id
stringlengths
17
17
text
stringlengths
426
1.67k
keyphrases
listlengths
4
46
relations
listlengths
0
20
S0032386109010386
From a general point of view, polymerization techniques can be divided into two types of chemical reactions: step-growth polymerization and free radical polymerization. Step-growth polymerization is widely used for synthesis of polyesters, polyamide and epoxies while the synthesis of polyacrylics requires the use of free radical polymerization. These polymerization reactions can be performed either in bulk or in solution or in dispersed media. Heterophase polymerizations (i.e. emulsion, dispersion and miniemulsion polymerizations) present the advantage of easier removal of the resulting product from the reactor compared to bulk polymerization thanks to the low viscosity of the reaction medium. Polymerization in solution also induces lower viscosity but also lower reaction rates due to dilution of the reactants and higher cost and environmental impact due to the use of organic solvents. These problems are solved in the case of heterophase polymerizations where the reactants are confined inside droplets (no dilution effect) and water is used as medium. The use of surfactant molecules are usually needed for the stabilization of the monomer droplet and subsequent polymer particles in the water phase.
[ { "id": "T1", "start": 30, "end": 55, "type": 1, "type_": "Process" }, { "id": "T2", "start": 89, "end": 107, "type": 1, "type_": "Process" }, { "id": "T3", "start": 109, "end": 135, "type": 1, "type_": "Process" }, { "id": "T4", "start": 140, "end": 167, "type": 1, "type_": "Process" }, { "id": "T5", "start": 169, "end": 195, "type": 1, "type_": "Process" }, { "id": "T6", "start": 215, "end": 238, "type": 2, "type_": "Task" }, { "id": "T7", "start": 228, "end": 238, "type": 0, "type_": "Material" }, { "id": "T8", "start": 240, "end": 249, "type": 0, "type_": "Material" }, { "id": "T9", "start": 254, "end": 261, "type": 0, "type_": "Material" }, { "id": "T10", "start": 272, "end": 297, "type": 1, "type_": "Process" }, { "id": "T11", "start": 285, "end": 297, "type": 0, "type_": "Material" }, { "id": "T12", "start": 318, "end": 345, "type": 1, "type_": "Process" }, { "id": "T13", "start": 353, "end": 377, "type": 1, "type_": "Process" }, { "id": "T14", "start": 416, "end": 424, "type": 0, "type_": "Material" }, { "id": "T15", "start": 405, "end": 409, "type": 0, "type_": "Material" }, { "id": "T16", "start": 431, "end": 446, "type": 0, "type_": "Material" }, { "id": "T17", "start": 448, "end": 475, "type": 1, "type_": "Process" }, { "id": "T18", "start": 482, "end": 490, "type": 1, "type_": "Process" }, { "id": "T19", "start": 492, "end": 502, "type": 1, "type_": "Process" }, { "id": "T20", "start": 507, "end": 535, "type": 1, "type_": "Process" }, { "id": "T21", "start": 569, "end": 601, "type": 2, "type_": "Task" }, { "id": "T22", "start": 611, "end": 618, "type": 0, "type_": "Material" }, { "id": "T23", "start": 631, "end": 650, "type": 1, "type_": "Process" }, { "id": "T24", "start": 686, "end": 701, "type": 0, "type_": "Material" }, { "id": "T25", "start": 703, "end": 717, "type": 1, "type_": "Process" }, { "id": "T26", "start": 721, "end": 729, "type": 0, "type_": "Material" }, { "id": "T27", "start": 774, "end": 782, "type": 1, "type_": "Process" }, { "id": "T28", "start": 796, "end": 804, "type": 1, "type_": "Process" }, { "id": "T29", "start": 812, "end": 821, "type": 0, "type_": "Material" }, { "id": "T30", "start": 881, "end": 897, "type": 0, "type_": "Material" }, { "id": "T31", "start": 940, "end": 967, "type": 1, "type_": "Process" }, { "id": "T32", "start": 978, "end": 987, "type": 0, "type_": "Material" }, { "id": "T33", "start": 1008, "end": 1016, "type": 0, "type_": "Material" }, { "id": "T34", "start": 1021, "end": 1036, "type": 1, "type_": "Process" }, { "id": "T35", "start": 1042, "end": 1047, "type": 0, "type_": "Material" }, { "id": "T36", "start": 1078, "end": 1098, "type": 0, "type_": "Material" }, { "id": "T37", "start": 1126, "end": 1162, "type": 1, "type_": "Process" }, { "id": "T38", "start": 1147, "end": 1162, "type": 0, "type_": "Material" }, { "id": "T39", "start": 1178, "end": 1195, "type": 0, "type_": "Material" }, { "id": "T40", "start": 1203, "end": 1208, "type": 0, "type_": "Material" } ]
[ { "arg1": "T3", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T4", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T20", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T19", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T18", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" } ]
S0378381215300674
The next phase of our current study is to use the parameters obtained from pure-component systems in a transferable manner to represent the corresponding mixtures. Mixtures of n-alkanes and H2O have been studied previously with SAFT-γ SW [82]. In general it is well known that the extreme nature of the phase separation [150] makes it challenging to model mixtures of H2O with non-polar compounds. Because of the large differences in the dielectric constant of the two phases as well as in the dipole moment of H2O and the hydrophobic molecules, it especially difficult to obtain phase-independent unlike interaction parameters [112] and thus to model simultaneously the equilibrium phases. In previous work [82], emphasis was placed on obtaining an accurate description of the alkane-rich phases (both liquid and vapour), while small absolute (but not relative) deviations for the aqueous phase composition were achieved. The systems of interest in our current work are typically aqueous mixtures containing a high proportion of H2O, alkylamine, and CO2. Consequently, in order to provide an improved overall description of the fluid-phase equilibria at the conditions of interest, refinements have been made to the unlike parameters presented in the previous study [129] relating to the interactions between H2O and the alkyl groups, CH3 and CH2, namely ϵCH3,H2O, ϵCH2,H2O and λCH3,H2O, λCH2,H2O.
[ { "id": "T4", "start": 228, "end": 237, "type": 1, "type_": "Process" }, { "id": "T6", "start": 377, "end": 396, "type": 1, "type_": "Process" }, { "id": "T1", "start": 164, "end": 193, "type": 0, "type_": "Material" }, { "id": "T2", "start": 75, "end": 97, "type": 1, "type_": "Process" }, { "id": "T3", "start": 356, "end": 371, "type": 0, "type_": "Material" }, { "id": "T5", "start": 759, "end": 796, "type": 2, "type_": "Task" }, { "id": "T7", "start": 981, "end": 997, "type": 0, "type_": "Material" }, { "id": "T8", "start": 1310, "end": 1313, "type": 0, "type_": "Material" }, { "id": "T9", "start": 1322, "end": 1334, "type": 0, "type_": "Material" }, { "id": "T10", "start": 1336, "end": 1347, "type": 0, "type_": "Material" }, { "id": "T11", "start": 1289, "end": 1334, "type": 1, "type_": "Process" }, { "id": "T12", "start": 511, "end": 514, "type": 0, "type_": "Material" }, { "id": "T13", "start": 803, "end": 809, "type": 0, "type_": "Material" }, { "id": "T14", "start": 814, "end": 820, "type": 0, "type_": "Material" }, { "id": "T15", "start": 1030, "end": 1033, "type": 0, "type_": "Material" }, { "id": "T16", "start": 1035, "end": 1045, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1051, "end": 1054, "type": 0, "type_": "Material" }, { "id": "T18", "start": 1356, "end": 1374, "type": 1, "type_": "Process" }, { "id": "T19", "start": 1379, "end": 1397, "type": 1, "type_": "Process" }, { "id": "T20", "start": 523, "end": 544, "type": 0, "type_": "Material" }, { "id": "T21", "start": 1129, "end": 1151, "type": 1, "type_": "Process" }, { "id": "T22", "start": 1129, "end": 1134, "type": 0, "type_": "Material" } ]
[ { "arg1": "T10", "arg2": "T9", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T18", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T19", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" } ]
S0038092X14000942
Thermal performance of the smart window has been predicted under different simulated parameters namely, direct solar radiation intensity, ambient temperature, water inlet temperature, and water flow rate. Fig. 11 shows a sample of temperature distribution of all window components at a plane passing through the horizontal window segment based on simulation physical conditions listed on Table 3. Simulation data has been collected from all successive simulations. The effect of increasing direct solar radiation on both solar cells and water temperatures is shown in Fig. 12. Three different simulations were performed assuming direct solar radiation intensities of 400, 600, and 800W/m2 incident on the window’s front pane with set ambient temperature, water inlet temperature and water flow rate of respectively 273K, 283K and 0.01kg/s. Water temperature was found to increase by 5°C as it passed through the tube, carrying the solar cells, from left to right for the bottom most units at 800W/m2 of direct incident solar radiation.
[ { "id": "T1", "start": 0, "end": 39, "type": 2, "type_": "Task" }, { "id": "T2", "start": 75, "end": 95, "type": 1, "type_": "Process" }, { "id": "T3", "start": 104, "end": 136, "type": 1, "type_": "Process" }, { "id": "T4", "start": 138, "end": 157, "type": 1, "type_": "Process" }, { "id": "T5", "start": 159, "end": 182, "type": 1, "type_": "Process" }, { "id": "T6", "start": 188, "end": 203, "type": 1, "type_": "Process" }, { "id": "T7", "start": 263, "end": 280, "type": 0, "type_": "Material" }, { "id": "T8", "start": 347, "end": 377, "type": 0, "type_": "Material" }, { "id": "T9", "start": 397, "end": 412, "type": 0, "type_": "Material" }, { "id": "T10", "start": 593, "end": 604, "type": 1, "type_": "Process" }, { "id": "T11", "start": 452, "end": 463, "type": 1, "type_": "Process" }, { "id": "T12", "start": 479, "end": 555, "type": 1, "type_": "Process" }, { "id": "T13", "start": 636, "end": 651, "type": 1, "type_": "Process" }, { "id": "T14", "start": 705, "end": 724, "type": 0, "type_": "Material" }, { "id": "T15", "start": 912, "end": 916, "type": 0, "type_": "Material" }, { "id": "T16", "start": 931, "end": 942, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1019, "end": 1034, "type": 1, "type_": "Process" } ]
[ { "arg1": "T3", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T4", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T5", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T6", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" } ]
S0888613X16300767
However this is not just a useful depiction of an apposite well-supported statistical model. If we are prepared to allow that the process is driven by a CRG and that the MAP model that we have discovered is indeed generating the idle process, then identifying the disconnected components of the system allows us to immediately make assertions about the impact of various controls we might apply to this regulatory process – just as we can were we to believe the model was a causal extension of a BN. In the context of microarrays, the objective of clustering is to identify patterns among the data and decide which genes to focus on in further, more gene-specific, experiments. It is therefore necessary for the scientist to make such causal conjectures about the effect of controls available to her on the expressions reflecting the underlying regulatory process she studies. These conjectures can be universal or nuanced by evoking ideas of parsimony.
[ { "id": "T1", "start": 47, "end": 91, "type": 2, "type_": "Task" }, { "id": "T2", "start": 153, "end": 156, "type": 0, "type_": "Material" }, { "id": "T3", "start": 170, "end": 179, "type": 0, "type_": "Material" }, { "id": "T4", "start": 229, "end": 241, "type": 1, "type_": "Process" }, { "id": "T5", "start": 248, "end": 301, "type": 1, "type_": "Process" }, { "id": "T6", "start": 349, "end": 379, "type": 2, "type_": "Task" }, { "id": "T7", "start": 496, "end": 498, "type": 0, "type_": "Material" }, { "id": "T8", "start": 548, "end": 558, "type": 2, "type_": "Task" }, { "id": "T9", "start": 565, "end": 597, "type": 1, "type_": "Process" }, { "id": "T10", "start": 760, "end": 782, "type": 2, "type_": "Task" }, { "id": "T11", "start": 926, "end": 952, "type": 0, "type_": "Material" }, { "id": "T12", "start": 883, "end": 894, "type": 2, "type_": "Task" } ]
[]
S0045782513001448
Traditionally, the simulation of incompressible fluid flow by the SPH method has been through a weakly compressible SPH formulation (WCSPH). In this approach, the pressure is treated as a thermodynamic variable and is calculated using an artificial equation of state. The sound speed is set to be sufficiently high to limit density variations to within a small fraction of the actual fluid density. In practice, this high sound speed places a limitation on the maximum permissible time-step size through the Courant–Friedrichs–Lewy (CFL) constraint. A particular weakness relates to noise in the pressure field since a small perturbation in the local density will yield a large variation in the local pressure. This can make WCSPH formulations ineffective for accurate force and pressure prediction, although recent developments which create more uniform particle distributions have improved this [1,2]. A review of the SPH method can be found in [3] while a review of the classical WCSPH formulation applied to free-surface flows can be found in [4].
[ { "id": "T1", "start": 19, "end": 58, "type": 2, "type_": "Task" }, { "id": "T2", "start": 66, "end": 76, "type": 1, "type_": "Process" }, { "id": "T3", "start": 96, "end": 131, "type": 1, "type_": "Process" }, { "id": "T4", "start": 133, "end": 138, "type": 1, "type_": "Process" }, { "id": "T5", "start": 238, "end": 266, "type": 1, "type_": "Process" }, { "id": "T6", "start": 725, "end": 730, "type": 1, "type_": "Process" }, { "id": "T7", "start": 920, "end": 930, "type": 1, "type_": "Process" }, { "id": "T8", "start": 983, "end": 988, "type": 1, "type_": "Process" }, { "id": "T9", "start": 48, "end": 58, "type": 1, "type_": "Process" }, { "id": "T10", "start": 48, "end": 53, "type": 0, "type_": "Material" }, { "id": "T11", "start": 384, "end": 389, "type": 0, "type_": "Material" }, { "id": "T12", "start": 835, "end": 877, "type": 2, "type_": "Task" }, { "id": "T13", "start": 855, "end": 863, "type": 0, "type_": "Material" }, { "id": "T14", "start": 1012, "end": 1030, "type": 1, "type_": "Process" } ]
[ { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" } ]
S0167931713002438
There have been suggestions that electrons can be trapped in the bulk and at surfaces of silica [15] but new models of electron trapping centres started to appear only recently. It has been suggested by Bersuker et al., who used molecular models, that electrons can be trapped by Si–O bonds in a-SiO2 leading to their weakening and thus facilitating Si–O bond dissociation [16]. Further calculations by Camellone et al. have shown that electrons can spontaneously trap in non-defective continuum random network model of a-SiO2 [17]. Recent calculations have also demonstrated that the two dominant neutral paramagnetic defects at surfaces of a-SiO2, the non-bridging oxygen centre and the silicon dangling bond, are deep electron traps and can form the corresponding negatively charged defects [18]. However, these theoretical predictions have not yet been confirmed experimentally, emphasising the challenges for identifying defect centres.
