx\u2014\n\u2014 a\nsin(0)x0 + E\ncos (0)x1 \u2014\n2!\n3111(0)); + . .\n"]], ["block_15", ["1\nl\n1\n1\nl\nl\n16\u2014!1\nl\n0\n\u2014\nx1\n\u2014\u2014\u2014\u2014\u20142\nn( +x)=\nn(\n+))cO +1!(1+0)x\n2!(1+0)2x\n+\n"]], ["block_16", ["1'x3 +\n(A.1.2)\n1\n:l+x+ 7x2 +3\u2014\n2\n"]], ["block_17", ["1\n:x__x3+_,5+...\n(A.1.3)\n"]], ["block_18", [{"image_4": "576_4.png", "coords": [74, 491, 323, 543], "fig_type": "molecule"}]], ["block_19", ["__\n12\n13\n"]], ["block_20", ["\u2014-X\u2014\u2014-2-X\n+\u00a7X \n"]], ["block_21", ["1\n1\n"]], ["block_22", ["ex(0)2+_.\n"]], ["block_23", ["Appendix\n"]], ["block_24", ["567\n"]]], "page_577": [["block_0", [{"image_0": "577_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "577_1.png", "coords": [27, 417, 229, 449], "fig_type": "molecule"}]], ["block_2", ["These results can be readily extended to related functions, for example by replacing x with \u2014x, ax,\nor a complex number 2.\n"]], ["block_3", ["These arise in several contexts, especially molecular weight distributions. For example, let x, be the\nmole fraction of i-mer in a polycondensation that follows the most probable distribution (Equation\n2.4.1),\n"]], ["block_4", ["(Note an important but subtle point: the mole fraction of i-mer only makes sense for i 2 1, but the\nsummation above runs from i O. This is because the sum ofp\"-1 starting from i: 1 is the same as\nthe sum ofpi starting from i: 0, and the solution is easier to obtain in the latter case.) To show that\nthis is, in fact, correct, consider a slightly different, \ufb01nite sum:\n"]], ["block_5", ["A2\nSummation Formulae\n"]], ["block_6", ["where p is the probability that a monomer has reacted. Are we sure that this distribution is\nnormalized, that is does\n"]], ["block_7", ["and therefore\n"]], ["block_8", ["If we multiply S] by p and subtract it from S 1, we have a term\u2014by-term cancellation:\n"]], ["block_9", ["Of course, for the polymerization case p will always be <1.\nTo obtain the number average degree of polymerization, we require the related summation\n(Equation 2.4.4)\n"]], ["block_10", ["Comparison with the distribution expression therefore requires that\n"]], ["block_11", ["568\nAppendix\n"]], ["block_12", [{"image_2": "577_2.png", "coords": [38, 264, 180, 300], "fig_type": "molecule"}]], ["block_13", [{"image_3": "577_3.png", "coords": [46, 633, 105, 673], "fig_type": "molecule"}]], ["block_14", ["x.- (1 \u2014p)p\"\u201c\n(A.2.1)\n"]], ["block_15", ["51=ZP\u2018=1+p+p2+p3+---+p\"\n(A23)\ni=0\n"]], ["block_16", ["82 : 2 ip\"\u201c\n(A.2.6)\n"]], ["block_17", ["Zpl :\n(A.2.2)\n.20\n1 \u2014p\n"]], ["block_18", ["00\nZx.=1=(1\u2014p)\u00a7jp\u201c\u2018 ?\ni=1\ni=1\n"]], ["block_19", ["1 \nn+1\nSi \n"]], ["block_20", [" 51-1951=(1+p+p2+'--+p\u201d)-(p+p2+p3 +-~+p\"+\u2018)\n= 1 P\n(A.2.4)\n"]], ["block_21", ["(1+x)\" = &(l +0)\"c0 +%n(l + OYHJCl +%n(n \u20141)(1\u2014l\u20140)\"\u201821r2 + \n"]], ["block_22", ["1\n:\n1\u2014:5\nas )2 \u2014> co (and assuming p < 1)\n(AZ-5)\n"]], ["block_23", ["i=1\n"]], ["block_24", ["1 p\n"]], ["block_25", ["=1+nx+%n(n\u2014l)x2+%n(nw1)(ne2)x3+---\n(AM)\n"]]], "page_578": [["block_0", [{"image_0": "578_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", [{"image_1": "578_1.png", "coords": [16, 216, 327, 344], "fig_type": "figure"}]], ["block_2", ["There are two steps required to transform this integral into spherical coordinates: transform\nf(x,y,z) itself, and transform the volume element dx dy dz. These steps are facilitated by the coordinate\naxes below.\n"]], ["block_3", [{"image_2": "578_2.png", "coords": [34, 82, 275, 150], "fig_type": "molecule"}]], ["block_4", ["The trick here is to recognize ip'.\u2018l as the derivative of pi with respect to p, and that the derivative\nwith respect to p can be taken outside the summation:\n"]], ["block_5", ["Transformation to Spherical Coordinates\n569\n"]], ["block_6", ["In situations where we need to integrate something over all space, and there is no preferred\ndirection, a transformation to spherical coordinates can be extremely useful. A prime example\noccurred in Chapter 6, where we convert the Gaussian distribution function for the end-to\u2014end\nvector into the distribution function for the end\u2014to\u2014end distance. Another instance arose in Chapter\n8, in considering the form factor for an arbitrary particle.\nSuppose we wish to find the integral over all space of some function off(x,y,z):\n"]], ["block_7", ["A3\nTransformation to Spherical Coordinates\n"]], ["block_8", ["Similarly, on the way to obtaining the weight average degree of polymerization we encountered the\nfollowing sum:\n"]], ["block_9", ["and this can be evaluated using the same \u201cderivative trick\u201d:\n"]], ["block_10", ["Figure A.1\nIllustration of the transformation from Cartesian to spherical coordinates.\n"]], ["block_11", [{"image_3": "578_3.png", "coords": [39, 241, 253, 281], "fig_type": "molecule"}]], ["block_12", [{"image_4": "578_4.png", "coords": [41, 512, 309, 643], "fig_type": "figure"}]], ["block_13", [{"image_5": "578_5.png", "coords": [42, 237, 316, 314], "fig_type": "molecule"}]], ["block_14", ["53 Z izpi\u2014l\n(A.2.8)\n"]], ["block_15", [{"image_6": "578_6.png", "coords": [47, 183, 108, 219], "fig_type": "molecule"}]], ["block_16", ["52 =Zi\u00e9\ufb01li\") =% ((23,) \u2018\n1)\nd\nd\n1\n1\n= am \u2018 1) = a (1\u20144;) =W\n(\u201d7)\n"]], ["block_17", ["-2\nt\u2014l\n-\n1\n-\nt\u20141\nl P\n= \nI10\n= \nP\n1P\ni=1\nd\nS)\nd(\np\n)=(l\u2014p)2+2(1\u2014P)P\n:E\n2 \u2018E\n(1\u20141\u00bb?\n (lap): \n1+p\n=\n(A.2.9)\n(1%?)3\n"]], ["block_18", ["2\nZ\nrsin\u00a2d9\n"]], ["block_19", ["J00 J00 J00 f(x,y,z)d.xdydz\n"]]], "page_579": [["block_0", [{"image_0": "579_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The last step was allowed because the argument of IO (and 1,, for all even values of n) is an even\nfunction, that is one for whichf(x) f(\u2014x). The integral of an even function from 0 to 00 is just half\nthe integral from\n\u201400 to 00. Now the integrals extend over all of space, and we make the\n"]], ["block_2", ["An arbitrary point (x, y, z) is represented by a distance from the origin, r, an angle away from the\nx-axis in the x y plane, 6, and an angle away from the z-axis, qb: (r, 6,qb). From the figure it can be\nseen that\n"]], ["block_3", ["Thus, in the case wheref(x, y, 2) can be written as\ufb02at2 + y2 + 22), as with the Gaussian distribution,\nthen f(r, 6, qb) becomes simply f(r).\nThe volume element dx dy dz is now replaced by a volume element with sides dr, r dqb, and r sin\n"]], ["block_4", ["These expressions can be substituted directly into f(x, y, z) to obtain f(r, 6, qb). Note also that\n"]], ["block_5", ["where n is an integer and a is a positive number. The results are quite simple, and can of course\nbe looked up in any table of integrals, but it is actually instructive to work out the answers. In\nso doing, we will utilize the transformation to spherical coordinates just described, as well as\nuse the two most common methods for simplifying integrals: change of variable and integration\nby parts.\nThe hardest one to do is actually the first, namely 10. All of the higher powers can be reduced\nback to this one, as we shall see. We begin by taking [3, and recognizing it can be written as the\nproduct of the same integrals along x, y, and z:\n"]], ["block_6", ["A.4\nSome Integrals of Gaussian Functions\n"]], ["block_7", ["(,1!) d6, as shown in the figure. For a function such as the Gaussian which is only a function of r, the\nintegral over all space can be reduced to a single integral:\n"]], ["block_8", ["A common class of integrals that arose for example in Chapter 6 are these:\n"]], ["block_9", ["570\nAppendix\n"]], ["block_10", [{"image_1": "579_1.png", "coords": [38, 529, 274, 605], "fig_type": "figure"}]], ["block_11", ["x=rsinqbcos\u00a29,\ny=rsia>sin6,\nz=rcosqb\n(A.3.1)\n"]], ["block_12", ["00\n1,, =J\nx\u201d exp(\u2014ax2) dx\n(A.4.1)\n0\n"]], ["block_13", ["x2 + y2 + 22 = r2 (sin2 (M cos2 6 + sin2 9] + 0052 <1?)