text
stringlengths
1
1k
Substituting into the previous expression for ,
Substituting this expression and (b1) in (b3) gives
Derivation of discrete formula and work expression.
The change in internal energy of a system, measured from state 1 to state 2, is equal to
At the same time, the work done by the pressure–volume changes as a result from this process, is equal to
Since we require the process to be adiabatic, the following equation needs to be true
By the previous derivation,
Rearranging (c4) gives
Substituting this into (c2) gives
Integrating we obtain the expression for work,
Substituting in second term,
Using the ideal gas law and assuming a constant molar quantity (as often happens in practical cases),
By the continuous formula,
Substituting into the previous expression for ,
Substituting this expression and (c1) in (c3) gives
An adiabat is a curve of constant entropy in a diagram. Some properties of adiabats on a "P"–"V" diagram are indicated. These properties may be read from the classical behaviour of ideal gases, except in the region where "PV" becomes small (low temperature), where quantum effects become important.
The right diagram is a "P"–"V" diagram with a superposition of adiabats and isotherms:
The isotherms are the red curves and the adiabats are the black curves.
The adiabats are isentropic.
Volume is the horizontal axis and pressure is the vertical axis.
The term "adiabatic" () is an anglicization of the Greek term ἀδιάβατος "impassable" (used by Xenophon of rivers). It is used in the thermodynamic sense by Rankine (1866), and adopted by Maxwell in 1871 (explicitly attributing the term to Rankine).
The etymological origin corresponds here to an impossibility of transfer of energy as heat and of transfer of matter across the wall.
The Greek word ἀδιάβατος is formed from privative ἀ- ("not") and διαβατός, "passable", in turn deriving from διά ("through"), and βαῖνειν ("to walk, go, come").
Conceptual significance in thermodynamic theory.
The adiabatic process has been important for thermodynamics since its early days. It was important in the work of Joule because it provided a way of nearly directly relating quantities of heat and work.
Energy can enter or leave a thermodynamic system enclosed by walls that prevent mass transfer only as heat or work. Therefore, a quantity of work in such a system can be related almost directly to an equivalent quantity of heat in a cycle of two limbs. The first limb is an isochoric adiabatic work process increasing the system's internal energy; the second, an isochoric and workless heat transfer returning the system to its original state. Accordingly, Rankine measured quantity of heat in units of work, rather than as a calorimetric quantity. In 1854, Rankine used a quantity that he called "the thermodynamic function" that later was called entropy, and at that time he wrote also of the "curve of no transmission of heat", which he later called an adiabatic curve. Besides its two isothermal limbs, Carnot's cycle has two adiabatic limbs.
For the foundations of thermodynamics, the conceptual importance of this was emphasized by Bryan, by Carathéodory, and by Born. The reason is that calorimetry presupposes a type of temperature as already defined before the statement of the first law of thermodynamics, such as one based on empirical scales. Such a presupposition involves making the distinction between empirical temperature and absolute temperature. Rather, the definition of absolute thermodynamic temperature is best left till the second law is available as a conceptual basis.
In the eighteenth century, the law of conservation of energy was not yet fully formulated or established, and the nature of heat was debated. One approach to these problems was to regard heat, measured by calorimetry, as a primary substance that is conserved in quantity. By the middle of the nineteenth century, it was recognized as a form of energy, and the law of conservation of energy was thereby also recognized. The view that eventually established itself, and is currently regarded as right, is that the law of conservation of energy is a primary axiom, and that heat is to be analyzed as consequential. In this light, heat cannot be a component of the total energy of a single body because it is not a state variable but, rather, a variable that describes a transfer between two bodies. The adiabatic process is important because it is a logical ingredient of this current view.
Divergent usages of the word "adiabatic".
This present article is written from the viewpoint of macroscopic thermodynamics, and the word "adiabatic" is used in this article in the traditional way of thermodynamics, introduced by Rankine. It is pointed out in the present article that, for example, if a compression of a gas is rapid, then there is little time for heat transfer to occur, even when the gas is not adiabatically isolated by a definite wall. In this sense, a rapid compression of a gas is sometimes approximately or loosely said to be "adiabatic", though often far from isentropic, even when the gas is not adiabatically isolated by a definite wall.