[ { "id": "T1", "start": 229, "end": 245, "type": 1, "type_": "Process" }, { "id": "T2", "start": 252, "end": 261, "type": 0, "type_": "Material" }, { "id": "T4", "start": 33, "end": 42, "type": 0, "type_": "Material" }, { "id": "T5", "start": 50, "end": 95, "type": 1, "type_": "Process" }, { "id": "T6", "start": 119, "end": 144, "type": 0, "type_": "Material" }, { "id": "T7", "start": 436, "end": 445, "type": 0, "type_": "Material" }, { "id": "T8", "start": 450, "end": 468, "type": 1, "type_": "Process" }, { "id": "T9", "start": 914, "end": 940, "type": 2, "type_": "Task" }, { "id": "T10", "start": 89, "end": 95, "type": 0, "type_": "Material" }, { "id": "T11", "start": 65, "end": 69, "type": 0, "type_": "Material" }, { "id": "T3", "start": 280, "end": 290, "type": 0, "type_": "Material" }, { "id": "T12", "start": 294, "end": 308, "type": 0, "type_": "Material" }, { "id": "T13", "start": 318, "end": 327, "type": 1, "type_": "Process" }, { "id": "T14", "start": 269, "end": 276, "type": 1, "type_": "Process" }, { "id": "T15", "start": 350, "end": 372, "type": 1, "type_": "Process" }, { "id": "T16", "start": 350, "end": 359, "type": 0, "type_": "Material" }, { "id": "T17", "start": 472, "end": 516, "type": 1, "type_": "Process" }, { "id": "T18", "start": 520, "end": 526, "type": 0, "type_": "Material" }, { "id": "T19", "start": 642, "end": 648, "type": 0, "type_": "Material" }, { "id": "T20", "start": 598, "end": 626, "type": 1, "type_": "Process" }, { "id": "T21", "start": 654, "end": 680, "type": 0, "type_": "Material" }, { "id": "T22", "start": 689, "end": 710, "type": 0, "type_": "Material" }, { "id": "T23", "start": 716, "end": 735, "type": 0, "type_": "Material" }, { "id": "T24", "start": 767, "end": 793, "type": 0, "type_": "Material" }, { "id": "T25", "start": 630, "end": 648, "type": 0, "type_": "Material" } ]
[ { "arg1": "T21", "arg2": "T25", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T22", "arg2": "T25", "relation": 2, "relation_": "Hyponym-of" } ]
S0022311513011951
The displacement cascade is a rapid process (of order picoseconds). Further migration of vacancies and SIAs, mainly by diffusion, happens over a timescale of order nanoseconds [17]. This is still short compared to operating times, so is important to consider the equilibrium result of such processes: If the vacancies and SIAs were likely to find their Frenkel partner, recombine, and annihilate, then the metal should essentially return to its original structure; however, if defects instead formed large clusters of a single type this could result in formation of voids, dislocation loops or swelling, possibly weakening the material in the process. Defects can be trapped at grain boundaries or surface, so for an ODS particle to effect the diffusion, there concentration must be such that there are many such particles in each grain.
[ { "id": "T1", "start": 4, "end": 24, "type": 2, "type_": "Task" }, { "id": "T2", "start": 454, "end": 463, "type": 0, "type_": "Material" }, { "id": "T3", "start": 477, "end": 484, "type": 0, "type_": "Material" }, { "id": "T4", "start": 566, "end": 571, "type": 0, "type_": "Material" }, { "id": "T5", "start": 573, "end": 590, "type": 0, "type_": "Material" }, { "id": "T6", "start": 594, "end": 602, "type": 0, "type_": "Material" }, { "id": "T7", "start": 119, "end": 128, "type": 1, "type_": "Process" }, { "id": "T8", "start": 76, "end": 107, "type": 1, "type_": "Process" }, { "id": "T9", "start": 717, "end": 729, "type": 0, "type_": "Material" }, { "id": "T10", "start": 678, "end": 683, "type": 0, "type_": "Material" }, { "id": "T11", "start": 831, "end": 836, "type": 0, "type_": "Material" } ]
[ { "arg1": "T4", "arg2": "T3", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T5", "arg2": "T3", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T6", "arg2": "T3", "relation": 2, "relation_": "Hyponym-of" } ]
S0168365913008766
Immunopotentiators activate innate immunity directly (for example, cytokines) or through pattern-recognition receptors (PRRs, such as those for bacterial components). The Toll-like receptors (TLRs) are a family of PRRs that are an important link between innate and adaptive immunity. Some studies have shown that TLR ligands have adjuvant activity and enhance antigen-specific antibody and cell-mediated immune responses, especially when they are combined with delivery systems that promote their uptake and delivery into antigen-presenting cells [22–24]. For clinical studies, TLR9 is generally stimulated with synthetic oligodeoxynucleotides containing one or more unmethylated CpG dinucleotides. In humans, CpG has been used as an adjuvant for infectious disease vaccination [25,26] and in the development of cancer therapy [27]. In a mouse model, CpG has also been shown to induce T helper 1 (Th1) immune responses, which are characterized by the production of IFN-γ and the generation of IgG2a [28,29]. Moreover, a previous study had demonstrated that different liposomes with CpG ODN significantly increased Th1-biased cytokines and augmented cell mediated immune response [30].
[ { "id": "T1", "start": 0, "end": 18, "type": 0, "type_": "Material" }, { "id": "T2", "start": 19, "end": 43, "type": 1, "type_": "Process" }, { "id": "T3", "start": 89, "end": 118, "type": 1, "type_": "Process" }, { "id": "T4", "start": 120, "end": 124, "type": 1, "type_": "Process" }, { "id": "T5", "start": 171, "end": 190, "type": 1, "type_": "Process" }, { "id": "T6", "start": 192, "end": 196, "type": 1, "type_": "Process" }, { "id": "T7", "start": 214, "end": 218, "type": 1, "type_": "Process" }, { "id": "T8", "start": 313, "end": 324, "type": 0, "type_": "Material" }, { "id": "T9", "start": 522, "end": 546, "type": 0, "type_": "Material" }, { "id": "T10", "start": 578, "end": 582, "type": 0, "type_": "Material" }, { "id": "T11", "start": 560, "end": 576, "type": 2, "type_": "Task" }, { "id": "T12", "start": 596, "end": 606, "type": 1, "type_": "Process" }, { "id": "T13", "start": 612, "end": 643, "type": 0, "type_": "Material" }, { "id": "T14", "start": 667, "end": 697, "type": 0, "type_": "Material" }, { "id": "T15", "start": 710, "end": 713, "type": 0, "type_": "Material" }, { "id": "T16", "start": 734, "end": 742, "type": 0, "type_": "Material" }, { "id": "T17", "start": 747, "end": 777, "type": 2, "type_": "Task" }, { "id": "T18", "start": 795, "end": 826, "type": 2, "type_": "Task" }, { "id": "T19", "start": 838, "end": 849, "type": 1, "type_": "Process" }, { "id": "T20", "start": 851, "end": 854, "type": 0, "type_": "Material" }, { "id": "T21", "start": 878, "end": 895, "type": 1, "type_": "Process" }, { "id": "T22", "start": 951, "end": 998, "type": 1, "type_": "Process" }, { "id": "T23", "start": 1067, "end": 1076, "type": 0, "type_": "Material" }, { "id": "T24", "start": 1114, "end": 1134, "type": 0, "type_": "Material" }, { "id": "T25", "start": 1139, "end": 1178, "type": 1, "type_": "Process" } ]
[ { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T5", "arg2": "T6", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T5", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T7", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T15", "arg2": "T16", "relation": 2, "relation_": "Hyponym-of" } ]
S0022311515303640
Ferritic and martensitic steels are candidate materials for use in nuclear reactors [1,2]. The transmutation-created inert gas, especially He, plays an important role in the microstructural evolution of these steels under neutron irradiation. In a previous paper [3] the mechanisms by which He in a perfect body-centred-cubic (bcc) Fe lattice, can agglomerate into bubbles was discussed. It was shown that small He interstitial clusters are highly mobile but become effectively pinned with the emission of Fe interstitials when the clusters contain 5 or more He atoms. Small bubbles up to around 1.5 nm in diameter can easily form at room temperature from such seed points but larger bubbles are more difficult to form by diffusion alone due to the induced strain in the bcc lattice which increases the energy barriers for diffusion towards the bubbles whilst reducing them in a direction away from the bubbles. Subsequent bubble enlargement can then only occur either through increased temperature or by radiation induced mechanisms which increase the number of vacancies in the bubble and reduce the lattice strain. Emission of interracial loops from such a bubble was not observed in molecular dynamics simulations.
[ { "id": "T1", "start": 271, "end": 372, "type": 2, "type_": "Task" }, { "id": "T2", "start": 0, "end": 31, "type": 0, "type_": "Material" }, { "id": "T3", "start": 174, "end": 241, "type": 2, "type_": "Task" }, { "id": "T4", "start": 412, "end": 436, "type": 0, "type_": "Material" }, { "id": "T5", "start": 1187, "end": 1217, "type": 1, "type_": "Process" }, { "id": "T6", "start": 575, "end": 582, "type": 0, "type_": "Material" }, { "id": "T7", "start": 684, "end": 691, "type": 0, "type_": "Material" }, { "id": "T8", "start": 823, "end": 832, "type": 0, "type_": "Material" }, { "id": "T9", "start": 506, "end": 522, "type": 0, "type_": "Material" }, { "id": "T10", "start": 559, "end": 567, "type": 0, "type_": "Material" }, { "id": "T11", "start": 722, "end": 731, "type": 1, "type_": "Process" }, { "id": "T12", "start": 749, "end": 763, "type": 1, "type_": "Process" }, { "id": "T13", "start": 771, "end": 782, "type": 0, "type_": "Material" }, { "id": "T14", "start": 845, "end": 852, "type": 0, "type_": "Material" }, { "id": "T15", "start": 903, "end": 910, "type": 0, "type_": "Material" }, { "id": "T16", "start": 1080, "end": 1086, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1102, "end": 1116, "type": 1, "type_": "Process" }, { "id": "T18", "start": 67, "end": 83, "type": 0, "type_": "Material" }, { "id": "T19", "start": 95, "end": 126, "type": 0, "type_": "Material" }, { "id": "T20", "start": 139, "end": 141, "type": 0, "type_": "Material" }, { "id": "T21", "start": 222, "end": 241, "type": 1, "type_": "Process" }, { "id": "T22", "start": 209, "end": 215, "type": 0, "type_": "Material" }, { "id": "T23", "start": 291, "end": 293, "type": 0, "type_": "Material" }, { "id": "T24", "start": 365, "end": 372, "type": 0, "type_": "Material" }, { "id": "T25", "start": 307, "end": 325, "type": 0, "type_": "Material" }, { "id": "T26", "start": 327, "end": 330, "type": 0, "type_": "Material" }, { "id": "T27", "start": 332, "end": 342, "type": 0, "type_": "Material" }, { "id": "T28", "start": 923, "end": 929, "type": 0, "type_": "Material" }, { "id": "T29", "start": 923, "end": 941, "type": 1, "type_": "Process" }, { "id": "T30", "start": 1005, "end": 1033, "type": 1, "type_": "Process" }, { "id": "T31", "start": 1091, "end": 1116, "type": 2, "type_": "Task" }, { "id": "T32", "start": 1160, "end": 1166, "type": 0, "type_": "Material" }, { "id": "T33", "start": 1118, "end": 1147, "type": 1, "type_": "Process" } ]
[ { "arg1": "T25", "arg2": "T26", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T26", "arg2": "T25", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T20", "arg2": "T19", "relation": 2, "relation_": "Hyponym-of" } ]
S0167931714000203
To restrict pollen tube growth to a single focal plane is an important subject to enable their accurate growth analysis under microscopic observation. In the conventional method to assay pollen tube growth, the pollen tubes grow in a disorderly manner on solid medium, rendering it impossible to observe their growth in detail. Here, we present a new method to assay pollen tube growth using poly-dimethylsiloxane microchannel device to isolate individual pollen tubes. The growth of the pollen tube is confined to the microchannel and to the same focal plane, allowing accurate microscopic observations. This methodology has the potential for analyses of pollen tube growth in microfluidic environments in response to chemical products and signaling molecules, which paves the way for various experiments on plant reproduction.
[ { "id": "T1", "start": 3, "end": 54, "type": 2, "type_": "Task" }, { "id": "T2", "start": 104, "end": 149, "type": 1, "type_": "Process" }, { "id": "T3", "start": 181, "end": 206, "type": 1, "type_": "Process" }, { "id": "T5", "start": 296, "end": 327, "type": 1, "type_": "Process" }, { "id": "T7", "start": 570, "end": 603, "type": 2, "type_": "Task" }, { "id": "T10", "start": 12, "end": 30, "type": 1, "type_": "Process" }, { "id": "T11", "start": 211, "end": 228, "type": 1, "type_": "Process" }, { "id": "T4", "start": 181, "end": 198, "type": 0, "type_": "Material" }, { "id": "T6", "start": 361, "end": 385, "type": 1, "type_": "Process" }, { "id": "T12", "start": 361, "end": 378, "type": 0, "type_": "Material" }, { "id": "T13", "start": 392, "end": 433, "type": 0, "type_": "Material" }, { "id": "T14", "start": 437, "end": 468, "type": 1, "type_": "Process" }, { "id": "T15", "start": 456, "end": 468, "type": 0, "type_": "Material" }, { "id": "T9", "start": 347, "end": 385, "type": 2, "type_": "Task" }, { "id": "T16", "start": 488, "end": 499, "type": 0, "type_": "Material" }, { "id": "T8", "start": 656, "end": 674, "type": 1, "type_": "Process" }, { "id": "T17", "start": 656, "end": 667, "type": 0, "type_": "Material" }, { "id": "T18", "start": 741, "end": 760, "type": 0, "type_": "Material" }, { "id": "T19", "start": 719, "end": 736, "type": 0, "type_": "Material" }, { "id": "T20", "start": 809, "end": 827, "type": 1, "type_": "Process" }, { "id": "T21", "start": 644, "end": 674, "type": 2, "type_": "Task" }, { "id": "T22", "start": 794, "end": 827, "type": 2, "type_": "Task" } ]
[]
S0301932215002037
There is also a lack of agreement as to what constitutes churn flow. It is fairly certainly a gas continuous flow. There is growing agreement that there are huge waves present and some of the liquid is carried as drops. Sekoguchi and Mori (1997) and Sawai et al. (2004) using measurements from their multiple probes (92 over an axial length of 2.325 m) obtained time/axial position/void fraction information. From this they were able to identify huge wave from amongst disturbance waves and slugs. They classified individual structures as huge waves from their size together with the fact that their velocities depended significantly on the corresponding axial length. This was in contrast to disturbance waves where the velocity of individual waves only increased slightly with the axial extent of these waves. They also found that the frequency of huge waves first increased and then decrease with increasing gas superficial velocity. Similarly, their velocities were found to deviate from the line for slug flow velocities and pass through a maximum and then a minimum.
[ { "id": "T1", "start": 57, "end": 67, "type": 1, "type_": "Process" }, { "id": "T2", "start": 309, "end": 315, "type": 0, "type_": "Material" }, { "id": "T3", "start": 362, "end": 407, "type": 1, "type_": "Process" }, { "id": "T4", "start": 437, "end": 455, "type": 1, "type_": "Process" }, { "id": "T5", "start": 469, "end": 486, "type": 0, "type_": "Material" }, { "id": "T8", "start": 693, "end": 710, "type": 0, "type_": "Material" }, { "id": "T9", "start": 721, "end": 749, "type": 1, "type_": "Process" }, { "id": "T10", "start": 837, "end": 860, "type": 1, "type_": "Process" }, { "id": "T11", "start": 911, "end": 935, "type": 1, "type_": "Process" }, { "id": "T12", "start": 1005, "end": 1025, "type": 1, "type_": "Process" }, { "id": "T13", "start": 94, "end": 113, "type": 1, "type_": "Process" }, { "id": "T14", "start": 94, "end": 97, "type": 0, "type_": "Material" }, { "id": "T15", "start": 57, "end": 62, "type": 0, "type_": "Material" }, { "id": "T16", "start": 157, "end": 167, "type": 0, "type_": "Material" }, { "id": "T17", "start": 192, "end": 198, "type": 0, "type_": "Material" }, { "id": "T18", "start": 446, "end": 455, "type": 0, "type_": "Material" }, { "id": "T19", "start": 491, "end": 496, "type": 0, "type_": "Material" }, { "id": "T20", "start": 539, "end": 549, "type": 0, "type_": "Material" }, { "id": "T6", "start": 744, "end": 749, "type": 0, "type_": "Material" }, { "id": "T7", "start": 805, "end": 810, "type": 0, "type_": "Material" }, { "id": "T21", "start": 850, "end": 860, "type": 0, "type_": "Material" }, { "id": "T22", "start": 911, "end": 914, "type": 0, "type_": "Material" }, { "id": "T23", "start": 1005, "end": 1009, "type": 0, "type_": "Material" } ]
[]
S0032386109004996
SPM, and AFM in particular, has been widely applied to questions in polymer crystallization. The technique has several strengths that make it ideally suited for such studies. It is a high resolution technology, routinely resolving sub 10nm features [13,14], and hence allowing the fundamental length scale of the polymer lamellar crystal, its thickness, to be observed. AFM requires no staining or metal coating of the sample, so sample preparation is relatively straightforward. Also, it is non-destructive under many circumstances. This allows images to be obtained while a process such as crystal growth or melting is occurring, giving time-resolved data at lamellar or sub-lamellar resolution [15–18]. It is this final feature that provides many of the most exciting possibilities of AFM for studying polymer crystallization, as it is now possible to watch crystal growth, crystal melting, and re-organisations within crystals at the lamellar scale, seeing how structure evolves and local conditions influence kinetics. AFM has a wide range of different measuring modes, and, with the ever increasing number of functional semicrystalline polymers available (e.g. [19]), the breadth of experiments that can be carried out with a single machine is also one of the techniques attractions.