\n2 r2(sin2qb + COS2 (I?) = \n(A.3.2)\n"]], ["block_14", ["oo\n3\n00\noo\n00\nI3 = (L\nexp(\u2014ax2) dx) (J0\nexp(\u2014ax2) dx) (J0\nexp(\u2014ay2) (1y) (J0\nexp(\u2014azz) dz)\n"]], ["block_15", ["J\nJ\nJ f(x\u2019y\u2019z)dxdd=J\nJ\nJf(r)r23in\u00a2drd6d\u00a2\n\u2014oo\n\u2014oo\n\u2014o-o\n0\nO\n0\n0.0\n211'\n7r\n0'0\n=J my? e\nJ\nsingbdd dqb J for)\"2 dr(21r(\u2014 cosc\ufb02lii)\n0\n0\n0\n0\n"]], ["block_16", [{"image_2": "579_2.png", "coords": [54, 245, 348, 322], "fig_type": "figure"}]], ["block_17", ["CO\n00\n00\n00\n27f\n7r\n"]], ["block_18", ["(X)\n00\n00\n= J\nJ\nJ\nexp(\u2014a[)c2 + y2 + 22]) dxdy dz\n(Pt-4.2)\n0\n0\n0\n"]], ["block_19", ["=1 J00\nJ00\nJ00\nexp(\u2014\u2014a[12 + y2 + 22]) dxdydz\n"]], ["block_20", ["2: 411'J f(r)r2 dr\n(A-3-3)\n0\n"]], ["block_21", ["\u2014OO\n\u201400\n\u201400\n"]], ["block_22", [{"image_3": "579_3.png", "coords": [183, 531, 403, 564], "fig_type": "molecule"}]]], "page_580": [["block_0", [{"image_0": "580_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The situation for odd n can be approached by a change of variable, e.g., u 2x2, du 2x dx:\n"]], ["block_2", ["01'\n_[\u201d:.\n__\nx/TT\nI0_2\\/E,\n12\u20144a\ufb01\n(A.4.7)\n"]], ["block_3", ["In this way, one can arrive at the general formula for even n:\n"]], ["block_4", ["Thus there is another simple relation between 10 and I2. Combining these, we see\n"]], ["block_5", ["transformation to spherical coordinates r, 6, 9b. This is particularly simple in this case, because the\nargument of the integral only depends on r2 =x2 + y2 + 22:\n"]], ["block_6", ["Continuing along this simple line, we apply integration by parts to 12, with u exp(\u2014- 0x2) again\n"]], ["block_7", ["but now v=x3/3 (so dv:x2dx):\n"]], ["block_8", ["and so forth. The general result for odd 1: becomes:\n"]], ["block_9", ["So far, this is not looking promising; we only have a simple relation between [a and 12. However,\nlet us attack 10 directly by integration by parts:\n"]], ["block_10", ["where we make the substitutions at exp(\u2014ax2), v =x, so du :\u20142ax exp(\u2014ax2) dx and dv : dx:\n"]], ["block_11", ["Some Integrals of Gaussian Functions\n571\n"]], ["block_12", [{"image_1": "580_1.png", "coords": [38, 543, 226, 612], "fig_type": "molecule"}]], ["block_13", [{"image_2": "580_2.png", "coords": [38, 335, 154, 374], "fig_type": "molecule"}]], ["block_14", [{"image_3": "580_3.png", "coords": [39, 288, 126, 326], "fig_type": "molecule"}]], ["block_15", [{"image_4": "580_4.png", "coords": [43, 490, 216, 522], "fig_type": "molecule"}]], ["block_16", ["11 = J\nx exp(\u2014ax2)dx J\nexp(\u2014au) du\n0\n2\no\n1\n\u20141\nO,\n1\n=\nE i\n(AA-10)\n"]], ["block_17", ["77\n77\n13=_1 2&1\nA.4.6\n0\n2\n2\n4a\n0\n(\n)\n"]], ["block_18", ["1\n0'0\n00\n00\n1\n0'0\n\u2014 J\nJ\nJ\nexp(\u2014a[x2 + y2 + 22])dx dy dz \u00a7477J\nr2 exp(\u2014\u2014ar2) dr 2 \u00a312\n(A.4.3)\n0\n8\n-00\n\u2014oo\n\u2014\u201400\n"]], ["block_19", ["b\nJudv=uv\n"]], ["block_20", ["00\nX3\n12 J\nA:2 expo\u2014(11:2) dx exp(\u2014ax2) \n"]], ["block_21", ["\u2014 J00 (\u2014 2a)x2 exp (\u2014 0x2) dx\nIO J\nexp (\u2014 axz) dx exp(\u2014\u2014 (13:32)):\n0\n0\n0\n(A45)\n2 0 + 2012\n"]], ["block_22", ["_(n\u20141)(n\u20143)~-(1)\n77\n#\n2(2a)\u201d/2\n3,\neven 11\n(A.4.9)\nn\n"]], ["block_23", ["a\na\n"]], ["block_24", ["2\n= 0 + \u00a314\n(A.4.8)\n"]], ["block_25", ["Nr\u2014I\n"]], ["block_26", ["OO\n1\n00\n"]], ["block_27", ["0\n3\n0\nJ0\n(\n3)\nP\n"]], ["block_28", [{"image_5": "580_5.png", "coords": [77, 402, 356, 451], "fig_type": "molecule"}]], ["block_29", ["b>bb>\nb>bb>\n_\nJ\nv \n(A.4.4)\n"]], ["block_30", ["a\n"]], ["block_31", ["1\n"]], ["block_32", ["DO\n00\n2\u20181\n4\n2\n\u2014\n\u2014\u2014 x ex (\u2014ax )dx\n"]]], "page_581": [["block_0", [{"image_0": "581_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["To divide by a complex number, it is helpful to multiply numerator and denominator by the\ncomplex conjugate of the denominator, thereby restricting complex numbers to the numerator\nalone:\n"]], ["block_2", ["As a complex number is represented by a pair of numbers ((1,1)) it can also be mapped uniquely\nonto a point on a Cartesian coordinate system, with horizontal axis re\ufb02ecting the real part and\nthe vertical axis representing the imaginary part. Similarly, as the following figure illustrates,\na complex number can be viewed as a vector from the origin, with a length given by A and a\ndirection specified by the angle 6:\n"]], ["block_3", ["A.5\nComplex Numbers\n"]], ["block_4", ["A complex number 2 can always be written as the sum of two parts, referred to as the real part, a,\nand the imaginary part, ib:\nz:a+w\n(AiD\n"]], ["block_5", ["where a and b are real numbers and i: \\/-1. The rules for addition and subtraction of two\ncomplex numbers are straightforward:\n"]], ["block_6", ["Multiplication also follows directly, recalling that i2 \u20141:\n"]], ["block_7", ["Division is a little more complicated, and is helped by introduction of the complex conjugate of a\ncomplex number, 2*, which is obtained by replacing i with\n\u2014\u2014i:\n"]], ["block_8", ["The product of a complex number and its complex conjugate is always purely real:\n"]], ["block_9", ["Figure A.2\nIllustration of the mapping of complex numbers onto a Cartesian axis system.\n"]], ["block_10", ["Imaginary axis\n"]], ["block_11", ["572\nAppendix\n"]], ["block_12", ["22* 611611 + blbl\n(A55)\n"]], ["block_13", ["21 :l: 22 (611 +1b1):l:((12 +1332) (611 :l: 612) + i(b1 :l: b2)\n(A52)\n"]], ["block_14", ["Z :a1+ib1,\n2* 611 ibl\n(A.5.4)\n"]], ["block_15", ["22 \u2014ag+ib2 w_Clg +ib2 612\u2014i\n"]], ["block_16", ["2122 =(611 + 1190012 + ibz) ((11612) + i(blaz) + i(611592) (blbz)\n= (61102 191192) + i(611192 + blaz)\n(A-5-3)\n"]], ["block_17", ["21 __a1+ib1 _a1+ib1 612\u2014i\n"]], ["block_18", [{"image_1": "581_1.png", "coords": [50, 502, 224, 643], "fig_type": "figure"}]], ["block_19", ["_ \n, blag\u00a31l\n1\nA.5.6\na\ufb01+b\u00a7\n0134\u2014193\na\u00a7+bg\n(\n)\n"]], ["block_20", ["b 4r\n"]], ["block_21", ["\u20186\n"]], ["block_22", ["m\n"]], ["block_23", ["___._..._____._.\n"]], ["block_24", ["Real axis\n"]]], "page_582": [["block_0", [{"image_0": "582_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["The standard trigonometric relations apply, such that\n"]], ["block_2", ["Recall the series expansions of ex, cos x, and sin at given above, and consider the complex\nnumber e\u201d:\n"]], ["block_3", ["Thus any complex number can also be written 2 =Aei9. This particular form is extremely useful in\nvarious mathematical operations, for example taking powers and roots:\n"]], ["block_4", ["The product of z and its complex conjugate 2* is easily seen to be A2\n"]], ["block_5", ["Complex Numbers\n573\n"]], ["block_6", ["and therefore any complex number can be written as\n"]], ["block_7", ["In this way the product of a complex number and its conjugate is analogous to taking the dot\nproduct of a vector with itself; the result is a real number (scalar), equal to the length squared.\n"]], ["block_8", [{"image_1": "582_1.png", "coords": [40, 216, 268, 289], "fig_type": "molecule"}]], ["block_9", ["b\ntan 6 ;\n(A.5.7)\n"]], ["block_10", [{"image_2": "582_2.png", "coords": [46, 48, 119, 99], "fig_type": "molecule"}]], ["block_11", ["b A sin 6\n(A.5.8)\n"]], ["block_12", ["a A cos 6\n"]], ["block_13", ["z=Acost9+iAsin6\n(A.5.9)\n"]], ["block_14", ["ix__\n-\n1-\n2\n1-\n3\ne _1+u+i(1x) +3-iax) +...