Quantum mechanics and quantum statistical mechanics, however, use the word "adiabatic" in a very different sense, one that can at times seem almost opposite to the classical thermodynamic sense. In quantum theory, the word "adiabatic" can mean something perhaps near isentropic, or perhaps near quasi-static, but the usage of the word is very different between the two disciplines.
On the one hand, in quantum theory, if a perturbative element of compressive work is done almost infinitely slowly (that is to say quasi-statically), it is said to have been done "adiabatically". The idea is that the shapes of the eigenfunctions change slowly and continuously, so that no quantum jump is triggered, and the change is virtually reversible. While the occupation numbers are unchanged, nevertheless there is change in the energy levels of one-to-one corresponding, pre- and post-compression, eigenstates. Thus a perturbative element of work has been done without heat transfer and without introduction of random change within the system. For example, Max Born writes "Actually, it is usually the 'adiabatic' case with which we have to do: i.e. the limiting case where the external force (or the reaction of the parts of the system on each other) acts very slowly. In this case, to a very high approximation
that is, there is no probability for a transition, and the system is in the initial state after cessation of the perturbation. Such a slow perturbation is therefore reversible, as it is classically."
On the other hand, in quantum theory, if a perturbative element of compressive work is done rapidly, it changes the occupation numbers and energies of the eigenstates in proportion to the transition moment integral and in accordance with time-dependent perturbation theory, as well as perturbing the functional form of the eigenstates themselves. In that theory, such a rapid change is said not to be "adiabatic", and the contrary word "diabatic" is applied to it.
Recent research suggests that the power absorbed from the perturbation corresponds to the rate of these non-adiabatic transitions. This corresponds to the classical process of energy transfer in the form of heat, but with the relative time scales reversed in the quantum case. Quantum adiabatic processes occur over relatively long time scales, while classical adiabatic processes occur over relatively short time scales. It should also be noted that the concept of 'heat' (in reference to the quantity of thermal energy transferred) breaks down at the quantum level, and the specific form of energy (typically electromagnetic) must be considered instead. The small or negligible absorption of energy from the perturbation in a quantum adiabatic process provides a good justification for identifying it as the quantum analogue of adiabatic processes in classical thermodynamics, and for the reuse of the term.
Furthermore, in atmospheric thermodynamics, a diabatic process is one in which heat is exchanged.
In classical thermodynamics, such a rapid change would still be called adiabatic because the system is adiabatically isolated, and there is no transfer of energy as heat. The strong irreversibility of the change, due to viscosity or other entropy production, does not impinge on this classical usage.
Thus for a mass of gas, in macroscopic thermodynamics, words are so used that a compression is sometimes loosely or approximately said to be adiabatic if it is rapid enough to avoid significant heat transfer, even if the system is not adiabatically isolated. But in quantum statistical theory, a compression is not called adiabatic if it is rapid, even if the system is adiabatically isolated in the classical thermodynamic sense of the term. The words are used differently in the two disciplines, as stated just above.
In organic chemistry, an amide, also known as an organic amide or a carboxamide, is a compound with the general formula , where R, R', and R′′ represent any group, typically organyl groups or hydrogen atoms. The amide group is called a peptide bond when it is part of the main chain of a protein, and an isopeptide bond when it occurs in a side chain, as in asparagine and glutamine. It can be viewed as a derivative of a carboxylic acid () with the hydroxyl group () replaced by an amine group (); or, equivalently, an acyl (alkanoyl) group () joined to an amine group.
Common of amides are formamide (), acetamide (), benzamide (), and dimethylformamide (). Some uncommon examples of amides are "N"-chloroacetamide () and chloroformamide ().
Amides are qualified as primary, secondary, and tertiary according to whether the amine subgroup has the form , , or , where R and R' are groups other than hydrogen.
The core of amides is called the amide group (specifically, carboxamide group).
In the usual nomenclature, one adds the term "amide" to the stem of the parent acid's name. For instance, the amide derived from acetic acid is named acetamide (CH3CONH2). IUPAC recommends ethanamide, but this and related formal names are rarely encountered. When the amide is derived from a primary or secondary amine, the substituents on nitrogen are indicated first in the name. Thus, the amide formed from dimethylamine and acetic acid is "N","N"-dimethylacetamide (CH3CONMe2, where Me = CH3). Usually even this name is simplified to dimethylacetamide. Cyclic amides are called lactams; they are necessarily secondary or tertiary amides.