[ { "id": "T1", "start": 0, "end": 3, "type": 1, "type_": "Process" }, { "id": "T2", "start": 9, "end": 12, "type": 1, "type_": "Process" }, { "id": "T3", "start": 68, "end": 91, "type": 1, "type_": "Process" }, { "id": "T4", "start": 68, "end": 75, "type": 0, "type_": "Material" }, { "id": "T5", "start": 183, "end": 209, "type": 1, "type_": "Process" }, { "id": "T6", "start": 221, "end": 248, "type": 1, "type_": "Process" }, { "id": "T7", "start": 313, "end": 337, "type": 0, "type_": "Material" }, { "id": "T8", "start": 370, "end": 373, "type": 1, "type_": "Process" }, { "id": "T9", "start": 386, "end": 394, "type": 1, "type_": "Process" }, { "id": "T10", "start": 398, "end": 411, "type": 1, "type_": "Process" }, { "id": "T11", "start": 430, "end": 448, "type": 1, "type_": "Process" }, { "id": "T12", "start": 592, "end": 606, "type": 1, "type_": "Process" }, { "id": "T13", "start": 610, "end": 617, "type": 1, "type_": "Process" }, { "id": "T14", "start": 592, "end": 599, "type": 0, "type_": "Material" }, { "id": "T15", "start": 639, "end": 657, "type": 0, "type_": "Material" }, { "id": "T16", "start": 788, "end": 791, "type": 1, "type_": "Process" }, { "id": "T17", "start": 796, "end": 828, "type": 2, "type_": "Task" }, { "id": "T18", "start": 805, "end": 828, "type": 1, "type_": "Process" }, { "id": "T19", "start": 861, "end": 875, "type": 1, "type_": "Process" }, { "id": "T20", "start": 877, "end": 892, "type": 1, "type_": "Process" }, { "id": "T21", "start": 898, "end": 952, "type": 1, "type_": "Process" }, { "id": "T22", "start": 922, "end": 930, "type": 0, "type_": "Material" }, { "id": "T23", "start": 861, "end": 868, "type": 0, "type_": "Material" }, { "id": "T24", "start": 877, "end": 884, "type": 0, "type_": "Material" }, { "id": "T25", "start": 1024, "end": 1027, "type": 0, "type_": "Material" }, { "id": "T26", "start": 1126, "end": 1150, "type": 0, "type_": "Material" }, { "id": "T27", "start": 55, "end": 91, "type": 2, "type_": "Task" } ]
[]
S2352179114200056
Power and particle exhaust are crucial for the viability of any future fusion power plant concept. Heat in fusion reactors must be extracted through a wall and cannot be exhausted volumetrically, which limits the allowed power density in fusion reactors [1] and is a severe technical challenge in itself [2]. In addition, structural material changes resulting from neutron irradiation cause degradation in the heat exhaust capabilities of existing designs [3] and static surfaces can suffer severely from erosion due to impinging plasma particles [4,5]. It is concluded that conventional concepts and materials for plasma facing components (PFCs) reach their limits in terms of material lifetime and power exhaust at approximately 20MW/m2, which is presumably dramatically reduced to <10MW/m2 due to neutron damage in a D-T reactor [6] or even only half that value [7].
[ { "id": "T1", "start": 107, "end": 122, "type": 1, "type_": "Process" }, { "id": "T2", "start": 238, "end": 253, "type": 1, "type_": "Process" }, { "id": "T3", "start": 464, "end": 479, "type": 1, "type_": "Process" }, { "id": "T4", "start": 0, "end": 26, "type": 0, "type_": "Material" }, { "id": "T5", "start": 365, "end": 384, "type": 1, "type_": "Process" }, { "id": "T6", "start": 615, "end": 639, "type": 2, "type_": "Task" }, { "id": "T7", "start": 641, "end": 645, "type": 2, "type_": "Task" }, { "id": "T8", "start": 820, "end": 831, "type": 0, "type_": "Material" }, { "id": "T9", "start": 322, "end": 349, "type": 1, "type_": "Process" }, { "id": "T10", "start": 322, "end": 341, "type": 0, "type_": "Material" }, { "id": "T11", "start": 520, "end": 546, "type": 0, "type_": "Material" }, { "id": "T12", "start": 800, "end": 807, "type": 0, "type_": "Material" }, { "id": "T13", "start": 800, "end": 814, "type": 1, "type_": "Process" }, { "id": "T14", "start": 71, "end": 89, "type": 0, "type_": "Material" }, { "id": "T15", "start": 267, "end": 293, "type": 2, "type_": "Task" } ]
[ { "arg1": "T7", "arg2": "T6", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T7", "relation": 1, "relation_": "Synonym-of" } ]
S0370269304009013
Several methods based on dynamical assumptions were suggested for determination of the P-parity of the Θ+ [13]. According to a general theorem [14], in order to determine the parity of one particle in a binary reaction one has to know polarizations at least of two fermions participating in this reaction. Model independent methods for determination of the P-parity of the Θ+ were suggested recently in Refs. [15,16] for pp-collision, and in Ref. [17] for photoproduction of the Θ+. The method of Refs. [15,16], based on the assumption that the spin of the Θ+ equals 12, suggests to measure the spin–spin correlation parameter in the reaction p→p→→Σ+Θ+ near the threshold. We generalize here this method for an arbitrary spin of the Θ+ and both isospins T=0 and T=1 of the NN channel of the NN→YΘ+ reaction. Furthermore, we consider a polarization transfer from a nucleon to the hyperon Y in this reaction. Our consideration is model independent, since it is based only on conservation of the P-parity, total angular momentum and isospin in the reaction and the generalized Pauli principle for nucleons.
[ { "id": "T1", "start": 66, "end": 95, "type": 2, "type_": "Task" }, { "id": "T2", "start": 127, "end": 142, "type": 1, "type_": "Process" }, { "id": "T3", "start": 161, "end": 181, "type": 2, "type_": "Task" }, { "id": "T4", "start": 203, "end": 218, "type": 1, "type_": "Process" }, { "id": "T5", "start": 306, "end": 331, "type": 1, "type_": "Process" }, { "id": "T6", "start": 336, "end": 365, "type": 2, "type_": "Task" }, { "id": "T7", "start": 583, "end": 626, "type": 1, "type_": "Process" }, { "id": "T8", "start": 835, "end": 856, "type": 1, "type_": "Process" }, { "id": "T9", "start": 265, "end": 273, "type": 0, "type_": "Material" }, { "id": "T10", "start": 291, "end": 304, "type": 1, "type_": "Process" }, { "id": "T11", "start": 189, "end": 197, "type": 0, "type_": "Material" }, { "id": "T12", "start": 1030, "end": 1037, "type": 1, "type_": "Process" }, { "id": "T13", "start": 1094, "end": 1102, "type": 0, "type_": "Material" }, { "id": "T14", "start": 973, "end": 1001, "type": 1, "type_": "Process" }, { "id": "T15", "start": 1003, "end": 1025, "type": 1, "type_": "Process" }, { "id": "T16", "start": 864, "end": 871, "type": 0, "type_": "Material" }, { "id": "T17", "start": 875, "end": 888, "type": 0, "type_": "Material" }, { "id": "T18", "start": 421, "end": 433, "type": 1, "type_": "Process" }, { "id": "T19", "start": 456, "end": 471, "type": 1, "type_": "Process" }, { "id": "T20", "start": 1045, "end": 1053, "type": 1, "type_": "Process" } ]
[ { "arg1": "T10", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T10", "relation": 1, "relation_": "Synonym-of" } ]
S0021999113005718
Numerical simulation of the gas flow through such non-trivial internal geometries is, however, extremely challenging. This is because conventional continuum fluid dynamics, which assumes that locally a gas is close to a state of thermodynamic equilibrium, becomes invalid or inaccurate as the smallest characteristic scale of the geometry (e.g. the channel height) approaches the mean distance between molecular collisions, λ [1]. An accurate and flexible modelling alternative for these cases is the direct simulation Monte Carlo method (DSMC) [2]. However, DSMC can be prohibitively expensive for internal-flow applications, which typically have a geometry of high-aspect ratio (i.e. are extremely long, relative to their cross-section). The high-aspect ratio creates a formidable multiscale problem: processes need to be resolved occurring over the smallest characteristic scale of the geometry (e.g. a channelʼs height), as well as over the largest characteristic scale of the geometry (e.g. the length of a long channel network), simultaneously.
[ { "id": "T1", "start": 134, "end": 171, "type": 2, "type_": "Task" }, { "id": "T2", "start": 0, "end": 20, "type": 2, "type_": "Task" }, { "id": "T3", "start": 0, "end": 36, "type": 1, "type_": "Process" }, { "id": "T4", "start": 402, "end": 422, "type": 1, "type_": "Process" }, { "id": "T5", "start": 501, "end": 537, "type": 1, "type_": "Process" }, { "id": "T6", "start": 599, "end": 625, "type": 2, "type_": "Task" }, { "id": "T7", "start": 229, "end": 254, "type": 2, "type_": "Task" }, { "id": "T8", "start": 349, "end": 363, "type": 0, "type_": "Material" }, { "id": "T9", "start": 744, "end": 761, "type": 1, "type_": "Process" }, { "id": "T10", "start": 539, "end": 543, "type": 1, "type_": "Process" }, { "id": "T11", "start": 559, "end": 563, "type": 1, "type_": "Process" }, { "id": "T12", "start": 772, "end": 801, "type": 2, "type_": "Task" }, { "id": "T13", "start": 202, "end": 205, "type": 0, "type_": "Material" } ]
[ { "arg1": "T5", "arg2": "T10", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T10", "arg2": "T5", "relation": 1, "relation_": "Synonym-of" } ]
S0029549314001551
An increase of neutron leakage from the core region can be achieved through modifications in the core geometry (usually by adopting a pan-cake geometry of the active core region at the expense of the general neutron economy). Extensive studies determined a set of core design modifications that optimised the total sodium void reactivity (becoming less positive). Among the most efficient design solutions identified is an enlarged sodium plenum above the active core region in combination with an absorber layer above the sodium plenum (to reduce neutron backscattering from the reflector region above the plenum). Fig. 19 shows the combined effect of different upper plenum thicknesses of the absorber and boron layers. It can be observed that the sequential increase of the layer's thickness converge to an asymptotic value of reactivity reduction slightly over 800pcm. The pair of values selected was 60cm for the sodium plenum and 30cm for the boron layer. These modifications implied a considerable increase in the sub-assembly length that was compensated by reducing the upper axial reflector width (Sun et al., 2013).
[ { "id": "T1", "start": 264, "end": 337, "type": 2, "type_": "Task" }, { "id": "T2", "start": 76, "end": 110, "type": 1, "type_": "Process" }, { "id": "T3", "start": 134, "end": 177, "type": 1, "type_": "Process" }, { "id": "T4", "start": 389, "end": 405, "type": 1, "type_": "Process" }, { "id": "T5", "start": 423, "end": 445, "type": 1, "type_": "Process" }, { "id": "T6", "start": 498, "end": 512, "type": 1, "type_": "Process" }, { "id": "T7", "start": 541, "end": 570, "type": 1, "type_": "Process" }, { "id": "T8", "start": 40, "end": 51, "type": 0, "type_": "Material" }, { "id": "T9", "start": 15, "end": 30, "type": 1, "type_": "Process" }, { "id": "T10", "start": 208, "end": 215, "type": 0, "type_": "Material" }, { "id": "T11", "start": 315, "end": 321, "type": 0, "type_": "Material" }, { "id": "T12", "start": 432, "end": 445, "type": 0, "type_": "Material" }, { "id": "T13", "start": 463, "end": 467, "type": 0, "type_": "Material" }, { "id": "T14", "start": 523, "end": 536, "type": 0, "type_": "Material" }, { "id": "T15", "start": 541, "end": 613, "type": 2, "type_": "Task" }, { "id": "T16", "start": 607, "end": 613, "type": 0, "type_": "Material" }, { "id": "T17", "start": 663, "end": 675, "type": 0, "type_": "Material" }, { "id": "T18", "start": 695, "end": 720, "type": 0, "type_": "Material" }, { "id": "T19", "start": 830, "end": 850, "type": 1, "type_": "Process" }, { "id": "T20", "start": 918, "end": 931, "type": 0, "type_": "Material" }, { "id": "T21", "start": 949, "end": 960, "type": 0, "type_": "Material" }, { "id": "T22", "start": 1005, "end": 1040, "type": 1, "type_": "Process" }, { "id": "T23", "start": 1065, "end": 1105, "type": 2, "type_": "Task" }, { "id": "T24", "start": 1084, "end": 1099, "type": 0, "type_": "Material" }, { "id": "T25", "start": 761, "end": 794, "type": 1, "type_": "Process" } ]
[ { "arg1": "T7", "arg2": "T6", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T7", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T5", "arg2": "T4", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T6", "arg2": "T4", "relation": 2, "relation_": "Hyponym-of" } ]
S2212667814000975
It has been acknowledged that megalopolises are playing a leading role in the processes of both economic development and culture change. Thereupon, the new emphases on sustainability of transportation system in megalopolis are creating new demands for adequate approach to measure its performance and diagnosis potential drawbacks. By examining the descriptions of sustainable transport system as well as its evaluating approach, a framework with the general applicability and easily accessible data resource for evaluating sustainability of transport system in megalopolis is developed based on nature of regional structure and the feature transport demand in megalopolis. The proposed framework is applied in the analysis and comparison of Jing-Jin-Ji and Yangtze River Delta..
[ { "id": "T1", "start": 96, "end": 116, "type": 1, "type_": "Process" }, { "id": "T2", "start": 121, "end": 135, "type": 1, "type_": "Process" }, { "id": "T3", "start": 168, "end": 207, "type": 1, "type_": "Process" }, { "id": "T4", "start": 377, "end": 393, "type": 1, "type_": "Process" }, { "id": "T5", "start": 542, "end": 558, "type": 1, "type_": "Process" }, { "id": "T6", "start": 678, "end": 707, "type": 2, "type_": "Task" }, { "id": "T7", "start": 261, "end": 330, "type": 2, "type_": "Task" }, { "id": "T8", "start": 432, "end": 441, "type": 0, "type_": "Material" } ]
[]
S0045782512000266
In this work, a numerical strategy for designing an optimal maintenance scheduling for a structure, accounting explicitly for the effects of uncertainty is suggested. This contribution, which can be regarded as an extension of the methods developed in [23], presents several novel aspects over similar approaches proposed in the literature. Firstly, the initiation and propagation of fatigue crack is modeled efficiently by means of cohesive zone elements [24–26]. The application of this class of elements allows modeling the crack initiation and propagation within a unified framework. It should be noted that cohesive zone elements have already been used for uncertainty quantification of the crack propagation phenomenon [27,28]. However its application within the context of maintenance scheduling constitutes a novelty. The second innovative aspect of this contribution refers to the assessment of the reliability sensitivity with respect to the variables that define the maintenance scheduling. The estimation of this sensitivity, which is required in order to determine the optimal maintenance schedule within the proposed framework, can be quite demanding as the model characterizing repair of a cracked structure leads to a discontinuous performance function associated with the failure probability. A new approach for modeling this function is proposed herein. The continuous and discontinuous parts respectively of the function are considered separately to estimate accurately the gradients of the failure events.