\n"]], ["block_15", ["2\u201d = (Ac-2w)\": A\u201d cine\n(A.5.11)\n"]], ["block_16", ["22* (x169) (Ac\u2014i9) A2 die\u2014i\") A2\n(A.5.12)\n"]], ["block_17", ["A \n612 + 102\n"]], ["block_18", ["21\u201451-31\n+4\u20141x +53:\n= cosx + isinx\n(A.5.10)\n"]], ["block_19", [{"image_3": "582_3.png", "coords": [60, 243, 251, 282], "fig_type": "molecule"}]]], "page_583": [["block_0", [{"image_0": "583_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["a-acetal\u2014hydrolyzable, 15\nB-acetal-hydrolyzable, 15\nAcid chlorides esterification, 61\nAcrylic, see Polyacrylonitrile\nAcrylonitrile, 88, 171, 173, 175\nAcrylonitrile\u2014methyl methacrylate copolymers\nchemical shift (from hexamethyldisiloxane)\nfor, 192\nActin, 18\nActivation energies for\ninitiator decomposition reaction, 86\npropagation, 92\ntermination, 88\nAddition polymers, 11\u201414\nAdipic acid esterification, 51\nAffine junction assumption, 402\nAlanine, l6\nAlbumin, 18\nAlternating copolymers, 172\nArnidation\nof acid chlorides, 64\u201465\nof amino acids, 64\u201466\nof diamine and diacid, 64\u201465\nAmorphous polyethylene, 291\nAmylopectin, 15\nAmylose, 15\nAnionic polymerization, 126\u2014137\nblock copolymers by, 129\u2014132\nend\u2014functional polymers by, 133\u2014135\nregular branched architectures,\n1354137\nAnisotropic liquid, 291\nArginine, l6\nArrhenius equation, 85\nAsparagine, 16\nAspartic acid, 16\nAtactic chain, 22\nAtactic polypropylene, 277, 291\nAtactic polystyrene, 511\n"]], ["block_2", ["Index\n"]], ["block_3", ["575\n"]], ["block_4", ["Band broadening, 368\nBead\u2014spring model (BSM), of linear viscoelasticity,\n432\u2014439, 461\nfreely draining, 439\ningredients of, 432\u2014434\nintrinsic functions, 438\nnormal mode concept, 434\u2014435\npredictions of, 434\u2014439\nreduced intrinsic dynamic moduli, 438\nRouse and Zimm versions of, 432\u2014433\nstress relaxation modulus, 437\nBenzoquinone [III] as inhibitor, 109\nBerthelot rule, 276\nBinodal curve, in phase behavior of polymer\nsolutions, 265, 268\u2014269\nBiological tissues, 41]\nBirefringent substances, 547\nBlock copolymers, 172\nby anionic polymerization, 129\u2014132\nmacromolecular surfactants, 130\nmicelles, 130\nnanostructured materials, 131\nself\u2014assembly, 130\nby sequential living anionic polymerization,\n131\u2014132\ntailored surfaces, 131\nthin \ufb01lms, 131\nBoltzmann superposition principle, of linear\nviscoelasticity, 430\u2014432\n"]], ["block_5", ["Atom transfer radical polymerization, 144\u2014145\nAvrami equation, 552\u2014557\nexponents in, 557\nAvrami parameters, crystallized fraction versus time\nfor, 556\nAxial diffusion, 368\nAxialites, 549\nAzeotropic polymerization, 170\nA20 compounds initiator, 80\u201481\n"]]], "page_584": [["block_0", [{"image_0": "584_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Calorimetry, glass transition temperature and,\n476\u2014477\nCapillary viscometers, 341\u2014345\nCarothers equation, 390\nCationic polymerization\naspects of, 138\u2014140\ncontrolled polymerization, 137\u2014142\nliving cationic polymerization, 140\u2014142\nCellulose, 15\nChain extension, in glassy polymers, 502\nChain-growth polymerization\nchain transfer, 105\u20141 10\ninitiation in, 78\u201486, 110\nmolecular weight distribution, 99\u2014104\nprincipal steps in chain-growth mechanism, 78\npropagation, 78, 90\u201496, 110\nradical lifetime, 96\u201499\nand step-growth polymerization, 77\u201479\ntermination, 78, 86\u201490, 110\nChain length distribution, kinetic analysis of,\n110, 119, 122\nChain molecules, see Polymers\nChain stiffness, in glassy polymers, 501\u2014504\nChain transfer\nin chain\u2014growth polymerization, 105\u2014109\nconstants, 106\u2014108\nevaluation of chain transfer constants, 106\u2014108\npolymer and, 108\u2014109\nreactions, 105\u2014106\nsuppressing polymerization, 109\nCharacteristic ratio, in polymer conformations,\n223\u2014225\nChemical shifts, 188\u2014189, 192\n"]], ["block_2", ["387\u2014388\nBravais lattices, 513\nBrillouin scattering, 292\nBrittle fracture, 498\nBrittle-to\u2014ductile transition, in glassy polymers,\n498\u2014501\nBrownian motion, 347\nBulk crystallization\nAvrami equation, 552\u2014556\nexperimental aspects, 556\u2014561\nkinetics of crystalline polymer, 551\u2014562\n1,3-Butadiene, 22, 148, 173\n1-Butene, 171\nButylated hydroxytoluene (BHT), 109\n"]], ["block_3", ["Bond rotation, in polymer conformations, 217\u2014219\nBond rupture, in glassy polymers, 502\nBragg diffraction, 293\nBragg\u2019s law, 294, 515\nand scattering vector, 294\u2014296\nBranched polymers, 7\u20149\nBranching coefficient, in multifunctional monomers,\n"]], ["block_4", ["576\n"]], ["block_5", ["1,3-Chloroprene, 22\nChromophores, 188\u2014189\nClausius\u2014Mosotti equation, 291\nCoexistence curve, in phase behavior of polymer\nsolutions, 265\nCoherent scattering, 292\u2014294\nCohesive energy density (CED), 276\nCollagen, 18\nCollapsed polymer film, interfacial polymerization\nwith, 66\nColligative properties, 32\nCombination termination, 86\u201488, 102\u2014104\nComb polymers, 8\nComplex modulus, in linear viscoelasticity, \nComplex viscosity, in linear viscoelasticity, 429\u2014430\nCompliance concept, in linear viscoelasticity,\n421\u2014422, 460\nConcentric cylinder viscometers, 345\u2014346\nCondensation polymers, 11\u201414\ndistribution of product molecules in, 44\u201446\nreactivity and reaction rates, 46\u201449\nstep-growth polymers, 43\u201444\nConditional probability, 184\nCone and plate rheometers, 458\u2014459\nControlled polymerization, 118\nanionic polymerization, 126\u2014137\nblock copolymers, 129\u2014132\nbranched polymers, 135\u2014137\ncontrol radical polymerization, 142\u2014147\ndendrimers, 156\u2014160\nend-functional polymers, 133\u2014135\nPoisson distribution for ideal living\npolymerization, 118\u2014122\npolymerization equilibrium, 147\u2014150\nregular branched architecture, 135\u2014137\nring-Opening polymerization, 150\u2014156, 160\nControl radical polymerization, 142-147\natom transfer radical polymerization, 144\u2014145\nparticular realization of, 144\u2014147\nprinciples of, 142\u2014144\nreversible addition-fragmentation transfer\npolymerization, 146\u2014147\nstable free-radical polymerization, 145\nCopolymerization equation, 167, 172\nCopolymers, 9\nalternating copolymers, 172\nblock copolymers, 172\ncomposition, 166\u2014170, 185-193\neffects of r values, 171\u2014172\neffects of variations on sequence distributions,\n181\u2014182\nequa\ufb01on,167,l72\nfeedstock and, 168\u2014169\nmicrostructure of, 179\u2014193\npenultimate models, 183\u2014185\n"]], ["block_6", ["Index\n"]]], "page_585": [["block_0", [{"image_0": "585_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["random copolymers, 172\nrate laws, 166\u2014168\nreactivity ratios, 1704175, 185\u2014186\nrelation of reactivity ratios to chemical structure,\n173\u2014175\nresonance and reactivity, 175\u2014179\nsequence distributions, 180\u2014183, 1904193\nsingle-site catalysts, 208\u2014211\nspectroscopic techniques, 188\u2014190\nstereoregularity in, 193\u2014205\nterminal models, 183\u2014185\nZiegler\u2014Natta catalysts, 205\u2014207\nCoupling agent, 132\nCraze in polystyrene, transmission electron\nmicrograph of, 500\nCrazing, in glassy polymers, 498\u2014501\nCreep compliance, 425\u2014426\nCritical point, in phase behavior of polymer\nsolutions, 265, 270\u2014271\nCrystal classes, 513\u2014514\nCrystalline and amorphous polymer mixture,\nproperties of, 558\nCrystalline polymer\nbehavior of Gibbs free energy, 523\u2014524\nbulk crystallization, 551\u2014562\ncrystal classes in, 513\u2014514\nkinetics of, 536\u2014545, 551\u2014562\nlamellae, 526\u2014536, 562\nlevels of structure in, 512\nmelting temperature to molecular structure\nrelation, 521\u2014526\nmorphology, 545\u2014551\nnucleation and growth, 536\u2014545, 562\nsemicrystalline polymers, 545\u2014551\nstructure and characterization of unit cells in,\n513\u2014521\nstructure and melting of lamellae,\n526\u2014536, 562\nx-ray diffraction, 515\u2014517\nCrystallization\nbehavior of Gibbs free energy, 523\u2014524\nkinetics of, 512\u2014513\nthermodynamics of, 512, 521\u2014526\nCycle polymers, 8\nCysteine, 16\n"]], ["block_2", ["Index\n"]], ["block_3", ["Dacron, 14\nDebye function, 313\nDehydrohalogenated copolymers, ultraviolet-visible\nspectrum of, 1904191\nDendrimers, 8, 156\u2014160\nDensity \ufb02uctuations, 319\nDeoxyribonucleic acid (DNA), 18\nmolecules, 6\u20147\nin T2 bacteriophage, 236\n"]], ["block_4", ["Diacid and diol esterification, 61\u201462\n1,1\u2014Dialkyl alkenes, 138\nDiapers, 411\n1,3-Dienes, 138\nDiethyl fumarate, 173\nDifferential refractometer, 320\nDifferential scanning calorimetry (DSC), 474, 476\nDiffusion coefficient, for dilute polymer solutions,\n346\u2014354\nDilatometer, 474\nDilatometry, of glass transition temperature,\n474\u2014476\nDilute polymer solutions\ndiffusion coefficient, 346\u2014354\ndraining, 357\u2014360, 373\ndynamic light scattering, 354\u2014357\ndynamics of, 327\u2014372\nEinstein\u2019s law, 330\u2014334\nFick\u2019s laws, 348\u2014354\nfriction, 327\u2014330, 373\nfriction factor, 