Amides are pervasive in nature and technology. Proteins and important plastics like Nylons, Aramid, Twaron, and Kevlar are polymers whose units are connected by amide groups (polyamides); these linkages are easily formed, confer structural rigidity, and resist hydrolysis. Amides include many other important biological compounds, as well as many drugs like paracetamol, penicillin and LSD. Low-molecular-weight amides, such as dimethylformamide, are common solvents.
Structure and bonding.
The lone pair of electrons on the nitrogen atom is delocalized into the carbonyl group, thus forming a partial double bond between nitrogen and carbon. In fact the O, C and N atoms have molecular orbitals occupied by delocalized electrons, forming a conjugated system. Consequently, the three bonds of the nitrogen in amides is not pyramidal (as in the amines) but planar. This planar restriction prevents rotations about the N linkage and thus has important consequences for the mechanical properties of bulk material of such molecules, and also for the configurational properties of macromolecules built by such bonds. The inability to rotate distinguishes amide groups from ester groups which allow rotation and thus create more flexible bulk material.
The C-C(O)NR2 core of amides is planar. The C=O distance is shorter than the C-N distance by almost 10%. The structure of an amide can be described also as a resonance between two alternative structures: neutral (A) and zwitterionic (B).
It is estimated that for acetamide, structure A makes a 62% contribution to the structure, while structure B makes a 28% contribution. (These figures do not sum to 100% because there are additional less-important resonance forms that are not depicted above). There is also a hydrogen bond present between the active groups hydrogen and nitrogen atoms. Resonance is largely prevented in the very strained quinuclidone.
In their IR spectra, amides exhibit a moderately intense "ν"CO band near 1650 cm−1. The energy of this band is about 60 cm-1 lower than for the "ν"CO of esters and ketones. This difference reflects the contribution of the zwitterionic resonance structure.
Compared to amines, amides are very weak bases. While the conjugate acid of an amine has a p"K"a of about 9.5, the conjugate acid of an amide has a p"K"a around −0.5. Therefore, amides do not have as clearly noticeable acid–base properties in water. This relative lack of basicity is explained by the withdrawing of electrons from the amine by the carbonyl. On the other hand, amides are much stronger bases than carboxylic acids, esters, aldehydes, and ketones (their conjugate acids' p"K"as are between −6 and −10).
The proton of a primary or secondary amide does not dissociate readily; its p"K"a is usually well above 15. Conversely, under extremely acidic conditions, the carbonyl oxygen can become protonated with a p"K"a of roughly −1. It is not only because of the positive charge on the nitrogen, but also because of the negative charge on the oxygen gained through resonance.
Hydrogen bonding and solubility.
Because of the greater electronegativity of oxygen, the carbonyl (C=O) is a stronger dipole than the N–C dipole. The presence of a C=O dipole and, to a lesser extent a N–C dipole, allows amides to act as H-bond acceptors. In primary and secondary amides, the presence of N–H dipoles allows amides to function as H-bond donors as well. Thus amides can participate in hydrogen bonding with water and other protic solvents; the oxygen atom can accept hydrogen bonds from water and the N–H hydrogen atoms can donate H-bonds. As a result of interactions such as these, the water solubility of amides is greater than that of corresponding hydrocarbons. These hydrogen bonds are also have an important role in the secondary structure of proteins.
The solubilities of amides and esters are roughly comparable. Typically amides are less soluble than comparable amines and carboxylic acids since these compounds can both donate and accept hydrogen bonds. Tertiary amides, with the important exception of "N","N"-dimethylformamide, exhibit low solubility in water.
Amides undergo many chemical reactions, although they are less reactive than esters. Amides hydrolyse in hot alkali as well as in strong acidic conditions. Acidic conditions yield the carboxylic acid and the ammonium ion while basic hydrolysis yield the carboxylate ion and ammonia. The protonation of the initially generated amine under acidic conditions and the deprotonation of the initially generated carboxylic acid under basic conditions render these processes non-catalytic and irreversible. Amides are also versatile precursors to many other functional groups. Electrophiles react with the carbonyl oxygen. This step often precedes hydrolysis, which is catalyzed by both Brønsted acids and Lewis acids. Enzymes, e.g. peptidases and artificial catalysts, are known to accelerate the hydrolysis reactions.