[ { "id": "T1", "start": 612, "end": 634, "type": 0, "type_": "Material" }, { "id": "T2", "start": 662, "end": 724, "type": 2, "type_": "Task" }, { "id": "T3", "start": 780, "end": 802, "type": 2, "type_": "Task" }, { "id": "T4", "start": 890, "end": 1000, "type": 1, "type_": "Process" }, { "id": "T5", "start": 1006, "end": 1036, "type": 1, "type_": "Process" }, { "id": "T6", "start": 1082, "end": 1140, "type": 2, "type_": "Task" }, { "id": "T7", "start": 1376, "end": 1410, "type": 0, "type_": "Material" }, { "id": "T9", "start": 52, "end": 98, "type": 2, "type_": "Task" }, { "id": "T8", "start": 16, "end": 34, "type": 1, "type_": "Process" }, { "id": "T10", "start": 1469, "end": 1524, "type": 1, "type_": "Process" }, { "id": "T11", "start": 354, "end": 397, "type": 1, "type_": "Process" }, { "id": "T12", "start": 433, "end": 455, "type": 0, "type_": "Material" }, { "id": "T13", "start": 489, "end": 506, "type": 0, "type_": "Material" }, { "id": "T14", "start": 514, "end": 586, "type": 2, "type_": "Task" }, { "id": "T15", "start": 100, "end": 152, "type": 1, "type_": "Process" }, { "id": "T16", "start": 1172, "end": 1222, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1234, "end": 1308, "type": 2, "type_": "Task" }, { "id": "T18", "start": 569, "end": 586, "type": 1, "type_": "Process" } ]
[]
S0301010414003516
Arrays of TFTs and circuits were fabricated on precleaned, 5cm×5cm, 125μm thick polyethylene naphthalate (PEN) substrates (Dupont-Teijin). Full details of our vacuum-fabrication procedures have been given in previous publications [17–19,23]. Briefly, aluminium gate electrodes and associated tracks were vacuum evaporated onto the substrates through shadow masks. Subsequently, the substrates were attached to a cooled web-coater drum (Aerre Machines). With the drum rotating at a linear speed of 25m/min under vacuum, flash-evaporated TPGDA monomer vapour which condensed onto the substrates was cross-linked by exposure, in situ, to a plasma. The resulting smooth, pinhole-free films were typically 500nm to 1μm thick with a measured dielectric constant varying in the range 4–5. For circuit fabrication, the insulator was patterned using shadow masks to define rectangular areas separated by 1mm gaps to act as vias for inter-layer metallic connections. The substrates were then transferred into an evaporator (Minispectros, Kurt Lesker) integrated into a nitrogen glovebox for the vacuum-deposition (2.4nm/min) of DNTT onto the insulator. Without exposing the substrates to ambient air, the gold source/drain metallisation layer was deposited through a shadow mask in the same evaporator.
[ { "id": "T1", "start": 10, "end": 14, "type": 0, "type_": "Material" }, { "id": "T2", "start": 19, "end": 27, "type": 0, "type_": "Material" }, { "id": "T3", "start": 80, "end": 104, "type": 0, "type_": "Material" }, { "id": "T4", "start": 106, "end": 109, "type": 0, "type_": "Material" }, { "id": "T5", "start": 80, "end": 121, "type": 0, "type_": "Material" }, { "id": "T6", "start": 159, "end": 188, "type": 1, "type_": "Process" }, { "id": "T7", "start": 251, "end": 276, "type": 0, "type_": "Material" }, { "id": "T8", "start": 292, "end": 298, "type": 0, "type_": "Material" }, { "id": "T9", "start": 304, "end": 321, "type": 1, "type_": "Process" }, { "id": "T10", "start": 331, "end": 341, "type": 0, "type_": "Material" }, { "id": "T11", "start": 350, "end": 362, "type": 0, "type_": "Material" }, { "id": "T12", "start": 382, "end": 392, "type": 0, "type_": "Material" }, { "id": "T13", "start": 419, "end": 434, "type": 0, "type_": "Material" }, { "id": "T14", "start": 462, "end": 466, "type": 0, "type_": "Material" }, { "id": "T15", "start": 519, "end": 556, "type": 0, "type_": "Material" }, { "id": "T16", "start": 582, "end": 592, "type": 0, "type_": "Material" }, { "id": "T17", "start": 637, "end": 643, "type": 0, "type_": "Material" }, { "id": "T18", "start": 597, "end": 609, "type": 1, "type_": "Process" }, { "id": "T19", "start": 613, "end": 621, "type": 1, "type_": "Process" }, { "id": "T20", "start": 680, "end": 685, "type": 0, "type_": "Material" }, { "id": "T21", "start": 786, "end": 805, "type": 2, "type_": "Task" }, { "id": "T22", "start": 811, "end": 820, "type": 0, "type_": "Material" }, { "id": "T23", "start": 841, "end": 853, "type": 0, "type_": "Material" }, { "id": "T24", "start": 923, "end": 955, "type": 1, "type_": "Process" }, { "id": "T25", "start": 961, "end": 971, "type": 0, "type_": "Material" }, { "id": "T26", "start": 1002, "end": 1012, "type": 0, "type_": "Material" }, { "id": "T27", "start": 1059, "end": 1076, "type": 0, "type_": "Material" }, { "id": "T28", "start": 1085, "end": 1091, "type": 0, "type_": "Material" }, { "id": "T29", "start": 1085, "end": 1102, "type": 1, "type_": "Process" }, { "id": "T30", "start": 1132, "end": 1141, "type": 0, "type_": "Material" }, { "id": "T31", "start": 1164, "end": 1174, "type": 0, "type_": "Material" }, { "id": "T32", "start": 1178, "end": 1189, "type": 0, "type_": "Material" }, { "id": "T33", "start": 1195, "end": 1232, "type": 0, "type_": "Material" }, { "id": "T34", "start": 1257, "end": 1268, "type": 0, "type_": "Material" }, { "id": "T35", "start": 1281, "end": 1291, "type": 0, "type_": "Material" } ]
[ { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" } ]
S2212671612001497
Our country is rich of line galloping, there are many important galloping data failed to collect systematically and completely because there is no unified management platform. After the galloping occurrence in 2009–2010's winter the department of productive of the State Grid Corporation organized a lot of human to carry out the research of galloping information, this work is time–consuming and inefficient. The State Grid Corporation has used the production management system (PMS) which is a powerful and easy to use. With the help of the system we can create a galloping database which can save resources and storage the galloping data. To build and put it into application of database can provide technical support for line galloping prevention and galloping research work.
[ { "id": "T2", "start": 23, "end": 37, "type": 1, "type_": "Process" }, { "id": "T3", "start": 64, "end": 78, "type": 0, "type_": "Material" }, { "id": "T4", "start": 89, "end": 126, "type": 2, "type_": "Task" }, { "id": "T5", "start": 147, "end": 174, "type": 1, "type_": "Process" }, { "id": "T6", "start": 186, "end": 206, "type": 1, "type_": "Process" }, { "id": "T7", "start": 316, "end": 364, "type": 1, "type_": "Process" }, { "id": "T8", "start": 450, "end": 478, "type": 0, "type_": "Material" }, { "id": "T9", "start": 480, "end": 483, "type": 0, "type_": "Material" }, { "id": "T10", "start": 557, "end": 584, "type": 2, "type_": "Task" }, { "id": "T11", "start": 595, "end": 640, "type": 1, "type_": "Process" }, { "id": "T12", "start": 626, "end": 640, "type": 0, "type_": "Material" }, { "id": "T13", "start": 667, "end": 690, "type": 1, "type_": "Process" }, { "id": "T14", "start": 703, "end": 720, "type": 1, "type_": "Process" }, { "id": "T15", "start": 725, "end": 750, "type": 1, "type_": "Process" }, { "id": "T16", "start": 755, "end": 778, "type": 1, "type_": "Process" }, { "id": "T1", "start": 64, "end": 73, "type": 1, "type_": "Process" }, { "id": "T17", "start": 342, "end": 351, "type": 1, "type_": "Process" }, { "id": "T18", "start": 566, "end": 575, "type": 1, "type_": "Process" } ]
[ { "arg1": "T8", "arg2": "T9", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T9", "arg2": "T8", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T1", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T1", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" } ]
S0168365912006207
Unlike conventional materials used in nerve tissue engineering, PAs can be directly injected in vivo into models and spontaneously self-assemble into nanofibers in aqueous solutions. Furthermore, PAs can function as biomimetic materials exemplified by collagen-mimetic PAs [92]. Conventional materials often rely on electrospinning as a manufacturing method to achieve fiber-like structures suitable for use in nerve regeneration. The self-assembly nature of PAs allows them to circumvent costly manufacturing methods. However, in contrast to conventional manufacturing methods like electrospinning where quality and batch-to-batch variability can be tightly controlled, merely relying on self-assembly as a method of large-scale commercial production is still an experimental concept. Perhaps the next step would be to carefully compare and contrast the robustness of self-assemled PAs to electrospun nanofibers. Given that the constituent elements in PAs and external factors like pH can affect its structural assembly, parameters must be finely tuned and optimized in order for PA nanofibers to be used as a full-fledged commercialized medical product [93].
[ { "id": "T1", "start": 64, "end": 67, "type": 0, "type_": "Material" }, { "id": "T2", "start": 84, "end": 181, "type": 1, "type_": "Process" }, { "id": "T3", "start": 196, "end": 199, "type": 0, "type_": "Material" }, { "id": "T4", "start": 216, "end": 236, "type": 0, "type_": "Material" }, { "id": "T5", "start": 252, "end": 272, "type": 0, "type_": "Material" }, { "id": "T6", "start": 316, "end": 331, "type": 1, "type_": "Process" }, { "id": "T7", "start": 335, "end": 357, "type": 1, "type_": "Process" }, { "id": "T8", "start": 361, "end": 390, "type": 2, "type_": "Task" }, { "id": "T9", "start": 459, "end": 462, "type": 0, "type_": "Material" }, { "id": "T10", "start": 489, "end": 517, "type": 1, "type_": "Process" }, { "id": "T11", "start": 543, "end": 577, "type": 1, "type_": "Process" }, { "id": "T12", "start": 583, "end": 598, "type": 1, "type_": "Process" }, { "id": "T13", "start": 689, "end": 702, "type": 1, "type_": "Process" }, { "id": "T14", "start": 718, "end": 751, "type": 2, "type_": "Task" }, { "id": "T15", "start": 830, "end": 850, "type": 1, "type_": "Process" }, { "id": "T16", "start": 855, "end": 912, "type": 2, "type_": "Task" }, { "id": "T17", "start": 953, "end": 956, "type": 0, "type_": "Material" }, { "id": "T18", "start": 1048, "end": 1067, "type": 1, "type_": "Process" }, { "id": "T19", "start": 1081, "end": 1094, "type": 0, "type_": "Material" }, { "id": "T20", "start": 1109, "end": 1154, "type": 2, "type_": "Task" } ]
[ { "arg1": "T6", "arg2": "T7", "relation": 2, "relation_": "Hyponym-of" } ]
S0375960112002885
First-principles calculations have clarified the electronic structure and stability of the W@Si12 cluster under O2 molecule adsorption and reaction. Our results show that the W-encapsulated Si12 hexagonal prism cage is very inert to oxidation. The O2 molecule only weakly adsorbs onto the cluster at relatively low temperatures, in the range of several tens meV. However, significant reaction barriers (0.593–1.118 eV) for the O2 molecule on the cluster are identified on different adsorption sites, nevertheless, these reaction paths are spin forbidden reactions according to Wingerʼs spin selection rule. These results imply that O2 readily desorb from the cluster surface rather than dissociate and oxide the W@Si12 cluster upon excitations. In high temperature and high pressure conditions, the O2 molecules may dissociate on the preferential edge site by overcoming a significantly large energy barrier.
[ { "id": "T1", "start": 0, "end": 29, "type": 1, "type_": "Process" }, { "id": "T2", "start": 49, "end": 83, "type": 2, "type_": "Task" }, { "id": "T3", "start": 91, "end": 105, "type": 0, "type_": "Material" }, { "id": "T4", "start": 112, "end": 114, "type": 0, "type_": "Material" }, { "id": "T5", "start": 112, "end": 147, "type": 2, "type_": "Task" }, { "id": "T6", "start": 233, "end": 242, "type": 2, "type_": "Task" }, { "id": "T7", "start": 175, "end": 215, "type": 1, "type_": "Process" }, { "id": "T8", "start": 248, "end": 250, "type": 0, "type_": "Material" }, { "id": "T9", "start": 311, "end": 327, "type": 1, "type_": "Process" }, { "id": "T10", "start": 372, "end": 401, "type": 2, "type_": "Task" }, { "id": "T11", "start": 403, "end": 417, "type": 2, "type_": "Task" }, { "id": "T12", "start": 427, "end": 429, "type": 0, "type_": "Material" }, { "id": "T13", "start": 446, "end": 453, "type": 0, "type_": "Material" }, { "id": "T14", "start": 289, "end": 296, "type": 0, "type_": "Material" }, { "id": "T15", "start": 539, "end": 563, "type": 2, "type_": "Task" }, { "id": "T16", "start": 632, "end": 634, "type": 0, "type_": "Material" }, { "id": "T17", "start": 659, "end": 666, "type": 0, "type_": "Material" }, { "id": "T18", "start": 712, "end": 726, "type": 0, "type_": "Material" }, { "id": "T19", "start": 748, "end": 793, "type": 1, "type_": "Process" }, { "id": "T20", "start": 799, "end": 801, "type": 0, "type_": "Material" }, { "id": "T21", "start": 860, "end": 907, "type": 1, "type_": "Process" }, { "id": "T22", "start": 482, "end": 492, "type": 1, "type_": "Process" }, { "id": "T23", "start": 520, "end": 528, "type": 1, "type_": "Process" }, { "id": "T24", "start": 124, "end": 134, "type": 1, "type_": "Process" }, { "id": "T25", "start": 139, "end": 147, "type": 1, "type_": "Process" }, { "id": "T26", "start": 577, "end": 605, "type": 1, "type_": "Process" } ]
[ { "arg1": "T11", "arg2": "T10", "relation": 2, "relation_": "Hyponym-of" } ]
S0377221716304258
Two-state models are often insufficient to fit complex traces, therefore we also study the approximate fitting of large M3PPs. In the single class setting, a known limitation of MMPPs is the inability to simultaneously fit many statistical descriptors due to the non-linearity of their underlying equations (Bodrog, Heindl, Horváth, & Telek, 2008; Heindl, Horváth, & Gross, 2006; Horváth & Telek, 2009). This has led to the definition of several approaches to fit complex traces by composing multiple small-sized MMPPs or MAPs using Kronecker operators (Andersen & Nielsen, 1998; Casale, Zhang, & Smirni, 2010; Horváth & Telek, 2002). These methods employ composition operators for moment fitting, offering a different trade-off between computational cost and fitting accuracy compared to fitting methods based on the EM algorithm (Breuer, 2002; Horváth & Okamura, 2013; Klemm, Lindemann, & Lohmann, 2003). In particular, the superposition operator allows one to describe a trace by the statistical multiplexing of several MMPPs, at the expense of an exponential growth of the number of states in the resulting process (Sriram & Whitt, 1986). This state space explosion is an obstacle for the application of MMPPs and MAPs to modeling real systems; for example it considerably slows down, or even renders infeasible, the numerical evaluation of queueing models by matrix geometric methods (Bini, Meini, Steffé, Pérez, & Houdt, 2012; Pérez, Velthoven, & Houdt, 2008).