346\u2014354\nhydrodynamic interactions, 357\u2014360, 373\nhydrodynamic radius, 347\u2014348\nintrinsic viscosity, 334\u2014340\nmutual diffusion, 348\u2014354\nrelaxation time versus molecular\nweight in, 443\nshear thinning, 329\nsize exclusion chromatography, 360\u2014372\nStokes\u2019 law, 330\u2014334\ntracer diffusion, 347\u2014348\nviscosity, 327\u2014330, 341\u2014346, 373\nviscosity measurement, 3414346\nZimm model for, 439\u2014444, 461\nDisk-shaped crystals, growth of, 552\nDisproportionation termination, 86\u201488, 99\u2014102\n2,6-di-tert-butyl\u20144-methylphenol (butylated\nhydroxytoluene), 109\n"]], ["block_5", ["as inhibitor, 109\nDuctile material, 498\nDuctile polymer, necking in, 501\nDynamic light scattering, 292\nin dilute polymer solutions, 354\u2014357\nDynamic mechanical analysis, of glass transition\ntemperature, 478\u2014479\nDynamic response\nin linear viscoelasticity, 426\u2014430\nDynamic scaling, 440\n"]], ["block_6", ["Einstein\u2019s law\nand dilute polymer solutions, 330\u2014334\nsuspension of spheres, 332\u2014334\nviscous forces on rigid spheres, 331\u2014332\nElastically effective strands, 403, 408\nElastic deformation, 392\u2014394\n"]], ["block_7", ["577\n"]]], "page_586": [["block_0", [{"image_0": "586_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Elastomers, 381\nequation of state for, 396\nideal elastomers, 396\u2014397\nElectromagnetic radiation initiator, 80\u201481\nEnd strand pullout, in glassy polymers, 502\nEngineering stress, 403\nEngineering thermoplastics, 491\nEntanglement phenomenology\ndependence of Me on molecular structure,\n447\u2014450\neffect, 441\nin linear viscoelasticity, 444\u2014450\nrubbery plateau, 444\u2014447\nspacing, 446\nviscoelastic response of polymer melts, 445\nEntanglements, role in glassy polymers, 501\u2014504\nEnthalpy, 248\nFlory\u2014Huggins theory, 257\u2014258\nof mixing, 251\u2014254, 257\u2014258\nregular solution theory, 251\u2014254\nEntrOpy, 248\nBoltzmann definition of, 249\nlonger route mixing, 255\u2014257\nof mixing, 249\u2014251, 255\u2014257\nquick route mixing, 255\nregular solution theory, 249\u2014251\nEpoxy formulation, 391\nEquation of motion, 330\nEquilibrium compliance, 426\nEquipartition theorem, of statistical\nmechanics, 436\nEsterification of\nacid chlorides, 61\ndiacid and diol, 61\u201462\nhydroxycarboxylic acid, 60\u201462, 64\nEster interchange with alcohol and ester, 61\nEthylene, 138, 148\nhomopolymerization of, 177\nExchange energy, 252\nExtent of reaction (p), 45\u201447, 51, 55\u201457, 65, 68\n"]], ["block_2", ["F-actin (fibrillar), 18\nFeedstock and copolymers composition, 168\u2014169\nFibrinogen, 18\nPicks laws, 348\u2014354\n"]], ["block_3", ["Finemann\u2014Ross plot, 186\nFirst-order order phase transition, 469\u2014471\nFlory\u2014Huggins theory, 254\u2014258, 263\u2014264, 284, 334,\n412, 414, 472\n"]], ["block_4", ["Elasticity\nequation of state, 394\u2014396\nexperiments on real rubber, 397\u2014398\nideal elastomers, 396\u2014397\nthermodynamics of, 394\u2014398\nElastic scattering process, 292\n"]], ["block_5", ["578\n"]], ["block_6", ["G-actin (globular), 18\nGaussian chain, force to extend in rubber,\n400\u2014402\nGaussian network, modulus of rubber elasticity,\n403\u2014405\nGaussian strands, network of rubber elasticity,\n402\u2014403\nGelation, see Network formation of polymers\nGel filtration chromatography (GFC), 361\nGel fraction, 383\nGel permeation chromatography (GPC), 361\nGel point, 383\u2014385, 388\u2014389, 415\nGel-spun polyethylene, 551\nGels swelling, 410\u2014415\nbiological tissues, 411\ndiapers, 411\nhot melt adhesives, 410\u2014411\n"]], ["block_7", ["assumptions, 258\nenthalpy of mixing, 257\u2014258\nentrOpy of mixing, 255\u2014257\ninteraction parameter()() and, 276\nlonger route entropy of mixing, 255\u2014257\nosmotic pressure, 263\u2014264\nquick route entropy of mixing, 255\nForm factor in scattering, 304\u20143 11\ndefinition, 304\nfor isotropic solutions, 306\u2014307\nmathematical expression, 305\u2014306\n"]], ["block_8", ["as n\u2014rO, 307\nscattering regimes and, 312\u2014315\nFour-arm star polymers, 8\nFox equation, 494\n"]], ["block_9", ["220\u2014221\nFreely rotating chain\nangles 0 and (I) for, 221\nin polymer conformations, 221\u2014222\nFree-radical\ncombination reactions yielding i-mers and rate\nlaws, 103\nfate during initiation, 81\u201482\nhomopolymerization rates, 177\ninitiation reactions, 80\u201482\npolymerization, 110 (see also Chain-growth\npolymerization)\nFree volume\nchanges inferred from viscosity, 481\u2014483\ndescription of glass transition, 479\u2014485\nfractional free volume, 483\ntemperature dependence, 480\u2014481\nWilliams\u2014Landel\u2014Ferry equation, 483\u2014485\nFriction factor, in dilute polymer solutions, 327\u2014330,\n346\u2014354, 373\n"]], ["block_10", ["Free energy of mixing, 258, 268\nFreely jointed chain, in polymer conformations,\n"]], ["block_11", ["Index\n"]]], "page_587": [["block_0", [{"image_0": "587_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["modulus of swollen rubber, 411\nsoft contact lenses, 411\nswelling equilibrium, 412\u2014414\nGeometrical isomerism, 22\u201424\nGibbs and DiMarzio theory, 472\u2014474\nGibbs free energy, 248\nGlass; see also Glass transition; Glassy polymers\ndefinition, 465\u2014466\nand melting transitions, 466\u2014468\nGlass transition\n\ufb01rst\u2014order order phase transition, 469\u2014471\nfree volume description of, 479\u2014485\nGibbs and DiMarzio theory, 472\u2014474\nKauzmann temperature, 471\u2014472\nsecond\u2014order order phase transition, 469\u2014471\ntemperature (see Glass transition temperature)\nthermodynamic aspects of, 468\u2014477, 504\ntime-temperature superposition, 486\u2014491\nGlass transition temperature, 465\nby calorimetry, 476\u2014477\ndependence on chemical structure, 491\u2014492\ndependence on composition, 492\u2014495\ndependence on molecular weight, 492\u2014493\nby dilatometry, 474\u2014476\nfactors affecting, value of, 491\u2014495\nmeasurements, 474\u2014479, 504\nand melting temperature, 468\nby thermal analysis method, 476\nGlassy polymers\nbasic properties of, 496\u2014498\nbond rupture, 502\nbrittle\u2014to\u2014ductile transition, 498\u2014501\nchain extension, 502\nchain stiffness role, 501\u2014504\ncrazing, 498\u2014501\nend strand pullout, 502\nentanglements role, 501\u2014504\nlong chain pullout, 502\nmechanical properties of, 496\u2014504\nmechanical strength of materials,\n496\u2014504\nmolecular separation, 502\nresponses to increasing strain, 502\nshort chain pullout, 502\nyielding, 498\u2014501\nGlutamic acid, 16\nGlutamine, 16\nGlycine, l6\nGlycogen, 15\nGlycolic acid polymer, 69\nGood solvent\nexcluded volume and chains in, 280\u2014283\nswelling of coil in, 281\nGraft distribution patterns, 10\nGrafting, 8\n"]], ["block_2", ["Index\n"]], ["block_3", ["as function of molecular weight, 338\nand generation number for polyether\ndendrimers, 340\nMark-Houwink equation, 336\u2014340, 373\nIsobutylene, 148, 171\nIsoleucine, l6\nIsoprene, 148\n"]], ["block_4", ["1,3\u2014130prene, 22\nIsOprene, polymerization of, 127\n"]], ["block_5", ["Ideal copolymerization, 169\nIdeal elastomers, 396\u2014397\nIncoherent scattering, 292\u2014294\nInelastic scattering process, 292\nInfinite network polymers, 383\nIn\ufb02ection point, in phase behavior of polymer\nsolutions, 269\nInherent viscosity, 335\nInhibitors, 109\nInitiation reactions, 80\u201481\nkinetics of, 82\u201484\nphotochemical initiation, 84\u201485\ntemperature dependence rates of, 85\u201486\nInitiator decomposition reaction, activation energies\nfor, 86\nInitiator efficiency, de\ufb01nition, 82\nInstrument response function, 368\nInteraction parameter (x), 252, 275\u2014280\napproaches to, 278\u2014280\nfrom experiment, 276\u2014278\nFlory\u2014Huggins theory and, 276\nfrom regular solution theory,\n275\u2014276, 284\nIntrinsic viscosity\nof dilute polymer solutions, 334\u2014340\n"]], ["block_6", ["Hedrites, 549\nHindered rotation chain, in polymer conformations,\n222\u2014223\nHistidine, l6\nHoffmann and Lauritzen theory, 543\nHoffmann\u2014Weeks plot, 523\nHomopolymerization\nof ethylene, 177\nof styrene, 177\nHomopolymers, 7\u201411\nHot