From carboxylic acids and related compounds.
Amides are usually prepared by coupling carboxylic acid with an amine. The direct reaction generally requires high temperatures to drive off the water:
Esters are far superior substrates relative to carboxylic acids
Further "activating" both acid chlorides (Schotten-Baumann reaction) and anhydrides (Lumière–Barbier method) react with amines to give amides:
Peptide synthesis use coupling agents such as HATU, HOBt, or PyBOP.
The hydrolysis of nitriles is conducted on an industrial scale to produce fatty amides. Laboratory procedures are also available.
Many specialized methods also yield amides. A variety of reagents, e.g. tris(2,2,2-trifluoroethyl) borate have been developed for specialized applications.
Animism (from Latin: "" meaning 'breath, spirit, life') is the belief that objects, places, and creatures all possess a distinct spiritual essence. Animism perceives all things—animals, plants, rocks, rivers, weather systems, human handiwork, and in some cases words—as being animated, having agency and free will. Animism is used in anthropology of religion as a term for the belief system of many Indigenous peoples in contrast to the relatively more recent development of organized religions. Animism is a metaphysical belief which focuses on the supernatural universe (beyond logical foundations and procedures): specifically, on the concept of the immaterial soul.
Although each culture has its own mythologies and rituals, animism is said to describe the most common, foundational thread of indigenous peoples' "spiritual" or "supernatural" perspectives. The animistic perspective is so widely held and inherent to most indigenous peoples that they often do not even have a word in their languages that corresponds to "animism" (or even "religion"). The term "animism" is an anthropological construct.
Largely due to such ethnolinguistic and cultural discrepancies, opinions differ on whether animism refers to an ancestral mode of experience common to indigenous peoples around the world or to a full-fledged religion in its own right. The currently accepted definition of animism was only developed in the late 19th century (1871) by Edward Tylor. It is "one of anthropology's earliest concepts, if not the first."
Animism encompasses beliefs that all material phenomena have agency, that there exists no categorical distinction between the spiritual and physical world, and that soul, spirit, or sentience exists not only in humans but also in other animals, plants, rocks, geographic features (such as mountains and rivers), and other entities of the natural environment. Examples include water sprites, vegetation deities, and tree spirits, among others. Animism may further attribute a life force to abstract concepts such as words, true names, or metaphors in mythology. Some members of the non-tribal world also consider themselves animists, such as author Daniel Quinn, sculptor Lawson Oyekan, and many contemporary Pagans.
English anthropologist Sir Edward Tylor initially wanted to describe the phenomenon as spiritualism, but he realized that it would cause confusion with the modern religion of spiritualism, which was then prevalent across Western nations. He adopted the term animism from the writings of German scientist Georg Ernst Stahl, who had developed the term "" in 1708 as a biological theory that souls formed the vital principle, and that the normal phenomena of life and the abnormal phenomena of disease could be traced to spiritual causes.
The origin of the word comes from the Latin word "anima", which means "life" or "soul."
The first known usage in English appeared in 1819.
"Old animism" definitions.
Earlier anthropological perspectives, which have since been termed the old animism, were concerned with knowledge on what is alive and what factors make something alive. The old animism assumed that animists were individuals who were unable to understand the difference between persons and things. Critics of the old animism have accused it of preserving "colonialist and dualistic worldviews and rhetoric."
Edward Tylor's definition.
The idea of animism was developed by anthropologist Sir Edward Tylor through his 1871 book "Primitive Culture", in which he defined it as "the general doctrine of souls and other spiritual beings in general." According to Tylor, animism often includes "an idea of pervading life and will in nature;" a belief that natural objects other than humans have souls. This formulation was little different from that proposed by Auguste Comte as "fetishism", but the terms now have distinct meanings.
For Tylor, animism represented the earliest form of religion, being situated within an evolutionary framework of religion that has developed in stages and which will ultimately lead to humanity rejecting religion altogether in favor of scientific rationality. Thus, for Tylor, animism was fundamentally seen as a mistake, a basic error from which all religions grew. He did not believe that animism was inherently illogical, but he suggested that it arose from early humans' dreams and visions and thus was a rational system. However, it was based on erroneous, unscientific observations about the nature of reality. Stringer notes that his reading of "Primitive Culture" led him to believe that Tylor was far more sympathetic in regard to "primitive" populations than many of his contemporaries and that Tylor expressed no belief that there was any difference between the intellectual capabilities of "savage" people and Westerners.