[ { "id": "T1", "start": 120, "end": 125, "type": 0, "type_": "Material" }, { "id": "T2", "start": 178, "end": 183, "type": 0, "type_": "Material" }, { "id": "T5", "start": 656, "end": 677, "type": 1, "type_": "Process" }, { "id": "T7", "start": 818, "end": 830, "type": 1, "type_": "Process" }, { "id": "T8", "start": 926, "end": 948, "type": 1, "type_": "Process" }, { "id": "T9", "start": 1364, "end": 1388, "type": 2, "type_": "Task" }, { "id": "T10", "start": 0, "end": 16, "type": 1, "type_": "Process" }, { "id": "T3", "start": 482, "end": 526, "type": 1, "type_": "Process" }, { "id": "T4", "start": 533, "end": 552, "type": 1, "type_": "Process" }, { "id": "T6", "start": 1023, "end": 1028, "type": 0, "type_": "Material" }, { "id": "T11", "start": 1208, "end": 1222, "type": 0, "type_": "Material" }, { "id": "T12", "start": 1226, "end": 1247, "type": 2, "type_": "Task" }, { "id": "T13", "start": 1321, "end": 1360, "type": 2, "type_": "Task" }, { "id": "T14", "start": 81, "end": 125, "type": 2, "type_": "Task" }, { "id": "T15", "start": 228, "end": 251, "type": 1, "type_": "Process" }, { "id": "T16", "start": 286, "end": 306, "type": 1, "type_": "Process" }, { "id": "T17", "start": 1148, "end": 1169, "type": 1, "type_": "Process" }, { "id": "T18", "start": 789, "end": 804, "type": 2, "type_": "Task" }, { "id": "T19", "start": 987, "end": 1028, "type": 1, "type_": "Process" } ]
[]
S0370269304009062
The microwave background is not the only universal photon field that has to be taken in consideration. Especially interesting is the isotropic infrared and optical background (IRB). The number density of IRB is smaller than that of MBR by more that two orders of magnitude. On the other hand, protons of lower energy can interact on the IRB, and the smaller number density has to be weighted with the higher flux of interacting protons. The present Universe is optically thin to 1019 eV and lower energy protons, but even at low redshifts the proton interaction rate quickly increases. This is different from the interactions on MBR, where the interacting protons quickly lose their energy even at z=0. The cosmological evolution of UHECR injection is thus of major importance for the contribution of such interactions to the flux of cosmogenic neutrinos.
[ { "id": "T1", "start": 4, "end": 24, "type": 1, "type_": "Process" }, { "id": "T2", "start": 133, "end": 174, "type": 1, "type_": "Process" }, { "id": "T3", "start": 176, "end": 179, "type": 1, "type_": "Process" }, { "id": "T4", "start": 346, "end": 435, "type": 1, "type_": "Process" }, { "id": "T5", "start": 441, "end": 457, "type": 2, "type_": "Task" }, { "id": "T6", "start": 543, "end": 566, "type": 2, "type_": "Task" }, { "id": "T7", "start": 707, "end": 748, "type": 1, "type_": "Process" }, { "id": "T8", "start": 826, "end": 854, "type": 2, "type_": "Task" }, { "id": "T9", "start": 79, "end": 101, "type": 1, "type_": "Process" }, { "id": "T10", "start": 613, "end": 632, "type": 2, "type_": "Task" }, { "id": "T11", "start": 644, "end": 663, "type": 0, "type_": "Material" }, { "id": "T12", "start": 664, "end": 701, "type": 1, "type_": "Process" }, { "id": "T13", "start": 204, "end": 207, "type": 1, "type_": "Process" }, { "id": "T14", "start": 232, "end": 235, "type": 1, "type_": "Process" }, { "id": "T15", "start": 337, "end": 340, "type": 1, "type_": "Process" }, { "id": "T16", "start": 416, "end": 435, "type": 0, "type_": "Material" }, { "id": "T17", "start": 629, "end": 632, "type": 0, "type_": "Material" }, { "id": "T18", "start": 543, "end": 549, "type": 0, "type_": "Material" }, { "id": "T19", "start": 293, "end": 300, "type": 0, "type_": "Material" }, { "id": "T20", "start": 41, "end": 63, "type": 0, "type_": "Material" }, { "id": "T21", "start": 491, "end": 511, "type": 0, "type_": "Material" } ]
[ { "arg1": "T2", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" } ]
S0009261412006513
Water is the most important liquid, and the nature of its structure remains a topic of keen debate and an active area of research [1–9]. Much of this debate centers around whether water has a mainly tetrahedral structure with a continuum of distorted hydrogen bonds, or if it contains a mixture of two distinct components. One major development in recent years is the application of inner-shell spectroscopic techniques, such as X-ray absorption spectroscopy (XAS) and X-ray emission spectroscopy (XES) at the oxygen K-edge to investigate the structure of water [2,10–12]. These methods can provide a direct structural probe of water, providing insight into the nature of its hydrogen bonding network. Theoretical studies play a critical role in these studies, since the analysis of the experimental data requires calculations to provide a link between the observed spectral features and the underlying structure. However, the simulation of the XAS or XES for liquid water presents a difficult challenge because it requires accurate molecular dynamics simulations to provide a correct description of the molecular structure coupled with accurate calculations of the spectral properties, i.e. excitation energies and line intensities. Furthermore, adequate sampling over molecular configurations also needs to be accounted for.
[ { "id": "T1", "start": 0, "end": 5, "type": 0, "type_": "Material" }, { "id": "T2", "start": 180, "end": 185, "type": 0, "type_": "Material" }, { "id": "T3", "start": 241, "end": 265, "type": 0, "type_": "Material" }, { "id": "T4", "start": 383, "end": 419, "type": 1, "type_": "Process" }, { "id": "T5", "start": 429, "end": 458, "type": 1, "type_": "Process" }, { "id": "T6", "start": 469, "end": 496, "type": 1, "type_": "Process" }, { "id": "T7", "start": 556, "end": 561, "type": 0, "type_": "Material" }, { "id": "T8", "start": 945, "end": 948, "type": 1, "type_": "Process" }, { "id": "T9", "start": 952, "end": 955, "type": 1, "type_": "Process" }, { "id": "T10", "start": 676, "end": 700, "type": 2, "type_": "Task" }, { "id": "T11", "start": 1024, "end": 1063, "type": 1, "type_": "Process" }, { "id": "T12", "start": 1104, "end": 1123, "type": 2, "type_": "Task" }, { "id": "T13", "start": 1166, "end": 1185, "type": 2, "type_": "Task" }, { "id": "T14", "start": 1192, "end": 1211, "type": 2, "type_": "Task" }, { "id": "T15", "start": 1216, "end": 1232, "type": 2, "type_": "Task" }, { "id": "T16", "start": 1247, "end": 1264, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1270, "end": 1294, "type": 1, "type_": "Process" }, { "id": "T18", "start": 787, "end": 804, "type": 0, "type_": "Material" }, { "id": "T19", "start": 814, "end": 826, "type": 0, "type_": "Material" }, { "id": "T20", "start": 857, "end": 883, "type": 1, "type_": "Process" }, { "id": "T21", "start": 892, "end": 912, "type": 2, "type_": "Task" }, { "id": "T22", "start": 601, "end": 633, "type": 1, "type_": "Process" }, { "id": "T23", "start": 510, "end": 523, "type": 0, "type_": "Material" }, { "id": "T24", "start": 28, "end": 34, "type": 0, "type_": "Material" }, { "id": "T25", "start": 460, "end": 463, "type": 1, "type_": "Process" }, { "id": "T26", "start": 498, "end": 501, "type": 1, "type_": "Process" }, { "id": "T27", "start": 628, "end": 633, "type": 0, "type_": "Material" }, { "id": "T28", "start": 960, "end": 972, "type": 0, "type_": "Material" } ]
[ { "arg1": "T5", "arg2": "T25", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T25", "arg2": "T5", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T26", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T26", "arg2": "T6", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T1", "arg2": "T24", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T4", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T14", "arg2": "T13", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T15", "arg2": "T13", "relation": 2, "relation_": "Hyponym-of" } ]
S1566253516300069
The above discussion also lays bare the difference of perspectives between the fusion of hard constraints and knowledge-base merging: the idea of Konieczny and Pino-Perez is to explain the fusion of plain epistemic states, understood as a set of plausible worlds, by the existence of underlying partial orderings or numerical plausibility degrees (obtained by distances), based on axioms that only use plausible sets attached to these orderings. In [67] the same authors use both hard (integrity) constraints and belief sets referring to plausible worlds, and try to extend both the AGM revision and knowledge-based merging. However, they do not envisage the merging of integrity constraints discussed in the previous section. The belief revision and merging literature takes an external point of view on cognitive processes under study. The underlying ordered structures are here a consequence of the merging postulates, but they do not appear explicitly in the axioms and they are not observable from the outside. On the contrary, our approach is to construct fusion rules that only rely on what is explicitly supplied by sources. In the sequel we consider the counterpart of our fusion postulates for ranked models, that can be expressed by means of total orders of possible worlds or by their encodings on a plausibility scale.
[ { "id": "T1", "start": 110, "end": 132, "type": 1, "type_": "Process" }, { "id": "T2", "start": 89, "end": 105, "type": 1, "type_": "Process" }, { "id": "T3", "start": 189, "end": 221, "type": 1, "type_": "Process" }, { "id": "T4", "start": 295, "end": 312, "type": 1, "type_": "Process" }, { "id": "T5", "start": 316, "end": 346, "type": 1, "type_": "Process" }, { "id": "T6", "start": 480, "end": 508, "type": 1, "type_": "Process" }, { "id": "T7", "start": 513, "end": 524, "type": 1, "type_": "Process" }, { "id": "T8", "start": 583, "end": 595, "type": 1, "type_": "Process" }, { "id": "T9", "start": 600, "end": 623, "type": 1, "type_": "Process" }, { "id": "T10", "start": 659, "end": 691, "type": 1, "type_": "Process" }, { "id": "T11", "start": 731, "end": 758, "type": 2, "type_": "Task" }, { "id": "T12", "start": 853, "end": 871, "type": 0, "type_": "Material" }, { "id": "T13", "start": 1052, "end": 1074, "type": 1, "type_": "Process" }, { "id": "T14", "start": 1204, "end": 1217, "type": 1, "type_": "Process" } ]
[]
S0370269304008809
The most ambitious goal may be stated as the one of detecting the location of, say, one missing level on an otherwise complete sequence. Dyson, in a recent review [7], uses information theory concepts and argues that correlations in a sequence may provide the necessary redundancy from which error correcting codes can be constructed. At one extreme where no correlations and therefore no redundancy are present (Poissonian sequence), there is no possibility of detecting one missing level. At the other extreme, a sequence of equally spaced levels (picket fence), there is a maximum redundancy and a missed level can be obviously detected as a hole in the spectrum. Eigenvalues of random matrices, which exhibit characteristic correlations, correspond to an intermediate situation between these two extremes. The attempts to locate in the last case a single missed level have remained unsuccessful so far. However, it should be mentioned that for two-dimensional chaotic systems where, besides correlations of the order of one mean spacing as described by random matrices, the presence and the role of long range correlations governed by the shortest periodic orbits and reflected in Weyl's law describing the average spectral density, is well understood. It is then possible to approximately locate, from the study of spectral fluctuations, a single missed level [8].
[ { "id": "T1", "start": 52, "end": 135, "type": 2, "type_": "Task" }, { "id": "T2", "start": 173, "end": 200, "type": 1, "type_": "Process" }, { "id": "T3", "start": 217, "end": 243, "type": 1, "type_": "Process" }, { "id": "T4", "start": 292, "end": 333, "type": 2, "type_": "Task" }, { "id": "T5", "start": 462, "end": 489, "type": 2, "type_": "Task" }, { "id": "T6", "start": 356, "end": 411, "type": 1, "type_": "Process" }, { "id": "T7", "start": 413, "end": 432, "type": 1, "type_": "Process" }, { "id": "T8", "start": 515, "end": 548, "type": 2, "type_": "Task" }, { "id": "T9", "start": 550, "end": 562, "type": 2, "type_": "Task" }, { "id": "T10", "start": 601, "end": 665, "type": 2, "type_": "Task" }, { "id": "T11", "start": 826, "end": 871, "type": 2, "type_": "Task" }, { "id": "T12", "start": 1280, "end": 1300, "type": 2, "type_": "Task" }, { "id": "T13", "start": 1343, "end": 1364, "type": 2, "type_": "Task" }, { "id": "T14", "start": 1311, "end": 1341, "type": 1, "type_": "Process" }, { "id": "T15", "start": 948, "end": 979, "type": 2, "type_": "Task" }, { "id": "T16", "start": 1185, "end": 1235, "type": 1, "type_": "Process" } ]
[ { "arg1": "T6", "arg2": "T7", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T8", "arg2": "T9", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T12", "arg2": "T13", "relation": 2, "relation_": "Hyponym-of" } ]
S0301010414003115
The optimised structure at the B3LYP/aug-cc-pVTZ level was then used to perform calculations of the lowest electronic singlet excited states with the coupled cluster linear response (LR) coupled cluster hierarchy CCS, CC2, CCSD and CC3, along with perturbative corrected methods CIS(D) and CCSDR(3). The correlated response methods were performed with an all-electron atomic natural orbital (ANO) basis set contracted to 6s5p4d3f1g on manganese, [47] together with the cc-pVTZ basis set on the oxygen atoms. The all-electron correlated calculations invoked a 13 orbital frozen core (O 1s, Mn 1s2s2p3s3p). Trial calculations correlating these orbitals only had a minor effect on excitation energies. For comparison the EOM-CCSD method with the cc-pVTZ basis on all atoms was tested to compare with LR-CCSD. These formally give exactly the same excitation energies, although the transition moments are more accurate for LR-CCSD. Abelian symmetry (D2) was used in all correlated excited state calculations.
[ { "id": "T1", "start": 31, "end": 48, "type": 1, "type_": "Process" }, { "id": "T2", "start": 100, "end": 125, "type": 1, "type_": "Process" }, { "id": "T3", "start": 166, "end": 181, "type": 1, "type_": "Process" }, { "id": "T4", "start": 183, "end": 185, "type": 1, "type_": "Process" }, { "id": "T5", "start": 187, "end": 212, "type": 1, "type_": "Process" }, { "id": "T6", "start": 248, "end": 278, "type": 1, "type_": "Process" }, { "id": "T7", "start": 304, "end": 331, "type": 1, "type_": "Process" }, { "id": "T8", "start": 368, "end": 390, "type": 1, "type_": "Process" }, { "id": "T9", "start": 392, "end": 395, "type": 1, "type_": "Process" }, { "id": "T10", "start": 435, "end": 444, "type": 0, "type_": "Material" }, { "id": "T11", "start": 494, "end": 500, "type": 0, "type_": "Material" }, { "id": "T12", "start": 421, "end": 431, "type": 1, "type_": "Process" }, { "id": "T13", "start": 469, "end": 476, "type": 1, "type_": "Process" }, { "id": "T14", "start": 512, "end": 548, "type": 1, "type_": "Process" }, { "id": "T15", "start": 583, "end": 584, "type": 0, "type_": "Material" }, { "id": "T16", "start": 589, "end": 591, "type": 0, "type_": "Material" }, { "id": "T17", "start": 678, "end": 697, "type": 2, "type_": "Task" }, { "id": "T18", "start": 718, "end": 733, "type": 1, "type_": "Process" }, { "id": "T19", "start": 743, "end": 750, "type": 1, "type_": "Process" }, { "id": "T20", "start": 797, "end": 804, "type": 1, "type_": "Process" }, { "id": "T21", "start": 843, "end": 862, "type": 2, "type_": "Task" }, { "id": "T22", "start": 877, "end": 895, "type": 2, "type_": "Task" }, { "id": "T23", "start": 918, "end": 925, "type": 1, "type_": "Process" }, { "id": "T24", "start": 927, "end": 943, "type": 1, "type_": "Process" }, { "id": "T25", "start": 976, "end": 1002, "type": 1, "type_": "Process" }, { "id": "T26", "start": 4, "end": 23, "type": 2, "type_": "Task" }, { "id": "T27", "start": 945, "end": 947, "type": 1, "type_": "Process" }, { "id": "T28", "start": 213, "end": 216, "type": 1, "type_": "Process" }, { "id": "T29", "start": 218, "end": 221, "type": 1, "type_": "Process" }, { "id": "T30", "start": 223, "end": 227, "type": 1, "type_": "Process" }, { "id": "T31", "start": 232, "end": 235, "type": 1, "type_": "Process" }, { "id": "T32", "start": 279, "end": 285, "type": 1, "type_": "Process" }, { "id": "T33", "start": 290, "end": 298, "type": 1, "type_": "Process" } ]
[ { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T8", "arg2": "T9", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T9", "arg2": "T8", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T24", "arg2": "T27", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T27", "arg2": "T24", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T28", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T29", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T30", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T31", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T32", "arg2": "T6", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T33", "arg2": "T6", "relation": 2, "relation_": "Hyponym-of" } ]
S0927025614007137
Structural adhesives are increasingly used for bonding components within critical load bearing engineering structures such as aerospace and automotives. Typically these adhesives are based on epoxy polymers. Epoxies are inherently brittle due to their homogeneous microstructure and highly cross linked nature. Thus, there has been much research focused on improving the fracture toughness of epoxy polymers by incorporating a second minority phase at the nano-scale. These modifiers fall into one of two main categories: inorganic additives, e.g. silica [1,2], glass [3], alumina [4], nano-clays [5] and carbon nanotubes [6,7] or organic, usually rubber particles. Rubbery additives can be either core–shell rubber particles [8–10] or can form during curing via reaction induced phase separation mechanisms [11,12]. The primary energy dissipation mechanisms for rubber toughened epoxies are known to be both plastic void growth and shear band development [13]. It has also been shown that a combination of the above additives to create a hybrid material can provide synergistic toughening effects, e.g. carbon nanotubes and silica nanoparticles [14] or rubber with silica nanoparticles [15–17].