melt adhesives, 410\u2014411\nHydrogenated (or deuterated) polybutadienes,\ndiffusion of, 454\nHydroxycarboxylic acid esteri\ufb01cation,\n60\u201462, 64\nHyperbranched polymers, 8\n"]], ["block_7", ["Gyration radii\nin polymer conformations, 230\u2014234, 241\nfor polystyrenes in benzene, 283\n"]], ["block_8", ["579\n"]]], "page_588": [["block_0", [{"image_0": "588_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Kauzmann paradox, 472\nKauzmann temperature, 471\u2014472\nKeratin, 18\nKevlar, 511\nKinematics viscosity, 345\nKinetic analysis of\nchain lengths distribution of, 110\ntermination, 87\nKinetic chain length, 119, 122\nfor propagation, 94\u201496\nin propagation, 94\u201496\n"]], ["block_2", ["detector, 371\nexperimental aspects, 314\u2014320\ninstrumentation, 316\u2014317\nlight waves, 289\u2014291\nphotometer, 316\nby polymer solutions, 289\u2014320\npreparation of samples and solutions to study, 319\nrefractive index increment, 319\u2014320\ntechnique for polymers, 289\nLight waves, 2894291\nelectric \ufb01eld component, 290\nincident beam polarization effect for, 297\nLinear polymers, 7\u201411\nLinear viscoelasticity, 419\u2014459\nadditional relaxation processes repetation model,\n456\u2014458\n"]], ["block_3", ["Lactone polymerization, 61, 64\u201465, 67\nLamellae, 511\nbright\u2014field operation, 532\ncrystal growth, 539\u2014545\ndark-field operation, 532\ndependence of T,n on molecular weight,\n"]], ["block_4", ["530\u2014531\ndependence of Tm on thickness, 527\u2014530\nexperimental characterization of, 532\u2014536\ninversion problem, 532\nkinetics of nucleation and growth, 536\u2014545, 562\nreal\u2014space image, 532\nreciprocal\u2014space image, 532\nshadow casting, 532\nstructure, 532\u2014536\nstructure and melting of, 526\u2014536, 562\nsurface contributions to phase transitions,\n526\u2014527\nLeucine, l6\nLexan, 14\nLight scattering; see also Scattering\ncalibration, 317\u2014319\n"]], ["block_5", ["Isotactic\nchain, 22\npolypropylene, 511\nIsotropic liquid, 291\n"]], ["block_6", ["580\n"]], ["block_7", ["stress and strain, 421, 429\ntransient response, 423\u2014426\nviscoelastic properties reptation model, 453\u2014456\nviscosity, 421\u2014422, 460\nviscous and elastic responses, 422\u2014423\nZimm model for dilute solutions, 439\u2014444, 46]\nLinear viscoelastic limit, 422\nLiving cationic polymerization, 140\u2014142\nLiving polymerization, 117\ndefinition, 119\nkinetic scheme, 119\u2014122\nPoisson distribution for, 118\u2014126\nLondon (dispersion) interactions, 275\nLondon forces, 276\nLong chain pullout, in glassy polymers, 502\nL00p, in network polymers, 383\nLoss and storage moduli, 426\u2014429\nLucite, see Poly(methyl methacrylate)\nLysine, 16\n"]], ["block_8", ["Macromolecular surfactants, 130\nMaleic anhydride, 173\nMaltese cross optical pattern, 546\nMark\u2014Houwink parameters, for polymer-solvent\nsystems, 336, 338\u2014339\nMatrix-assisted laser desorption/ionization (MALDI)\nmass spectrometry, 31, 35\u201438\nMaxwell and Voigt elements, linear viscoelasticity of\ncomplex modulus, 429\u2014430\ncomplex viscosity, 429\u2014430\ncreep compliance, 425\u2014426\ndynamic response, 426\u2014430\nloss and storage moduli, 426\u2014429\n"]], ["block_9", ["stress relaxation, 423\u2014425\ntransient response, 423\u2014426\nMean-square fluctuations in polarizability, 299\nMechanical strength of materials, 497\nMelting temperature and glass transition\ntemperature, 468\n"]], ["block_10", ["basic concepts, 419\u2014423\nbead\u2014spring model, 432\u2014439, 461\nBoltzmann superposition principle, 430\u2014432\n"]], ["block_11", ["compliance, 421\u2014422, 460\ndiffusivity reptation model, 451\u2014453\ndynamic response, 426\u2014430\nentanglement phenomenology, 444\u2014450\nexperimental rheometry, 458\u2014460\nlongest relaxation time, 451\u2014453\nMaxwell and Voigt elements response, 423\u2014430\nmodulus, 421\u2014422, 460\nreduced intrinsic moduli, 442\nreptation model, 450\u2014458\n"]], ["block_12", ["Rouse and Zimm versions of, 432\u2014433\nRouse model for unentangled melts,\n439\u2014444, 461\n"]], ["block_13", ["Index\n"]]], "page_589": [["block_0", [{"image_0": "589_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Melt viscosity, dependence on molecular weight, 450\nMethacrylic acid, 171\nMethionine, 17\nMethyl acrylate, 88, 173, 175\nMethyl acrylate-vinyl chloride system, 187\nMethyl methacrylate, 88, 148, 171, 173\nacceleration of polymerization rate for, 89\nor-Methyl styrene, 148, 171\nMethyl vinyl ketone, 17], 173\nMicelles, 130\nMiktoarm star, 137\nMixing versus composition curves, free\nenergy of, 268\nMolecular separation, in glassy polymers, 502\nMolecular size distribution of, in step-growth\npolymerization, 55\u201460\nMolecular weight between entanglements, 446, 448\nMolecular weight distribution of\naddition polymer molecules, 99\u2014104\ndistribution of i-mers termination by\ncombination, 102\u2014104\ndistribution of i-mers termination by\ndisproportionation, 99\u2014102\nMole fraction, 25\nin step-growth polymerization, 56\u201458\nMolten poly(or-methyl styrene)\nstress relaxation modulus, 420\nviscosity versus molecular weight for, 420\nMoments of distribution, 28\nMomentum transfer vector, 295\nMonodisperse, 26\nMonomer-radical combinations\ncross-propagation constants values for, 175\nsubstituents effects on reactivity, 176\nMonosubstituted ethylene, stereoregularity, 193\u2014205\nMooney\u2014Rivlin equation, 409\u20144 10\nMultifunctional monomers, branching coefficient,\n387\u2014388\nMutual diffusion, dilute polymer solutions, 348\u2014354\nMylar, 14\n"]], ["block_2", ["Natural polymers, 13\u201418\nNatural rubber\nstress at constant length for, 397\nstress for cross-linked, 404\nstress versus elongation for, 399\ntemperature change during adiabatic\nextension of, 400\nNeoprene, 23\nNetwork formation of polymers\ndangling end, 383\ndefinition, 381\u2014383\nelements, 381\u2014383\ngel fraction, 383\njuncuon,381\n"]], ["block_3", ["Index\n"]], ["block_4", ["Packing length, 446\nPartial molar volume, 248\nPenultimate models of copolymers, 183\u2014185\nPerlon, l4\nPersistence length\nof \ufb02exible chains, 227\u2014228\nand semi\ufb02exible chains in polymers, 225\u2014230\nPhantom network, 408\nPhase behavior of polymer solutions, 264\u2014275\nbinodal curve, 265, 268\u2014269\ncoexistence curve, 265\ncritical point, 265, 270\u2014271\nin\ufb02ection point, 269\nphase diagram, 265\u2014268\nspinodal curve, 265, 269\u2014270\nstability limit (see Spinodal curve)\nstable, unstable, and metastable states, 269\u2014270\n"]], ["block_5", ["loop, 383\nby multifunctional monomers, 386-392, 415\nby random cross-linking, 381\u2014385\nsol fraction, 383\nNetwork materials of polymers, 381\nNewtonian \ufb02uids, 329, 422\nNewton\u2019s law of viscosity, 329\nNon-Gaussian force law, 406\u2014407\nNon-Newtonian \ufb02uid, 422\nNucleation and growth in crystalline polymer\ncritical nucleus, 538\ncrystal growth, 539\u2014545\nheterogeneous nucleation, 537\nhomogeneous nucleation, 537\nkinetics of, 536-545, 562\nprimary nucleation, 537\u2014539\nsaddle point, 539\nsecondary nucleation process, 540\nNumber-average molecular weight, 25\nNylon-6, see Poly(a-caprolactam)\nNylon-6,6, 14, 19, 511, 519\nNylon-6,10, 19\nNylon salt f\"\u2014 nylon equilibrium, 65\n"]], ["block_6", ["Ole\ufb01n and transition metal, orbital overlaps\nbetween, 206\nor-Ole\ufb01ns, 138\nOligomeric polystyrene, viscosity of, 482\nOrganic peroxides or hydroperoxides initiator, 80\u201481\nOsmotic pressure, 3]\neffect of concentration polymer solutions on,\n260\u2014261\nexperimental approach, 259\u2014261\nFlory\u2014Huggins theory, 263\u2014264\nnumber average molecular weight, 261\u2014263\noperational de\ufb01nition, 259\u2014260\nfor polystyrene in cyclohexane, 263\n"]], ["block_7", ["581\n"]]], "page_590": [["block_0", [{"image_0": "590_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Phase diagram\nfinding binodal curve, 265, 268\u2014269\nfinding critical point, 265, 2704271\nfinding spinodal curve, 265, 269\u2014270\nfrom Flory\u2014Huggins Theory, 271\u2014275\nlower critical solution temperature, 274\nfor polymer solutions, 265\u2014268\nfor