The idea that there had once been "one universal form of primitive religion" (whether labelled "animism", "totemism", or "shamanism") has been dismissed as "unsophisticated" and "erroneous" by archaeologist Timothy Insoll, who stated that "it removes complexity, a precondition of religion now, in "all" its variants."
Social evolutionist conceptions.
Tylor's definition of animism was part of a growing international debate on the nature of "primitive society" by lawyers, theologians, and philologists. The debate defined the field of research of a new science: "anthropology". By the end of the 19th century, an orthodoxy on "primitive society" had emerged, but few anthropologists still would accept that definition. The "19th-century armchair anthropologists" argued that "primitive society" (an evolutionary category) was ordered by kinship and divided into exogamous descent groups related by a series of marriage exchanges. Their religion was animism, the belief that natural species and objects had souls.
With the development of private property, the descent groups were displaced by the emergence of the territorial state. These rituals and beliefs eventually evolved over time into the vast array of "developed" religions. According to Tylor, as society became more scientifically advanced, fewer members of that society would believe in animism. However, any remnant ideologies of souls or spirits, to Tylor, represented "survivals" of the original animism of early humanity.
Confounding animism with totemism.
In 1869 (three years after Tylor proposed his definition of animism), Edinburgh lawyer John Ferguson McLennan, argued that the animistic thinking evident in fetishism gave rise to a religion he named "totemism". Primitive people believed, he argued, that they were descended from the same species as their totemic animal. Subsequent debate by the "armchair anthropologists" (including J. J. Bachofen, Émile Durkheim, and Sigmund Freud) remained focused on totemism rather than animism, with few directly challenging Tylor's definition. Anthropologists "have commonly avoided the issue of animism and even the term itself, rather than revisit this prevalent notion in light of their new and rich ethnographies."
According to anthropologist Tim Ingold, animism shares similarities with totemism but differs in its focus on individual spirit beings which help to perpetuate life, whereas totemism more typically holds that there is a primary source, such as the land itself or the ancestors, who provide the basis to life. Certain indigenous religious groups such as the Australian Aboriginals are more typically totemic in their worldview, whereas others like the Inuit are more typically animistic.
From his studies into child development, Jean Piaget suggested that children were born with an innate animist worldview in which they anthropomorphized inanimate objects and that it was only later that they grew out of this belief. Conversely, from her ethnographic research, Margaret Mead argued the opposite, believing that children were not born with an animist worldview but that they became acculturated to such beliefs as they were educated by their society.
Stewart Guthrie saw animism—or "attribution" as he preferred it—as an evolutionary strategy to aid survival. He argued that both humans and other animal species view inanimate objects as potentially alive as a means of being constantly on guard against potential threats. His suggested explanation, however, did not deal with the question of why such a belief became central to the religion. In 2000, Guthrie suggested that the "most widespread" concept of animism was that it was the "attribution of spirits to natural phenomena such as stones and trees."
"New animism" non-archaic definitions.
Many anthropologists ceased using the term "animism", deeming it to be too close to early anthropological theory and religious polemic. However, the term had also been claimed by religious groups—namely, Indigenous communities and nature worshippers—who felt that it aptly described their own beliefs, and who in some cases actively identified as "animists." It was thus readopted by various scholars, who began using the term in a different way, placing the focus on knowing how to behave toward other beings, some of whom are not human. As religious studies scholar Graham Harvey stated, while the "old animist" definition had been problematic, the term "animism" was nevertheless "of considerable value as a critical, academic term for a style of religious and cultural relating to the world."
Hallowell and the Ojibwe.
The "new animism" emerged largely from the publications of anthropologist Irving Hallowell, produced on the basis of his ethnographic research among the Ojibwe communities of Canada in the mid-20th century. For the Ojibwe encountered by Hallowell, "personhood" did not require human-likeness, but rather humans were perceived as being like other persons, who for instance included rock persons and bear persons. For the Ojibwe, these persons were each willful beings, who gained meaning and power through their interactions with others; through respectfully interacting with other persons, they themselves learned to "act as a person".