[ { "id": "T1", "start": 0, "end": 20, "type": 0, "type_": "Material" }, { "id": "T2", "start": 47, "end": 65, "type": 0, "type_": "Material" }, { "id": "T3", "start": 73, "end": 117, "type": 0, "type_": "Material" }, { "id": "T4", "start": 126, "end": 135, "type": 0, "type_": "Material" }, { "id": "T5", "start": 140, "end": 151, "type": 0, "type_": "Material" }, { "id": "T6", "start": 169, "end": 178, "type": 0, "type_": "Material" }, { "id": "T7", "start": 192, "end": 206, "type": 0, "type_": "Material" }, { "id": "T8", "start": 208, "end": 215, "type": 0, "type_": "Material" }, { "id": "T9", "start": 264, "end": 278, "type": 0, "type_": "Material" }, { "id": "T10", "start": 371, "end": 389, "type": 1, "type_": "Process" }, { "id": "T11", "start": 357, "end": 407, "type": 2, "type_": "Task" }, { "id": "T12", "start": 393, "end": 407, "type": 0, "type_": "Material" }, { "id": "T13", "start": 411, "end": 466, "type": 1, "type_": "Process" }, { "id": "T14", "start": 522, "end": 541, "type": 0, "type_": "Material" }, { "id": "T15", "start": 548, "end": 554, "type": 0, "type_": "Material" }, { "id": "T16", "start": 562, "end": 567, "type": 0, "type_": "Material" }, { "id": "T17", "start": 573, "end": 580, "type": 0, "type_": "Material" }, { "id": "T18", "start": 586, "end": 596, "type": 0, "type_": "Material" }, { "id": "T19", "start": 605, "end": 621, "type": 0, "type_": "Material" }, { "id": "T20", "start": 631, "end": 638, "type": 0, "type_": "Material" }, { "id": "T21", "start": 648, "end": 664, "type": 0, "type_": "Material" }, { "id": "T22", "start": 666, "end": 683, "type": 0, "type_": "Material" }, { "id": "T23", "start": 698, "end": 725, "type": 0, "type_": "Material" }, { "id": "T24", "start": 752, "end": 758, "type": 1, "type_": "Process" }, { "id": "T25", "start": 763, "end": 807, "type": 1, "type_": "Process" }, { "id": "T26", "start": 829, "end": 858, "type": 1, "type_": "Process" }, { "id": "T27", "start": 863, "end": 887, "type": 0, "type_": "Material" }, { "id": "T28", "start": 909, "end": 928, "type": 2, "type_": "Task" }, { "id": "T29", "start": 933, "end": 955, "type": 2, "type_": "Task" }, { "id": "T30", "start": 909, "end": 928, "type": 1, "type_": "Process" }, { "id": "T31", "start": 933, "end": 955, "type": 1, "type_": "Process" }, { "id": "T32", "start": 1017, "end": 1026, "type": 0, "type_": "Material" }, { "id": "T33", "start": 1039, "end": 1054, "type": 0, "type_": "Material" }, { "id": "T34", "start": 1067, "end": 1097, "type": 1, "type_": "Process" }, { "id": "T35", "start": 1104, "end": 1120, "type": 0, "type_": "Material" }, { "id": "T36", "start": 1125, "end": 1145, "type": 0, "type_": "Material" }, { "id": "T37", "start": 1154, "end": 1186, "type": 0, "type_": "Material" }, { "id": "T38", "start": 1104, "end": 1145, "type": 0, "type_": "Material" } ]
[ { "arg1": "T4", "arg2": "T3", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T5", "arg2": "T3", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T15", "arg2": "T14", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T16", "arg2": "T14", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T17", "arg2": "T14", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T18", "arg2": "T14", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T19", "arg2": "T14", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T21", "arg2": "T20", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T23", "arg2": "T22", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T30", "arg2": "T26", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T31", "arg2": "T26", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T38", "arg2": "T33", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T37", "arg2": "T33", "relation": 2, "relation_": "Hyponym-of" } ]
S2212667814000550
The low-carbon economic development has become the trend and orientation of regional economic development. As the residents of Heilongjiang province, their consumption is the most direct way to achieve the low-carbon lifestyle. Based on the research and discussion of the connotation of low-carbon consumption and its culture, behaviour, preferences and habits, it is concluded that the low-carbon consumption requires us to abide by the life-style of knowledge and culture. Therefore, it is obvious that the development of low-carbon economy is a complex and systematic project, involved the economic development mode, technological innovation mode, consumption values and changes of lifestyle.
[ { "id": "T1", "start": 387, "end": 409, "type": 1, "type_": "Process" }, { "id": "T2", "start": 4, "end": 35, "type": 1, "type_": "Process" }, { "id": "T3", "start": 76, "end": 105, "type": 1, "type_": "Process" }, { "id": "T4", "start": 287, "end": 309, "type": 1, "type_": "Process" }, { "id": "T5", "start": 524, "end": 542, "type": 1, "type_": "Process" }, { "id": "T6", "start": 509, "end": 542, "type": 2, "type_": "Task" }, { "id": "T7", "start": 593, "end": 613, "type": 1, "type_": "Process" }, { "id": "T8", "start": 620, "end": 644, "type": 1, "type_": "Process" }, { "id": "T9", "start": 8, "end": 14, "type": 0, "type_": "Material" }, { "id": "T10", "start": 210, "end": 216, "type": 0, "type_": "Material" }, { "id": "T11", "start": 291, "end": 297, "type": 0, "type_": "Material" }, { "id": "T12", "start": 391, "end": 397, "type": 0, "type_": "Material" }, { "id": "T13", "start": 528, "end": 534, "type": 0, "type_": "Material" } ]
[]
S0370269304009049
One of the challenges in quantum chromodynamics (QCD) is the relativistic bound state problem. In the light-cone Hamiltonian approach [1] light-cone wave functions can be constructed in a boost invariant way. It is necessary to have reliable light-cone wave functions if one wants to calculate high energy scattering, especially exclusive reactions. Many parametrizations assume separability of the dependence on the longitudinal momentum fraction and transverse momentum which is very unlikely since the two momenta are coupled in the kinetic energy operator. Various approaches have been tried to compute such wave functions. One can use the usual equal time Hamiltonian [2] and transform the resulting wave functions into light-cone form with the help of kinematical on-shell equations. The light-cone Hamiltonian in a string picture is formulated in Ref. [3]. More ambitious is the construction of an effective Hamiltonian including the gauge degrees of freedom explicitly and then solving the bound state problem. For mesons this approach [4,5] still needs many parameters to be fixed. Attempts have been made to solve the valence quark wave function for mesons in a simple Hamiltonian with a two-body potential [6].
[ { "id": "T1", "start": 25, "end": 47, "type": 2, "type_": "Task" }, { "id": "T2", "start": 49, "end": 52, "type": 2, "type_": "Task" }, { "id": "T3", "start": 61, "end": 93, "type": 2, "type_": "Task" }, { "id": "T4", "start": 102, "end": 133, "type": 1, "type_": "Process" }, { "id": "T5", "start": 138, "end": 163, "type": 2, "type_": "Task" }, { "id": "T6", "start": 233, "end": 267, "type": 0, "type_": "Material" }, { "id": "T7", "start": 284, "end": 316, "type": 2, "type_": "Task" }, { "id": "T8", "start": 650, "end": 672, "type": 1, "type_": "Process" }, { "id": "T9", "start": 681, "end": 740, "type": 1, "type_": "Process" }, { "id": "T10", "start": 886, "end": 976, "type": 2, "type_": "Task" }, { "id": "T11", "start": 986, "end": 1017, "type": 1, "type_": "Process" }, { "id": "T12", "start": 1118, "end": 1166, "type": 2, "type_": "Task" }, { "id": "T13", "start": 1172, "end": 1217, "type": 1, "type_": "Process" }, { "id": "T14", "start": 758, "end": 788, "type": 1, "type_": "Process" }, { "id": "T15", "start": 1160, "end": 1166, "type": 0, "type_": "Material" }, { "id": "T16", "start": 1128, "end": 1155, "type": 1, "type_": "Process" }, { "id": "T17", "start": 915, "end": 926, "type": 1, "type_": "Process" }, { "id": "T18", "start": 1023, "end": 1029, "type": 0, "type_": "Material" }, { "id": "T19", "start": 339, "end": 348, "type": 1, "type_": "Process" } ]
[ { "arg1": "T1", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T1", "relation": 1, "relation_": "Synonym-of" } ]
S2212667814000070
This paper suggests a design of high quality real-time rotation face detection architecture for gesture recognition of smart TV. For high performance rotated face detection, the multiple-MCT(Modified Census Transform) architecture, which is robust against lighting change, was used. The Adaboost learning algorithm was used for creating optimized learning data. The proposed hardware structure was composed of Color Space Converter, Image Resizer, Noise Filter, Memory Controller Interface, Image Rotator, Image Scaler, MCT Generator, Candidate Detector, Confidence Switch, Confidence Mapper, Position Resizer, Data Grouper, Overlay Processor and Color Overlay Processer. As a result, suggested face detection device can conduct real-time processing at speed of at least 30 frames per second.
[ { "id": "T1", "start": 22, "end": 127, "type": 2, "type_": "Task" }, { "id": "T2", "start": 158, "end": 172, "type": 1, "type_": "Process" }, { "id": "T3", "start": 187, "end": 190, "type": 1, "type_": "Process" }, { "id": "T4", "start": 191, "end": 216, "type": 1, "type_": "Process" }, { "id": "T5", "start": 287, "end": 314, "type": 1, "type_": "Process" }, { "id": "T6", "start": 328, "end": 360, "type": 2, "type_": "Task" }, { "id": "T7", "start": 410, "end": 431, "type": 0, "type_": "Material" }, { "id": "T8", "start": 433, "end": 447, "type": 0, "type_": "Material" }, { "id": "T9", "start": 448, "end": 460, "type": 0, "type_": "Material" }, { "id": "T10", "start": 462, "end": 489, "type": 0, "type_": "Material" }, { "id": "T11", "start": 491, "end": 504, "type": 0, "type_": "Material" }, { "id": "T12", "start": 506, "end": 518, "type": 0, "type_": "Material" }, { "id": "T13", "start": 520, "end": 533, "type": 0, "type_": "Material" }, { "id": "T14", "start": 535, "end": 554, "type": 0, "type_": "Material" }, { "id": "T15", "start": 555, "end": 572, "type": 0, "type_": "Material" }, { "id": "T16", "start": 574, "end": 590, "type": 0, "type_": "Material" }, { "id": "T17", "start": 593, "end": 610, "type": 0, "type_": "Material" }, { "id": "T18", "start": 611, "end": 623, "type": 0, "type_": "Material" }, { "id": "T19", "start": 625, "end": 642, "type": 0, "type_": "Material" }, { "id": "T20", "start": 647, "end": 670, "type": 0, "type_": "Material" }, { "id": "T21", "start": 695, "end": 709, "type": 1, "type_": "Process" }, { "id": "T22", "start": 729, "end": 749, "type": 1, "type_": "Process" }, { "id": "T23", "start": 375, "end": 393, "type": 0, "type_": "Material" }, { "id": "T24", "start": 55, "end": 91, "type": 0, "type_": "Material" }, { "id": "T25", "start": 96, "end": 115, "type": 1, "type_": "Process" } ]
[ { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" } ]
S0749603615302184
However, the measured reflectivity is less than the predicted value (∼96%), which is likely to relate to, amongst other factors, the roughness of the GaN/AlN interfaces particularly for the first layer in the DBR stack and the non-uniformity of the DBR layer thicknesses. Using STEM measurements of the thickness of each layer (on the a-plane) through the thickness of the stack, we calculate a new model (green curve) in which the overall reflectivity is reduced to 85%. This implies that variations in layer thickness through the stack are the main source of the reduced reflectivity in comparison to the model. In fact, a closer look at the cross-sectional STEM data and a careful extraction of layer thickness have revealed that whilst the layer thicknesses are fairly consistent through the DBR stack in the wing regions, there is a monotonic variation in the measured layer thicknesses in the window regions. (The GaN layer width smoothly increases, while the AlN layer thickness decreases through the DBR stack.). This observation could potentially be of practical importance, for samples grown on templates with a uniform defect density, as one could achieve much better reflectivities simply by altering the growth time to counteract the change in growth rate. This possibility is the subject of ongoing investigations. In addition, the presence of cracks and trenches in the top surface may also reduce the measured reflectivity further.
[ { "id": "T2", "start": 278, "end": 295, "type": 1, "type_": "Process" }, { "id": "T3", "start": 383, "end": 404, "type": 1, "type_": "Process" }, { "id": "T4", "start": 644, "end": 669, "type": 0, "type_": "Material" }, { "id": "T5", "start": 684, "end": 713, "type": 1, "type_": "Process" }, { "id": "T6", "start": 1204, "end": 1228, "type": 1, "type_": "Process" }, { "id": "T1", "start": 209, "end": 218, "type": 0, "type_": "Material" }, { "id": "T7", "start": 796, "end": 825, "type": 0, "type_": "Material" }, { "id": "T8", "start": 895, "end": 913, "type": 0, "type_": "Material" }, { "id": "T9", "start": 1008, "end": 1017, "type": 0, "type_": "Material" }, { "id": "T10", "start": 150, "end": 168, "type": 0, "type_": "Material" }, { "id": "T11", "start": 249, "end": 258, "type": 0, "type_": "Material" }, { "id": "T12", "start": 920, "end": 929, "type": 0, "type_": "Material" }, { "id": "T13", "start": 966, "end": 975, "type": 0, "type_": "Material" } ]
[]
S2212671612002120
A design method for network attack and defense simulation platform is discussed in this paper. Firstly the component and function of the platform are analyzed. Then Visio second development method is used to construct the virtual network topology. The parsing of virtual network topology is also researched and the relative flow sheet is described. Lastly an example is carried out to test performance of the platform. Simulation results show the effectiveness of the proposed method.