polymer-solvent system, 273\nfor polystyrene in acetone, 275\nupper critical solution temperature, 274\nPhenylalanine, 17\nPhotochemical initiation, 84\u201485\nPhthalic acid polymerization, 61\nPlastic deformation, 498\nPlateau modulus, 446, 448\nPlate rheometers, 458\u2014459\nPlexiglass, see Poly(methyl methacrylate)\nPoiseuille equation, 341\u2014345\nPoisson distribution for ideal living polymerization,\n1184126\nrequirements for, 126\nPoisson\u2019s ratio, 394\nPoly( 1 \u2014acetoxyethy1ene), 19\nPolyacrylonitrile, 13, 336\nPolyactide polymerization, 153\u2014154\nPolyamides, 14, 44, 64\u201467\ninterchange reactions, 64\u201465\nlactam polymerization, 64\u201465, 67\nreactions for formulation of, 64\u201465\nstoichiometric balance between reactive\ngroups, 65\nPolyamides poly(hexamethy1ene\nsebacamide), 19\nPoly(amino acid), 44\nPoly(6-aminocaproic acid), 19\nPoly(y\u2014benzyl-L-glutamate), 226, 229\n1,4\u2014Polybutadiene, 224, 277, 291, 336, 446, 468, 484\nmechanism for vulcanization with sulfur, 385\nPoly(e\u2014caprolactam), 13, 33, 336\nPolycarbonate, 14\nof bisphenol, 468\nstress-strain behavior of, 501\nPoly( 1 \u2014chloroethylene), 19\nPoly(decamethylene azelamide), 524\nPoly(decamethylene azelate), 524\nPoly(decamethylene sebacamide), 524\nPoly(decamethylene sebacate), 524\nPolydienes, cross-linking of, 384\nPoly(dimethylsiloxane), 14, 154\u2014155, 224, 277, 291,\n336, 446, 468, 484\nPolydispersity index (PDI), 26\u201428\nPolyesters, 14, 43\nreactions for formulation, 61\nstep-growth polymerization, 6%\nsynthesis, 60\u201462, 64\n"]], ["block_2", ["582\n"]], ["block_3", ["Polyethylene, 13, 224, 277, 336, 446, 468, 511,\n518\u2014519, 524\u2014525, 530, 547\nbackbone bond conformations for, 218\ncrystallization rates for, 542\ncrystals electron micrographs of, 534\u2014536\ncrystal structure of, 519\ndendrite, 550\nhedrite, 550\nshish kebabs transmission electron micrograph of,\n551\ntrans and gauche arrangements of backbone\nbonds, 217\u2014218\nPoly(ethylene glycol), 13\nPoly(ethylene oxide), 13, 224, 277, 291, 336, 446,\n468, 511, 518, 524\u2014525, 547\ncrystal growth rate for low molecular-weight, 536\npolymerization, 152\u2014153\nPoly(ethylene terephthalate), 14, 19, 62, 336, 468,\n511, 559, 561\nPoly(hexamethylene adipamide), 14, 19, 336, 468,\n511, 518\u2014520\nPoly( 1 \u2014hydroxyethy1ene), 19\nPoly(12-hydroxystearic acid), 14\nPolyisobutylene, 13, 224, 277, 291, 336,\n446, 468, 484\nstress relaxation modulus for, 487\nPolyis0prene, 128, 277\n1,4-PolyiSOprene, 24, 224, 291, 336, 446, 468, 484,\n518, 522, 524\u2014525\nPoly(4,4\u2014isopropy]idenediphenylene carbonate)\nbisphenol, 14\nPoly(L-lactide), spherulites of, 546\nPolymer conformations\nbond rotation, 217\u2014219\ncharacteristic ratio, 223\u2014225\ncoils, 234\u2014235\nconformations, 217\u2014219\ndistribution about center of mass, 240\u2014241\nend-to-end distance and segment density\ndistributions for, 235\u2014241\nend-to-end distance distribution for, 239\u2014240\n"]], ["block_4", ["end\u2014to\u2014end distance for model chains in, 219\u2014223\nend-to\u2014end vector distribution for, 236\u2014239\nfreely jointed chain, 220\u2014221\nfreely rotating chain, 221\u2014222\nhindered rotation chain, 222\u2014223\nmodel chains in, 219\u2014223\nradius of gyration, 230\u2014234, 241\nrandom coil, 217\u2014218\nrods, 234\u2014235\nsegment density distributions for, 235\u2014241\nself-avoiding chains, 241\u2014242\nsemi\ufb02exible chains and persistence length,\n225\u2014230\nsize, 217\u2014219\n"]], ["block_5", ["Index\n"]]], "page_591": [["block_0", [{"image_0": "591_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["1,4\u2014Polymerizations, 23\n3,4\u2014Polymerizations, 23\nPolymerization with multifunctional monomers,\n386\u2014392, 415\nbranching coef\ufb01cient calculation, 387\u2014388\ngel point, 388\u2014389\nmolecular-weight averages, 389\u2014392\nreaction schemes for monomer mixtures,\n386\u2014387\nPolymers\naddition, 11\u201414\nalternating distribution patterns, 10\narchitectures, 8\nblock distribution patterns, 10, 129\u2014135\nbranched, 7\u20149, 129\u2014135\nchain-growth, 77\u20141 10\nchain transfer to, 108\u2014109\nchanges in specific volume with temperature, 467\ncharacteristic ratio, 224\nclasses of crystals, 513\u2014514\ncomb, 8\ncondensation, 11\u201414, 43\u201446\ncopolymers, 9, 165\u2014211\ncrystal, representation of, 528\ncrystalline, 511\u2014561\ncrystal structure, 511, 513\u20145 14\ncycle, 8\ndegree of polymerization, 3\u20144\ndendrimers, 8, 155\u2014156\nend group analysis, 32\u201434\nfour-arm star polymers, 8\ngeometrical isomerism, 22\u201424\ngraft distribution patterns, 10\ngrafting, 8\nhomopolymers, 7\u201411\nhyperbranched, 8\n"]], ["block_2", ["worm\u2014like chain, 226, 228\u2014230\nPolymer glass, see Glassy polymers\nPolymer radicals, initiation of, 83\u201484\nPolymer solutions\ndraining, 357\u2014360, 373\nhydrodynamic interactions, 357\u2014360, 373\nhydrodynamic radius, 347\u2014348\nhydrodynamic volume, 335\nPolymerization\nchain\u2014growth, 105\u2014109\nfree-radical, 110\nmacroinitiators for, 131\npreparation, 131\nsuppressing of, 109\n1,2\u2014Polymerizations, 23\n"]], ["block_3", ["Index\n"]], ["block_4", ["spheres, 234\u2014235\nstatistical segment length, 223\u2014225\ntrans and gauche arrangements of backbone\nbonds, 217\u2014218\n"]], ["block_5", ["stereo isomerism, 21\u201422\nstress-strain curves for, 497\nstructural isomerism, 20\u201424\nstructure and characterization of unit cells,\n513\u2014521\nsynthetic, 19\nterpolymer, 9\nunit cells, 511, 513\u2014521\nweight, 25\u201426\nworm-like chain, 226, 228\u2014230\nz\u2014average molecular weights, 25\u201426\nPolymer solutions\nenthalpy of mixing, 251\u2014254, 257\u2014258\nentropy of mixing, 249\u2014251, 255\u2014257\nexcluded volume and chains in good solvent,\n280\u2014283\nFlory\u2014Huggins theory, 254\u2014258, 263\u2014264, 284\ninteraction parameter (x), 252, 275\u2014280\nlight scattering (see Light scattering)\nosmotic pressure, 258\u2014264\nphase behavior of, 264\u2014275\nphase diagram, 265\u2014268\nregular solution theory, 249\u2014254\nthermodynamics, 247\u2014249, 283\nPolymer-solvent systems, Mark\u2014Houwink\nparameters for, 336\nPolymers viscoelasticity, see Linear viscoelasticity\nPoly(methy1 methacrylate), 13, 224, 277, 291, 336,\n446, 468, 484, 511\nstress-strain response for, 499\nPolymorphism, 519\nPoly(n\u2014hexyl isocyanate), 226, 229\nradius of gyration versus molecular weight\nfor, 234\nPoly(oxy\u20142,6-dimethyl- l ,4\u2014phenylene), 468\nPoly(oxyethylene oxyterphthaloyl), 19\n"]], ["block_6", ["linear, 7\u201411\nmolecular weight, 3\u20144\nmolecular weight measurement, 31\u201437\nmolecular weights and molecular weight\naverages, 24\u201431\nnatural, 13\u201418\nnomenclature, 18\u201419\nnumber weight z\u2014average molecular weights,\n25\u201426\npositional isomerism, 20\nradicals initiation of, 83\u201484, 142\u2014147\nrandom distribution patterns, 10\nrefractive indices of, 291\nsemi\ufb02exible chains and persistence length\nin, 225\u2014230\nsignificance, 1\u20142\nspatial extent, 5\u20147\nstatistical distribution patterns, 10\nstep\u2014growth, 43\u201471\n"]], ["block_7", ["583\n"]]], "page_592": [["block_0", [{"image_0": "592_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["Quasielastic scattering process, 292\n"]], ["block_2", ["Radiation pressure, 295\nRadical lifetime, in chain-growth polymerization,\n96\u201498\nRadius of gyration, in polymer conformations,\n230\u2014234, 241\nRaman scattering, 292\nRandom coil, in polymer conformations, 217\u2014218\nRandom copolymers, 172\nRandom cross\u2014linking, network formation by,\n381\u2014385\nRandom \ufb02uctuations, 293\nRate constant, 46\nRate laws for\nc0polymers, 166\u2014168\npropagation, 91\u201492\nRayleigh ratio, 302\nReactivity and reaction rates, in condensation\npolymers, 46\u201449\nReactivity ratios\nfor copolymers, 170\u2014175, 185\u2014186\neffects of r values, 171\u2014172\n"]], ["block_3", ["Polypeptides, 15\nPoly(p-phenylene), 225\u2014226, 229\nPoly(p-phenylene terephthalamide), 226, 229, 468,\n511, 523\nPolypropylene, 224, 336, 468, 511, 518, 521\nl3C-NMR assignments for, 