Hallowell's approach to the understanding of Ojibwe personhood differed strongly from prior anthropological concepts of animism. He emphasized the need to challenge the modernist, Western perspectives of what a person is, by entering into a dialogue with different worldwide views. Hallowell's approach influenced the work of anthropologist Nurit Bird-David, who produced a scholarly article reassessing the idea of animism in 1999. Seven comments from other academics were provided in the journal, debating Bird-David's ideas.
Postmodern anthropology.
More recently, postmodern anthropologists are increasingly engaging with the concept of animism. Modernism is characterized by a Cartesian subject-object dualism that divides the subjective from the objective, and culture from nature. In the modernist view, animism is the inverse of scientism, and hence, is deemed inherently invalid by some anthropologists. Drawing on the work of Bruno Latour, some anthropologists question modernist assumptions and theorize that all societies continue to "animate" the world around them. In contrast to Tylor's reasoning, however, this "animism" is considered to be more than just a remnant of primitive thought. More specifically, the "animism" of modernity is characterized by humanity's "professional subcultures", as in the ability to treat the world as a detached entity within a delimited sphere of activity.
Human beings continue to create personal relationships with elements of the aforementioned objective world, such as pets, cars, or teddy bears, which are recognized as subjects. As such, these entities are "approached as communicative subjects rather than the inert objects perceived by modernists." These approaches aim to avoid the modernist assumption that the environment consists of a physical world distinct from the world of humans, as well as the modernist conception of the person being composed dualistically of a body and a soul.
Nurit Bird-David argues that:
She explains that animism is a "relational epistemology" rather than a failure of primitive reasoning. That is, self-identity among animists is based on their relationships with others, rather than any distinctive features of the "self". Instead of focusing on the essentialized, modernist self (the "individual"), persons are viewed as bundles of social relationships ("dividuals"), some of which include "superpersons" (i.e. non-humans).
Stewart Guthrie expressed criticism of Bird-David's attitude towards animism, believing that it promulgated the view that "the world is in large measure whatever our local imagination makes it." This, he felt, would result in anthropology abandoning "the scientific project."
Like Bird-David, Tim Ingold argues that animists do not see themselves as separate from their environment:
Rane Willerslev extends the argument by noting that animists reject this Cartesian dualism and that the animist self identifies with the world, "feeling at once "within" and "apart" from it so that the two glide ceaselessly in and out of each other in a sealed circuit". The animist hunter is thus aware of himself as a human hunter, but, through mimicry, is able to assume the viewpoint, senses, and sensibilities of his prey, to be one with it. Shamanism, in this view, is an everyday attempt to influence spirits of ancestors and animals, by mirroring their behaviors, as the hunter does its prey.
Ethical and ecological understanding.
Cultural ecologist and philosopher David Abram proposed an ethical and ecological understanding of animism, grounded in the phenomenology of sensory experience. In his books "The Spell of the Sensuous" and "Becoming Animal," Abram suggests that material things are never entirely passive in our direct perceptual experience, holding rather that perceived things actively "solicit our attention" or "call our focus," coaxing the perceiving body into an ongoing participation with those things.
In the absence of intervening technologies, he suggests that sensory experience is inherently animistic in that it discloses a material field that is animate and self-organizing from the beginning. David Abram used contemporary cognitive and natural science, as well as the perspectival worldviews of diverse indigenous oral cultures, Abram proposed a richly pluralist and story-based cosmology in which matter is alive. He suggested that such a relational ontology is in close accord with humanity's spontaneous perceptual experience by drawing attention to the senses, and to the primacy of sensuous terrain, enjoining a more respectful and ethical relation to the more-than-human community of animals, plants, soils, mountains, waters, and weather-patterns that materially sustains humanity.
In contrast to a long-standing tendency in the Western social sciences, which commonly provide rational explanations of animistic experience, Abram develops an animistic account of reason itself. He holds that civilised reason is sustained only by intensely animistic participation between human beings and their own written signs. For instance, as soon as someone reads letters on a page or screen, they can "see what it says"—the letters speak as much as nature spoke to pre-literate peoples. Reading can usefully be understood as an intensely concentrated form of animism, one that effectively eclipses all of the other, older, more spontaneous forms of animistic participation in which humans were once engaged.