[ { "id": "T1", "start": 2, "end": 15, "type": 1, "type_": "Process" }, { "id": "T2", "start": 20, "end": 34, "type": 1, "type_": "Process" }, { "id": "T3", "start": 39, "end": 66, "type": 1, "type_": "Process" }, { "id": "T5", "start": 165, "end": 196, "type": 1, "type_": "Process" }, { "id": "T6", "start": 208, "end": 246, "type": 2, "type_": "Task" }, { "id": "T7", "start": 252, "end": 287, "type": 2, "type_": "Task" }, { "id": "T8", "start": 315, "end": 334, "type": 0, "type_": "Material" }, { "id": "T9", "start": 385, "end": 417, "type": 2, "type_": "Task" }, { "id": "T10", "start": 385, "end": 401, "type": 1, "type_": "Process" }, { "id": "T4", "start": 137, "end": 145, "type": 1, "type_": "Process" }, { "id": "T12", "start": 222, "end": 237, "type": 0, "type_": "Material" }, { "id": "T13", "start": 263, "end": 278, "type": 0, "type_": "Material" } ]
[]
S0021999115003939
The extrapolation of the upwind value required for TVD differencing is a particular hurdle for the application on unstructured meshes. As discussed in Section 3.2, two methods to extrapolate the value at the virtual upwind node, using data readily available on unstructured meshes, are considered. Given how the virtual upwind node is incorporated in the gradient ratio rf, the extrapolation method of Darwish and Moukalled [13] is referred to as implicit extrapolation and the method introduced by Ubbink and Issa [12] as explicit extrapolation. Both methods precisely reconstruct the upwind value for equidistant, rectilinear meshes but fail to do so on non-equidistant or non-rectilinear meshes, as discussed in Section 3.2. Using the explicit extrapolation method this issue can be rectified by imposing appropriate limits on the extrapolated upwind value.
[ { "id": "T1", "start": 4, "end": 37, "type": 2, "type_": "Task" }, { "id": "T2", "start": 402, "end": 423, "type": 1, "type_": "Process" }, { "id": "T3", "start": 447, "end": 469, "type": 1, "type_": "Process" }, { "id": "T4", "start": 478, "end": 514, "type": 1, "type_": "Process" }, { "id": "T5", "start": 523, "end": 545, "type": 1, "type_": "Process" }, { "id": "T6", "start": 312, "end": 372, "type": 1, "type_": "Process" }, { "id": "T7", "start": 738, "end": 760, "type": 1, "type_": "Process" }, { "id": "T8", "start": 114, "end": 133, "type": 0, "type_": "Material" }, { "id": "T9", "start": 235, "end": 280, "type": 0, "type_": "Material" }, { "id": "T10", "start": 603, "end": 634, "type": 0, "type_": "Material" }, { "id": "T11", "start": 675, "end": 697, "type": 0, "type_": "Material" }, { "id": "T12", "start": 656, "end": 671, "type": 0, "type_": "Material" }, { "id": "T13", "start": 51, "end": 67, "type": 1, "type_": "Process" } ]
[ { "arg1": "T2", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T12", "arg2": "T11", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T11", "arg2": "T12", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T5", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T2", "arg2": "T6", "relation": 2, "relation_": "Hyponym-of" } ]
S2214657115000052
The exponential relationships reported in the plots may be used to convert the dielectric values to air void values as prescribed in previous studies [1–3]. The AC pavement composite permittivity reduces, along with the reflection coefficient, as the volumetric proportion of air increases as compared to the remaining components. However, the method relies on an empirical fit, determined on a case-by-case basis, since the permittivity of the remaining components depends on the mix design (aggregate type, binder content, etc.). Long term studies in Finland concluded that this empirical fit is an exponential relationship [1]. The exponential fits, using a sufficient amount of cores, can be used to map the air void content variation in a similar manner to the dielectric maps shown in Fig. 4. Only 4 cores were feasible due to various factors involved with testing the final lift of an in-service pavement. More cores are needed for more stable exponential coefficients, although the limited cores show that the predicted relationships are similar for the measured dielectric range in this case-study. Since both regressions predict air void content at a maximum difference of 0.56%, which is within the uncertainty of the core measurement precision of 0.7%, use of either the initial or repeat run regression predictions are appropriate.
[ { "id": "T1", "start": 4, "end": 29, "type": 1, "type_": "Process" }, { "id": "T2", "start": 67, "end": 115, "type": 2, "type_": "Task" }, { "id": "T3", "start": 161, "end": 195, "type": 1, "type_": "Process" }, { "id": "T4", "start": 251, "end": 279, "type": 1, "type_": "Process" }, { "id": "T5", "start": 481, "end": 491, "type": 0, "type_": "Material" }, { "id": "T6", "start": 532, "end": 560, "type": 1, "type_": "Process" }, { "id": "T7", "start": 601, "end": 625, "type": 2, "type_": "Task" }, { "id": "T8", "start": 661, "end": 687, "type": 1, "type_": "Process" }, { "id": "T9", "start": 704, "end": 738, "type": 2, "type_": "Task" }, { "id": "T10", "start": 766, "end": 781, "type": 0, "type_": "Material" }, { "id": "T11", "start": 863, "end": 885, "type": 2, "type_": "Task" }, { "id": "T12", "start": 1018, "end": 1041, "type": 1, "type_": "Process" }, { "id": "T13", "start": 939, "end": 975, "type": 2, "type_": "Task" }, { "id": "T14", "start": 1062, "end": 1087, "type": 1, "type_": "Process" }, { "id": "T15", "start": 1119, "end": 1130, "type": 1, "type_": "Process" }, { "id": "T16", "start": 1139, "end": 1155, "type": 2, "type_": "Task" }, { "id": "T17", "start": 1229, "end": 1255, "type": 1, "type_": "Process" }, { "id": "T18", "start": 1283, "end": 1327, "type": 1, "type_": "Process" } ]
[]
S0370269304009086
The ART model is a hadronic transport model that includes baryons such as N, Δ(1232), N∗(1440), N∗(1535), Λ, Σ, and mesons such as π, ρ, ω, η, K, K∗. Both elastic and inelastic collisions among most of these particles are included by using the experimental data from hadron–hadron collisions. The ART model has been quite successful in explaining many experimental observations, including the surprisingly large kaon antiflow [11,12] in heavy ion collisions at AGS energies. The ART model also allows us to understand whether or not the strongly interacting matter formed in these collisions reaches chemical and/or thermal equilibrium. In the present study, we extend the ART model to include perturbatively the Ξ particle as in the studies for other rare particles using the transport model [6,13,14].
[ { "id": "T1", "start": 4, "end": 13, "type": 1, "type_": "Process" }, { "id": "T2", "start": 19, "end": 43, "type": 1, "type_": "Process" }, { "id": "T3", "start": 244, "end": 291, "type": 0, "type_": "Material" }, { "id": "T4", "start": 155, "end": 187, "type": 2, "type_": "Task" }, { "id": "T5", "start": 297, "end": 306, "type": 1, "type_": "Process" }, { "id": "T7", "start": 479, "end": 488, "type": 1, "type_": "Process" }, { "id": "T8", "start": 507, "end": 635, "type": 2, "type_": "Task" }, { "id": "T9", "start": 662, "end": 723, "type": 2, "type_": "Task" }, { "id": "T10", "start": 777, "end": 792, "type": 0, "type_": "Material" }, { "id": "T11", "start": 58, "end": 65, "type": 0, "type_": "Material" }, { "id": "T12", "start": 74, "end": 84, "type": 0, "type_": "Material" }, { "id": "T13", "start": 86, "end": 94, "type": 0, "type_": "Material" }, { "id": "T14", "start": 96, "end": 104, "type": 0, "type_": "Material" }, { "id": "T15", "start": 106, "end": 107, "type": 0, "type_": "Material" }, { "id": "T16", "start": 109, "end": 110, "type": 0, "type_": "Material" }, { "id": "T17", "start": 116, "end": 122, "type": 0, "type_": "Material" }, { "id": "T18", "start": 131, "end": 132, "type": 0, "type_": "Material" }, { "id": "T19", "start": 134, "end": 135, "type": 0, "type_": "Material" }, { "id": "T20", "start": 137, "end": 138, "type": 0, "type_": "Material" }, { "id": "T21", "start": 140, "end": 141, "type": 0, "type_": "Material" }, { "id": "T22", "start": 143, "end": 144, "type": 0, "type_": "Material" }, { "id": "T23", "start": 146, "end": 148, "type": 0, "type_": "Material" }, { "id": "T24", "start": 202, "end": 217, "type": 0, "type_": "Material" }, { "id": "T25", "start": 267, "end": 291, "type": 1, "type_": "Process" }, { "id": "T26", "start": 267, "end": 273, "type": 0, "type_": "Material" }, { "id": "T27", "start": 274, "end": 280, "type": 0, "type_": "Material" }, { "id": "T28", "start": 336, "end": 377, "type": 2, "type_": "Task" }, { "id": "T6", "start": 412, "end": 425, "type": 1, "type_": "Process" }, { "id": "T29", "start": 437, "end": 457, "type": 1, "type_": "Process" }, { "id": "T30", "start": 412, "end": 416, "type": 0, "type_": "Material" }, { "id": "T31", "start": 437, "end": 446, "type": 0, "type_": "Material" }, { "id": "T32", "start": 461, "end": 473, "type": 1, "type_": "Process" }, { "id": "T33", "start": 537, "end": 564, "type": 0, "type_": "Material" }, { "id": "T34", "start": 581, "end": 591, "type": 1, "type_": "Process" }, { "id": "T35", "start": 673, "end": 682, "type": 1, "type_": "Process" }, { "id": "T36", "start": 713, "end": 723, "type": 0, "type_": "Material" }, { "id": "T37", "start": 752, "end": 766, "type": 0, "type_": "Material" } ]
[ { "arg1": "T1", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T12", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T13", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T14", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T15", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T16", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T18", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T19", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T20", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T21", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T22", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T23", "arg2": "T17", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T36", "arg2": "T37", "relation": 2, "relation_": "Hyponym-of" } ]
S0045782513001473
In this paper, however, we prefer the simpler ‘framed’ cell employed by Hadjiconstantinou and Patera [16], where the shear stress is generated by constraining the velocity in a ‘frame’ rather than by modifying the shape of the box. The framed cell is periodic, but we cannot simply calculate the average stress in the whole box because the presence of an external buffer would produce spurious results. We need the local stress in the core region, but this complicates the Oij term in Eq. (3). There are other methods to calculate the stress tensor such as the method of planes [32], the volume-average approach [26,14], or the method derived from the Schweitz virial relation [25], but, in general, we must choose between a complicated computational cell (i.e. Lees–Edwards cell) and simplifying the calculation of the momentum flux, or a simple cell (i.e. framed cell) and complicating the calculation of the momentum flux. The new method we propose here does not need the direct calculation of the flux, so it avoids this issue altogether: we can use the framed cell and, at the same time, avoid the calculation of the IK equation.
[ { "id": "T1", "start": 521, "end": 548, "type": 2, "type_": "Task" }, { "id": "T2", "start": 47, "end": 59, "type": 1, "type_": "Process" }, { "id": "T3", "start": 117, "end": 129, "type": 1, "type_": "Process" }, { "id": "T4", "start": 146, "end": 171, "type": 1, "type_": "Process" }, { "id": "T5", "start": 1058, "end": 1069, "type": 1, "type_": "Process" }, { "id": "T6", "start": 236, "end": 247, "type": 1, "type_": "Process" }, { "id": "T7", "start": 561, "end": 577, "type": 1, "type_": "Process" }, { "id": "T8", "start": 521, "end": 548, "type": 1, "type_": "Process" }, { "id": "T9", "start": 588, "end": 611, "type": 1, "type_": "Process" }, { "id": "T10", "start": 652, "end": 676, "type": 1, "type_": "Process" }, { "id": "T11", "start": 725, "end": 755, "type": 1, "type_": "Process" }, { "id": "T12", "start": 762, "end": 779, "type": 1, "type_": "Process" }, { "id": "T13", "start": 840, "end": 851, "type": 1, "type_": "Process" }, { "id": "T14", "start": 858, "end": 869, "type": 1, "type_": "Process" }, { "id": "T15", "start": 1122, "end": 1133, "type": 1, "type_": "Process" } ]
[ { "arg1": "T3", "arg2": "T4", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T7", "arg2": "T8", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T9", "arg2": "T8", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T10", "arg2": "T8", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T11", "arg2": "T8", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T12", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T14", "arg2": "T13", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T13", "arg2": "T8", "relation": 2, "relation_": "Hyponym-of" } ]
S221450951400031X
This phase was completed in 2005. Previous contracts had been procured with the contractor providing the detailed design. For this system the design was undertaken by Mott MacDonald. It was developed by looking at the systems installed previously and calculating what was actually required to achieve cathodic protection of the piers. This resulted in a significant reduction in the number of zones and monitoring probes. The varying amounts of steelwork in the beams had previously lead to up to 5 zones per beam, with multiple layers of mesh to achieve the design current density. On review of the data the operating current density was similar in all zones and so this was reduced to a single zone per beam. The encapsulation was susceptible to ASR and contained post tensioning and so it was decided to use a galvanic system based on Galvashield CC anodes from Fosroc. Our design did not include an option to allow depolarization of the galvanic system, but the contractor supplied one, such that the anodes could be remotely disconnected. The control unit was from Electrotech CP and operated via a broadband connection provided by the contractor.
[ { "id": "T1", "start": 105, "end": 120, "type": 0, "type_": "Material" }, { "id": "T2", "start": 203, "end": 262, "type": 1, "type_": "Process" }, { "id": "T3", "start": 218, "end": 246, "type": 0, "type_": "Material" }, { "id": "T4", "start": 354, "end": 420, "type": 1, "type_": "Process" }, { "id": "T5", "start": 426, "end": 454, "type": 1, "type_": "Process" }, { "id": "T6", "start": 520, "end": 543, "type": 1, "type_": "Process" }, { "id": "T8", "start": 600, "end": 604, "type": 0, "type_": "Material" }, { "id": "T10", "start": 301, "end": 333, "type": 2, "type_": "Task" }, { "id": "T11", "start": 715, "end": 728, "type": 1, "type_": "Process" }, { "id": "T12", "start": 813, "end": 828, "type": 0, "type_": "Material" }, { "id": "T14", "start": 919, "end": 956, "type": 2, "type_": "Task" }, { "id": "T15", "start": 1104, "end": 1124, "type": 1, "type_": "Process" }, { "id": "T7", "start": 748, "end": 751, "type": 1, "type_": "Process" }, { "id": "T9", "start": 838, "end": 859, "type": 0, "type_": "Material" }, { "id": "T13", "start": 941, "end": 956, "type": 0, "type_": "Material" } ]
[]
S2212671612001618
This paper presents a non-fragile controller design method based on system quadratic performance optimization. For the additive controller gain variations, the necessary and sufficient conditions for the existence of non-fragile state feedback controller are given and transformed to the LMI problems, which simplifies the solutions to obtain non-fragile state feedback controllers. The flight control simulation results prove the reliability and validity of the method.
[ { "id": "T1", "start": 22, "end": 58, "type": 1, "type_": "Process" }, { "id": "T2", "start": 68, "end": 109, "type": 1, "type_": "Process" }, { "id": "T3", "start": 119, "end": 154, "type": 1, "type_": "Process" }, { "id": "T4", "start": 217, "end": 254, "type": 1, "type_": "Process" }, { "id": "T5", "start": 288, "end": 300, "type": 2, "type_": "Task" }, { "id": "T6", "start": 269, "end": 280, "type": 1, "type_": "Process" }, { "id": "T7", "start": 308, "end": 332, "type": 1, "type_": "Process" }, { "id": "T8", "start": 343, "end": 381, "type": 1, "type_": "Process" }, { "id": "T9", "start": 387, "end": 412, "type": 1, "type_": "Process" } ]
[]
S1875952116300209
In order to test whether haptic patterns can convey or enhance the mood music of a movie, an affective movie clip corpus was required consisting of clips labeled according to the emotion conveyed in the mood music. The following database collections were examined as possible sources for the corpus: the Emotional Movie Database (EMDB) [19], and Film Stim [20]. However, these were discarded after review as unsuitable. The aim of this study is to enhance the mood in the film score, and in the case of the clips in the EMDB, no audio is provided which deemed the clips unsuitable. In the case of the Film Stim database, the clips are in French rather than English, and with no subtitles which where also deemed unsuitable since the studies are carried out with English speaking participants. Furthermore, the Film Stim selection is based on the affective content of the narrative as in most of them there is no music which is also unsuitable as discussed. From our review of available database collections, it was found that at present there is no standard corpus of affective movie clips where the affective indexing referred to the musical score of the clip.