204\nisospecific polymerization mechanism of, 210\nPolystyrene, 13, 127, 224, 277, 291, 336, 446,\n468, 484\nPoly(tetra\ufb02uoroethylene), 13, 277, 291, 468, 511,\n518\u2014520\nPoly(tetramethylenehexamethylene urethane), 14\nPoly(tetramethyl-p-phenylene siloxane), crystals\ngrowth rates of, 544\nPolyurethane, 14\nPoly(vinyl acetate), 19, 224, 277, 291, 336, 468, 484\nPoly(vinyl alcohol), 19, 336, 468, 511, 518, 520\nPoly(vinyl chloride), 13, 19, 277, 336, 468, 547\nPoly(vinyl \ufb02uoride), 511, 518, 520\nPositional isomerism, 20\nPrinciple of time-temperature superposition, in\nrheometry, 460\nProline, 17\nPropagation\nactivation energies for, 92\nin chain-growth polymerization, 78, 90\u201496\nkinetic chain length, 94\u201496\nrate laws for, 91-92\ntemperature dependence of rates of, 92\u201493\nPropylene, 148\nProtein molecules, structure in, 17\u201418\n"]], ["block_4", ["584\n"]], ["block_5", ["evaluation from composition data, 185\u2014186\nrelation to chemical structure, 173\u2014175\nReal rubber, elasticity experiments on, 397\u2014398\nReciprocal lattice vector, 295\nRecoverable compliance, 426\nRedox initiator systems, 80\u201481\nReduced viscosity, 335\nRefractive index increment, 300\nin light scattering, 319\u2014320\nRefractive indices, of polymers and solvents, 291\nRegular solution theory, 249\u2014254\nenthalpy of mixing, 251\u2014254\nentropy of mixing, 249\u2014251\ninteraction parameter (x), 275\u2014276, 284\nphase diagram, 266\npredictions of, 267\nRelative viscosity, 335\nRelaxation processes reptation model in linear\nviscoelasticity, 456\u2014458\nconstraint release process in, 456\ncontour length \ufb02uctuations in, 456\nReptation model in linear viscoelasticity, 450\u2014458\nadditional relaxation processes, 456\u2014458\nlongest relaxation time and diffusivity, 451\u2014453\nviscoelastic properties, 453\u2014456\nResonance\nand reactivity in copolymers, 175\u2014179\nstabilization energies, 177\nRetarders, 109\nReversible addition-fragmentation transfer\npolymerization, 146\u2014147\nRheometry\ncone and plate rheometers, 458\u2014459\nexperimental aspects, 458\u2014460\nplate rheometers, 458\u2014459\nprinciple of time-temperature superposition, 460\nshear sandwich rheometers, 458\u2014459\nRibonucleic acid (RNA), 18\nRI detector, 369\u2014370\nRigid spheres, viscous forces on, 331\u2014332\nRing-opening metathesis polymerization, 155\u2014156\nRing-opening polymerization, 12\ncontrolled polymerization, 150\u2014156, 160\npolyactide polymerization, 153\u2014154\npoly(dimethylsiloxane) polymerization, 154\u2014155\npoly(ethyleneoxide) polymerization, 152\u2014153\nring-opening metathesis polymerization, 155\u2014156\nRouse model for unentangled melts, 439\u2014444, 46]\nRubber, 383\nRubber elasticity\ndeveIOpments in, 406\u2014410\nexperiments on, 397\u2014398\nforce to extend Gaussian chain, 400\u2014402\nfront factor, 407\u2014408\nmodulus of Gaussian network, 403\u2014405\n"]], ["block_6", ["Index\n"]]], "page_593": [["block_0", [{"image_0": "593_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["for, 229\nSequence distributions\nin 00polymers, 180\u2014183, 190\u2014191\nexperimental determination, 190\u2014193\nSerine, 17\nShear rate, 421\nShear sandwich rheometers, 458\u2014459\nShear thinning, in dilute polymer solutions, 329\nShort chain pullout, in glassy polymers, 502\n"]], ["block_2", ["Scattering; see also Light scattering\nbasic concepts of, 291\u2014296, 320\nBragg\u2019s Law and, 29-3.1294\ncoherent scattering, 293\u2014294\nfrom dilute polymer solution, 298\u2014303\nform factor, 304\u2014311\nincoherent scattering, 293\u2014294\nby isolated small molecule, 296\u2014298\nfrom perfect crystal, 292\u2014293\nfrom perfectly homogeneous material, 292\nfrom randomly placed objects, 292\nregimes and particular form factors, 312\u2014315\nstructure factor, 305\nvector, 294\u2014296\nvector length, 295\nvolume, 299, 318\nZimm equation, 304, 307\u2014308\nZimm plot, 308\u2014311\nSchizophyllan, 229\nSchotten\u2014Baumann reaction, 61\nSchulz\u2014Zimm distribution, 30\u201431\nSecond\u2014order order phase transition, 469\u2014471\nSegment density distributions, for polymer\nconformations, 235\u2014241\nSelf-avoiding chains, in polymer conformations,\n241\u2014242\nSemicrystalline polymers\ndendritic structures, 549\ndiffusion limited aggregation, 549\nMaltese cross pattern, 546\u2014547\nmelt, 513\nmorphology of, 545\u2014551\nnonspherulitic morphologies, 548\u2014551\nspherulites, 545\u2014548\nSemiflexible chains and persistence length\nin polymers, 225\u2014230\nworm\u2014like chain, 226, 228\u2014230\nSemi\ufb02exible polymers, persistence lengths values\n"]], ["block_3", ["Mooney plot for, 410\nMooney-Rivlin equation, 409\u2014410\nnetwork defects, 408\u2014409\nnetwork of Gaussian strands, 402\u2014403\nnon\u2014Gaussian force law, 406\u2014407\nstatistical mechanical theory, 398\u2014405, 415\nRubbery plateau, 444\u2014447\n"]], ["block_4", ["Index\n"]], ["block_5", ["Simple shear, 421\nSingle-site catalysts, 208\u2014211\nmetallocene systems, 209\nSize exclusion chromatography (SEC), 31\nbasic separation process, 361\u2014365, 373\ncalibration curve for, 362\nchromatogram, 363\u2014364\ndetectors, 369\u2014372\nin dilute polymer solutions, 360\u2014372\nlight scattering detector, 371\nlimitations of calibration by standards, 367\u2014368\nfor molecular weight determination, 362\nRI detector, 369\u2014370\nseparation mechanism, 365\u2014367\ntwo calibration strategies, 367\u2014369\nuniversal calibration, 368\u2014369\nUV-vis detector, 370\u2014371\nviscometer, 372\nSoft contact lenses, 411\nSolubility parameter (8), 276\nfor common polymers and solvents, 277\nSolubility value, for common polymers\nand solvents, 277\nSolvents, refractive indices of, 291\nSpace group, 514\nSpandex,l4\nSpecific viscosity, 335\nSpheres, suspension of, 332\u2014334\nSpherulites, 545\u2014548\nleading edge of lathlike crystal within, 547\nof poly(L-lactide), 546\nof poly(l\u2014propylene oxide), 549\ny-polarized light incident on, 548\nSpinodal curve, in phase behavior of polymer\nsolutions, 265, 269\u2014270\nSpinodal decomposition, 269\nSplit cell prism based refractometer, 320\nSquare-law detectors, 290\nStable free-radical polymerization, 145\nStationary\u2014state radical concentration termination, 89\nStatistical mechanical theory, of rubber elasticity,\n398\u2014405, 415\nStatistical mechanics, equipartition theorem of, 436\nStatistical segment length, in polymer\nconformations, 223\u2014225\nStatistical thermodynamic concepts, 247\u2014249, 283\n"]], ["block_6", ["Steady\u2014state compliance, 426\nStep-growth polymerization, 44\ncatalyzed step\u2014growth reactions, 50\u201452, 56\n"]], ["block_7", ["and chain\u2014growth polymerization, 77\u201479, 110\ncondensation polymers, 43\u201449\ndistribution of molecular sizes, 55\u201460\nexperimental vs. theoritical data, 52\u201453\nkinetics of, 49\u201455\nmole fractions, 56\u201458\n"]], ["block_8", ["585\n"]]], "page_594": [["block_0", [{"image_0": "594_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["92\u201493\nTerminal control mechanism, 166\nTerminal models for copolymers, 183\u2014185\nTermination\nactivation energies for, 88\nin chain-growth polymerization, 78, 86\u201490\nby combination, 86\u201488, 102\u2014104\nby disproportionation, 86\u201488, 99\u2014102\n"]], ["block_2", ["Te\ufb02on, 511, 519; see also Poly(tetra\ufb02uoroethylene)\nTelechelic polymers, 133\nTemperature dependence of rates, of propagation,\n"]], ["block_3", ["polyamides, 64\u201467\npolyesters, 60\u201464\nstoichiometric imbalance, 67\u201470\nuncatalyzed step-growth reactions, 53\u201456\nweight fractions of species, 58\u201460\nStep\u2014growth polymers, classes of, 43\u201444\nStereoblock copolymer, 211\nStereocontrol, role of catalyst symmetry in, 209\nStereo isomerism, 21-22\nStereoregular homopolymers, 194\nStereoregularity\nassessment by NMR, 200\u2014205\ncharacteristics, 193\u2014196\nin copolymers, 193\u2014205\nsplitting of meso and racemic dyads, 197\u2014198\nstatistical description of, 196\u2014200\nsystem of notation, 196\u2014197\nStokes\u2014Einstein\u2014Debye equation, 440\nStokes\u2014Einstein relation, 348\nStokes\u2019 law\nand dilute polymer solutions, 330\u2014334\nsuspension of spheres, 332-334\nviscous forces on rigid spheres, 331\u2014332\nStrain, 421, 429\nStrain rate, 421, 429\nStrand, in network polymers, 381\nStress, 421, 429\nrelaxation, 423\u2014425\nrelaxation modulus, 419\u2014420\nStructural isomerism, 20\u201424\nStyrene, 88, 138, 148, 171, 173, 175, 177\nbutadiene rubber (SBR), 23\nhomopolymerization of, 177\nvinyl acetate system, 178\nSuspension of spheres, 332\u2014334\nSwelling equilibrium, 412\u2014414\nSwelling of gels, 410\u2014415\nmodulus of swollen rubber, 411\nswelling equilibrium, 412\u2014414\nSwollen rubber, modulus of, 411\nSyndiotactic chain, 22\nSyndiotactic polypropylene, 511\nSynthetic polymers, 19\n"]], ["block_4", ["586\n"]], ["block_5", ["Valine, 17\nVinyl, see Poly(vinyl chloride)\nVinyl acetate, 88, 148, 171, 173, 177\nVinyl chloride, 171, 173, 175\nVinylidene chloride, 171, 173\nVinyl monomer (CH2 2 CHX), polymerization of, 12\n"]], ["block_6", ["effect on conversion to polymer, 88\u201489\nkinetic analysis of, 87\nstationary\u2014state radical concentration, 89\nTerpolymer, 9\nTerylene, l4\nTetra\ufb02uoroethylene, 148\nThermal analysis, of glass transition\ntemperature, 476\nThermodynamic concepts, 247\u2014249, 283\nThermodynamic of glass transition, 468\u2014477, 504\n\ufb01rst-order order phase transition, 469\u2014471\nGibbs and DiMarzio theory, 472\u2014474\nKauzmann temperature, 471\u2014472\nsecond-order order phase transition, 469\u2014471\nThermodynamics\nof crystallization, 521\u2014526\nof elasticity, 394\u2014398\nThermoplastics, 491\nelastomers, 131\nmechanical properties for, 498\nThermosets, 381\nThompson\u2014Gibbs equation, 529\nThree-arm star polymers, atomic force microscopy\nimages of, 6\u20147\nThreonine, 17\nTime-temperature superposition (TTS)\nof dynamic moduli for polyisoprene, 490\nfor glass transition, 486\u2014491\nshift factor, 486\nTracer diffusion, 347\u2014348\nTransition metal and olefin, orbital overlaps\nbetween, 206\nTrapped entanglement, 408\nTrommsdorff effect, 88, 90\nTrue stress, 403\nTryptophan, 17\nTyrosine, 17\n"]], ["block_7", ["x\u2014ray diffraction, 515\u2014517\nUV-vis detector, 370\u2014371\n"]], ["block_8", ["Uncatalyzed step\u2014growth reactions, 53\u2014-56\nUnentangled melts, Rouse model for, 439\u2014444, 461\nUnit cells\nconstraints on, 514\nparameters, 518\nin polymers, 511, 513\u2014521\nspace group, 514\nstructure and characterization of, 513\u2014521\n"]], ["block_9", ["Index\n"]]], "page_595": [["block_0", [{"image_0": "595_0.png", "coords": [0, 0, 463, 716], "fig_type": 0}]], ["block_1", ["2-Vinyl pyridine, 88, 171\nViscoelastic liquids, 419\nViscoelastic properties reptation model, in linear\nviscoelasticity, 453\u2014456\nViscoelastic solid, 426\nViscometer, 372\nViscosity\n"]], ["block_2", ["Index\n"]], ["block_3", ["average molecular weight, 337\ndefinition, 421\u2014422, 460\nin dilute polymer solutions, 327\u2014330, 373\nmeasurement, 341\u2014346\nand molecular weight for molten poly(0t-methyl\nstyrene), 420\nViscosity measurement\ncapillary viscometers, 341\u2014345\nconcentric cylinder viscometers, 345\u2014346\nof dilute polymer solutions, 341\u2014346\nPoiseuille equation, 341\u2014345\nViscous forces on rigid spheres, 331\u2014332\nViscous heating, 330\nVogel\u2014Fulcher\u2014Tammann\u2014Hesse (VFTH) equation,\n483\u2014484\nrepresentative values of parameters for, 484\nVogel temperature, 483\nVulcanization, 384\u2014385\n"]], ["block_4", ["Weight\u2014average molecular weight, 26\nWeight fraction, 25\n"]], ["block_5", ["mechanism for, 385\n"]], ["block_6", ["z-average molecular weights, 25\u201426\nZero-order Bernoulli statistics, 199\nZero-order Markov statistics, 199\nZero shear viscosity, 329\nZiegler\u2014Natta catalysts, 205\u2014207\nbimetallic mechanism, 207\ncatalyst solubility, 206\ncrystal structure of solids, 206\nmonometallic mechanism, 207\nrate of polymerization, 206\ntacticity of products, 206\nZimm equation, for scattering, 304, 307\u2014308\nZimm model for dilute solutions, 439\u2014444, 461\nZimm plot, for scattering, 308\u2014311\n"]], ["block_7", ["Xanthan, 229\nX-ray diffraction\n"]], ["block_8", ["Weight fractions of species, in step-growth\npolymerization, 58\u201460\nWilliams\u2014Landel\u2014Ferry (WLF) equation, 483\u2014485\nrepresentative values of parameters for, 484\n"]], ["block_9", ["Yielding, in glassy polymers, 498\u2014501\nYield point, 498\nYoung\u2019s modulus, 394\n"]], ["block_10", ["structure of unit cells, 515\u2014517\ndiffraction patterns on area detector,\n"]], ["block_11", ["5 15\u20145 16\n"]], ["block_12", ["587\n"]]], "page_596": [["block_0", [{"image_0": "596_0.png", "coords": [0, 0, 468, 731], "fig_type": 0}]], ["block_1", ["\ufb01\u2014_-\u2014\u2014=.\n"]], ["block_2", ["mm\n"]], ["block_3", ["-.\n19:!\n"]], ["block_4", ["-:_-.r-\u2018-gr\u2018.\n"]], ["block_5", [",_\n"]], ["block_6", ["_\n"]], ["block_7", ["The book covers topics that appear prominently in current polymer\n"]], ["block_8", ["Chemistry\n"]], ["block_9", ["Written by well-established professors in the field, Polymer Chemistry,\nSecond Edition provides a well\u2014rounded and articulate examination of\npolymer properties at the molecular level. it focuses on fundamental\nprinciples based on underlying chemical structures, polymer synthesis,\n"]], ["block_10", ["science journals. It also provides mathematical tools as needed and\nfully derived problems for advanced calculations. This new edition\n"]], ["block_11", ["transition, crystallization, solution properties, thermodynamics, and light\nscattering.\n"]], ["block_12", ["Polymer Chemistry, Second Edition offers a logical presentation of\ntopics that can be scaled to meet the needs of introductory as well as\nmore advanced courses in chemistry, materials science, and chemical\n"]], ["block_13", ["chain conformations, while expanding and updating material on topics\nsuch as catalysis and synthesis, viscoelasticity, rubber elasticity, glass\n"]], ["block_14", ["characterization, and properties.\n"]], ["block_15", ["Consistent with the previous edition, the authors emphasize the logical\nprogression of concepts, rather than presenting just a catalog of facts.\n"]], ["block_16", ["engineering.\n"]], ["block_17", ["integrates \u2019new theories and experiments made possible by advances in\ninstrumentation. It adds new chapters on controlled polymerization and\n"]], ["block_18", ["Features:\nt\nCovers topics closest to what the authors use in their own courses\n-\nBuilds upon principles taught in undergraduate chemistry courses,\n"]], ["block_19", ["0\nContains mathematical tools and step-by-step derivations for\n"]], ["block_20", ["particularly organic and physical chemistry\n-\nIntegrates concepts from physics, biology, materials science,\n"]], ["block_21", ["0\nIncorporates new theories and experiments using the latest tools\n"]], ["block_22", ["and instrumentation\n"]], ["block_23", ["chemical engineering, and statistics as needed\n"]], ["block_24", ["example problems\n"]], ["block_25", ["60m Broken Sound Parkway. NW\nSuite 300. Bocu Raton. FL 33487\n"]], ["block_26", ["270 Madison Avenue\nNew York. NY I00\")\n"]], ["block_27", ["2 Park Square. Milton Park\nAbingdnn. Oxon OX l4 JRN. UK\nwww.crcpress.com\n"]], ["block_28", ["CRC Press\nTaylor & Francis Group\nan informa business\n-\n"]], ["block_29", ["www.lnylorandfrancisgroup.conr\n"]], ["block_30", [{"image_1": "596_1.png", "coords": [197, 543, 418, 716], "fig_type": "figure"}]], ["block_31", [{"image_2": "596_2.png", "coords": [209, 567, 359, 686], "fig_type": "figure"}]], ["block_32", [{"image_3": "596_3.png", "coords": [356, 5, 462, 221], "fig_type": "figure"}]], ["block_33", [{"image_4": "596_4.png", "coords": [361, 431, 467, 729], "fig_type": "figure"}]], ["block_34", [{"image_5": "596_5.png", "coords": [368, 620, 466, 708], "fig_type": "figure"}]], ["block_35", [{"image_6": "596_6.png", "coords": [371, 8, 461, 92], "fig_type": "figure"}]], ["block_36", [{"image_7": "596_7.png", "coords": [374, 223, 458, 302], "fig_type": "figure"}]], ["block_37", [{"image_8": "596_8.png", "coords": [375, 424, 458, 507], "fig_type": "figure"}]]]} \ No newline at end of file