[ { "id": "T1", "start": 93, "end": 120, "type": 0, "type_": "Material" }, { "id": "T2", "start": 304, "end": 328, "type": 0, "type_": "Material" }, { "id": "T3", "start": 330, "end": 334, "type": 0, "type_": "Material" }, { "id": "T4", "start": 346, "end": 355, "type": 0, "type_": "Material" }, { "id": "T5", "start": 12, "end": 77, "type": 2, "type_": "Task" }, { "id": "T6", "start": 229, "end": 249, "type": 0, "type_": "Material" }, { "id": "T7", "start": 398, "end": 404, "type": 1, "type_": "Process" }, { "id": "T8", "start": 448, "end": 464, "type": 2, "type_": "Task" }, { "id": "T9", "start": 520, "end": 524, "type": 0, "type_": "Material" }, { "id": "T10", "start": 601, "end": 619, "type": 0, "type_": "Material" }, { "id": "T11", "start": 762, "end": 791, "type": 0, "type_": "Material" }, { "id": "T12", "start": 810, "end": 829, "type": 0, "type_": "Material" }, { "id": "T13", "start": 966, "end": 1006, "type": 1, "type_": "Process" } ]
[ { "arg1": "T3", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T6", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T4", "arg2": "T6", "relation": 2, "relation_": "Hyponym-of" } ]
S2212667814001294
Knowledge Management (KM) is one of the hotspots for research in the past decade. In most cases, the number of users in a Knowledge Management System (KMS) is very large, and they are from varied departments, even other companies. In this paper, some defects when existing methods about access control and recommendation are deployed in KMS are analyzed to show that these widely-used approaches need to be extended. To overcome the deficiencies of previous work, this paper proposes an extended Role-Based Access Control (RBAC) method and a hybrid recommendation approach for Knowledge Management System. Also, a real-life system is presented to verify the proposed methodology.
[ { "id": "T1", "start": 0, "end": 20, "type": 2, "type_": "Task" }, { "id": "T2", "start": 22, "end": 24, "type": 2, "type_": "Task" }, { "id": "T3", "start": 122, "end": 149, "type": 1, "type_": "Process" }, { "id": "T4", "start": 151, "end": 154, "type": 1, "type_": "Process" }, { "id": "T5", "start": 251, "end": 353, "type": 2, "type_": "Task" }, { "id": "T6", "start": 407, "end": 415, "type": 1, "type_": "Process" }, { "id": "T7", "start": 496, "end": 520, "type": 1, "type_": "Process" }, { "id": "T8", "start": 523, "end": 527, "type": 1, "type_": "Process" }, { "id": "T9", "start": 542, "end": 572, "type": 1, "type_": "Process" }, { "id": "T10", "start": 577, "end": 604, "type": 1, "type_": "Process" }, { "id": "T11", "start": 614, "end": 630, "type": 1, "type_": "Process" }, { "id": "T12", "start": 647, "end": 678, "type": 2, "type_": "Task" }, { "id": "T13", "start": 122, "end": 142, "type": 2, "type_": "Task" }, { "id": "T14", "start": 337, "end": 340, "type": 1, "type_": "Process" } ]
[ { "arg1": "T1", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T1", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T7", "arg2": "T8", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T8", "arg2": "T7", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" } ]
S0010938X12002508
The study outlines a trial of transient response analysis on full-scale motorway bridge structures to obtain information concerning the steel–concrete interface and is part of a larger study to assess the long-term sustained benefits offered by Impressed Current Cathodic Protection (ICCP) after the interruption of the protective current [1]. These structures had previously been protected for 5–16years by an ICCP system prior to the start of the study. The protective current was interrupted, in order to assess the long-term benefits provided by ICCP after it has been turned off. This paper develops and examines a simplified approach for the on-site use of transient response analysis and discusses the potential advantages of the technique as a tool for the assessment of the corrosion condition of steel in reinforced concrete structures.
[ { "id": "T1", "start": 30, "end": 57, "type": 2, "type_": "Task" }, { "id": "T2", "start": 136, "end": 160, "type": 2, "type_": "Task" }, { "id": "T3", "start": 205, "end": 233, "type": 1, "type_": "Process" }, { "id": "T4", "start": 300, "end": 338, "type": 1, "type_": "Process" }, { "id": "T6", "start": 460, "end": 494, "type": 1, "type_": "Process" }, { "id": "T7", "start": 508, "end": 537, "type": 1, "type_": "Process" }, { "id": "T8", "start": 596, "end": 639, "type": 2, "type_": "Task" }, { "id": "T9", "start": 765, "end": 802, "type": 2, "type_": "Task" }, { "id": "T10", "start": 695, "end": 729, "type": 1, "type_": "Process" }, { "id": "T11", "start": 72, "end": 98, "type": 0, "type_": "Material" }, { "id": "T12", "start": 245, "end": 282, "type": 1, "type_": "Process" }, { "id": "T13", "start": 284, "end": 288, "type": 1, "type_": "Process" }, { "id": "T14", "start": 411, "end": 422, "type": 1, "type_": "Process" }, { "id": "T5", "start": 550, "end": 554, "type": 1, "type_": "Process" }, { "id": "T15", "start": 663, "end": 690, "type": 1, "type_": "Process" }, { "id": "T16", "start": 806, "end": 811, "type": 0, "type_": "Material" }, { "id": "T17", "start": 815, "end": 845, "type": 0, "type_": "Material" } ]
[ { "arg1": "T4", "arg2": "T6", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T6", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T7", "arg2": "T3", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T7", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T13", "arg2": "T12", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T12", "arg2": "T13", "relation": 1, "relation_": "Synonym-of" } ]
S0022311513001165
Our simulations confirm experimental observations that W net erosion represents only tiny fraction (in our simulation ∼1%) of the W gross erosion. The estimated upstream W fluxes, FWupstrem, are in good agreement with the experimentally observed values ⩽1019m-2s-1 [16]. Moreover, this value is not very sensitive to the divertor plasma temperature. For low temperatures the energy of D and C ions hitting to the divertor plates is too low to sputter sufficient amount of W. With increasing energy the W sputtering increases, but the potential drop in the divertor plasma increases too. As a result, most of the W atoms are ionized in the vicinity of the divertor and return back to the plates. There are two effects leading to the observed prompt redeposition of W ions: first is the “near-divertor” ionization of W due to low ionization potential −7.86eV (for comparison the ionization potentials for D and C are 13,6 and 10.6eV), second, W+n ions have large Larmor radius ∼2/nmm, so that they are redeposited within the distance of a Larmor radius. Important to note that a significant fraction of W ions escaping this prompt redeposition are returned back due to the friction with the main ions.
[ { "id": "T1", "start": 4, "end": 15, "type": 0, "type_": "Material" }, { "id": "T2", "start": 55, "end": 68, "type": 2, "type_": "Task" }, { "id": "T3", "start": 130, "end": 145, "type": 2, "type_": "Task" }, { "id": "T4", "start": 151, "end": 178, "type": 1, "type_": "Process" }, { "id": "T5", "start": 222, "end": 252, "type": 0, "type_": "Material" }, { "id": "T6", "start": 398, "end": 428, "type": 1, "type_": "Process" }, { "id": "T7", "start": 480, "end": 497, "type": 1, "type_": "Process" }, { "id": "T9", "start": 624, "end": 631, "type": 1, "type_": "Process" }, { "id": "T10", "start": 668, "end": 674, "type": 1, "type_": "Process" }, { "id": "T11", "start": 732, "end": 760, "type": 2, "type_": "Task" }, { "id": "T12", "start": 786, "end": 811, "type": 1, "type_": "Process" }, { "id": "T13", "start": 955, "end": 974, "type": 1, "type_": "Process" }, { "id": "T14", "start": 1171, "end": 1198, "type": 1, "type_": "Process" }, { "id": "T15", "start": 385, "end": 397, "type": 0, "type_": "Material" }, { "id": "T16", "start": 180, "end": 189, "type": 1, "type_": "Process" }, { "id": "T17", "start": 472, "end": 473, "type": 0, "type_": "Material" }, { "id": "T18", "start": 170, "end": 178, "type": 0, "type_": "Material" }, { "id": "T19", "start": 502, "end": 503, "type": 0, "type_": "Material" }, { "id": "T20", "start": 556, "end": 571, "type": 0, "type_": "Material" }, { "id": "T21", "start": 815, "end": 816, "type": 0, "type_": "Material" }, { "id": "T22", "start": 612, "end": 619, "type": 0, "type_": "Material" }, { "id": "T23", "start": 903, "end": 904, "type": 0, "type_": "Material" }, { "id": "T24", "start": 1101, "end": 1107, "type": 0, "type_": "Material" }, { "id": "T25", "start": 909, "end": 910, "type": 0, "type_": "Material" }, { "id": "T26", "start": 941, "end": 949, "type": 0, "type_": "Material" }, { "id": "T8", "start": 502, "end": 514, "type": 1, "type_": "Process" }, { "id": "T27", "start": 828, "end": 838, "type": 1, "type_": "Process" }, { "id": "T28", "start": 413, "end": 428, "type": 0, "type_": "Material" }, { "id": "T29", "start": 764, "end": 770, "type": 0, "type_": "Material" }, { "id": "T30", "start": 655, "end": 663, "type": 0, "type_": "Material" }, { "id": "T31", "start": 687, "end": 693, "type": 0, "type_": "Material" }, { "id": "T32", "start": 321, "end": 329, "type": 0, "type_": "Material" }, { "id": "T33", "start": 877, "end": 887, "type": 1, "type_": "Process" } ]
[ { "arg1": "T16", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T16", "relation": 1, "relation_": "Synonym-of" } ]
S0022311515002664
The primary benefit of using a 3D model is that it allows the application of anisotropic material properties. As a hexagonal close packed lattice structure, a single zirconium grain is plastically anisotropic due to the difficulty of activating slip with a 〈c〉 component [23–26]. Abaqus allows this to be represented by setting plasticity potential ratios. The anisotropic elastic and plastic constants are shown in Table 1. Zirconium alloys can often have a bimodal basal pole distribution, with a tilt on the basal normal or c direction of ±30° in the normal direction being quoted for recrystallized Zircaloy-4 [27,28]. However, for simplicity the basal normal or c direction has been taken as being parallel to the normal direction. As such the 1, 2 and 3 directions in Table 1 correlate with the X, Y and Z global coordinate system for the 3D simulations, with the 3 direction correlating to the c direction of a zirconium unit lattice. Table 1 also shows the elastic properties incorporated into the simulations. The oxide layer has been simulated as a purely elastic material. Although it is known that the oxide is strongly textured [29], it is still simulated as a homogenous solid therefore isotropic material properties have been used for the oxide in all simulations.
[ { "id": "T1", "start": 31, "end": 39, "type": 1, "type_": "Process" }, { "id": "T2", "start": 77, "end": 108, "type": 2, "type_": "Task" }, { "id": "T3", "start": 280, "end": 286, "type": 1, "type_": "Process" }, { "id": "T4", "start": 328, "end": 355, "type": 1, "type_": "Process" }, { "id": "T5", "start": 459, "end": 490, "type": 1, "type_": "Process" }, { "id": "T6", "start": 1201, "end": 1230, "type": 2, "type_": "Task" }, { "id": "T7", "start": 361, "end": 402, "type": 0, "type_": "Material" }, { "id": "T8", "start": 166, "end": 181, "type": 0, "type_": "Material" }, { "id": "T9", "start": 425, "end": 434, "type": 0, "type_": "Material" }, { "id": "T10", "start": 588, "end": 613, "type": 0, "type_": "Material" }, { "id": "T11", "start": 845, "end": 859, "type": 1, "type_": "Process" }, { "id": "T12", "start": 918, "end": 940, "type": 0, "type_": "Material" }, { "id": "T13", "start": 1006, "end": 1017, "type": 1, "type_": "Process" }, { "id": "T14", "start": 1023, "end": 1034, "type": 0, "type_": "Material" }, { "id": "T15", "start": 1059, "end": 1082, "type": 0, "type_": "Material" }, { "id": "T16", "start": 1114, "end": 1119, "type": 0, "type_": "Material" }, { "id": "T17", "start": 1254, "end": 1259, "type": 0, "type_": "Material" }, { "id": "T18", "start": 1267, "end": 1278, "type": 1, "type_": "Process" } ]
[]
S0370269304009347
The presence of chaotic motion in nuclear systems has been firmly related with the statistics of high-lying energy levels [8,9]. Poisson distributions of normalized spacings of successive nuclear or atomic excited levels with the same spin and parity correspond to integrable classical dynamics, while Wigner's statistics signal chaotic motion in the corresponding classical regime [10]. Intermediate situations are more difficult to assess. Very recently a proposal has been made to treat the spectral fluctuations δn as discrete time series [11]. Defining (1)δn=∫−∞En+1ρ˜(E)dE−n, with ρ˜(E) the mean level density which allows the mapping to dimensionless levels with unitary average level density, and analyzing the energy fluctuations as a discrete time series, they found that nuclear power spectra behave like 1f noise, postulating that this might be a characteristic signature of generic quantum chaotic systems. In the present work we implement this idea, using the 1f spectral behavior as a test for the presence of chaos in nuclear mass errors.
[ { "id": "T1", "start": 16, "end": 49, "type": 2, "type_": "Task" }, { "id": "T2", "start": 83, "end": 121, "type": 1, "type_": "Process" }, { "id": "T3", "start": 129, "end": 150, "type": 1, "type_": "Process" }, { "id": "T4", "start": 206, "end": 220, "type": 1, "type_": "Process" }, { "id": "T5", "start": 235, "end": 239, "type": 1, "type_": "Process" }, { "id": "T6", "start": 244, "end": 250, "type": 1, "type_": "Process" }, { "id": "T7", "start": 265, "end": 294, "type": 1, "type_": "Process" }, { "id": "T8", "start": 302, "end": 321, "type": 0, "type_": "Material" }, { "id": "T9", "start": 329, "end": 343, "type": 1, "type_": "Process" }, { "id": "T10", "start": 365, "end": 381, "type": 2, "type_": "Task" }, { "id": "T11", "start": 494, "end": 515, "type": 1, "type_": "Process" }, { "id": "T12", "start": 516, "end": 518, "type": 1, "type_": "Process" }, { "id": "T13", "start": 522, "end": 542, "type": 1, "type_": "Process" }, { "id": "T17", "start": 633, "end": 664, "type": 1, "type_": "Process" }, { "id": "T18", "start": 670, "end": 699, "type": 1, "type_": "Process" }, { "id": "T19", "start": 705, "end": 738, "type": 1, "type_": "Process" }, { "id": "T20", "start": 744, "end": 764, "type": 1, "type_": "Process" }, { "id": "T21", "start": 782, "end": 803, "type": 2, "type_": "Task" }, { "id": "T22", "start": 895, "end": 918, "type": 2, "type_": "Task" }, { "id": "T23", "start": 974, "end": 994, "type": 1, "type_": "Process" }, { "id": "T24", "start": 1025, "end": 1053, "type": 1, "type_": "Process" } ]
[ { "arg1": "T12", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T13", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" } ]
S0165168416300603
Despite the fact that SRC-HE reduces the number of FEs, audio measurements extraction based on SRC would still be not suitable for real-time applications [39]. The previous SRC-HE module is then replaced by the generalised cross correlation phase transform (GCC-PHAT) introduced in Section 2.1, as this does not involve cumbersome point function estimations. The drawback is that the basic GCC algorithm can only detect one source at a time and it is known to be sensitive to room reverberations [5], however it is still effective under moderate reverberant environments (T60≈0.3s) [40]. For these reasons, at first experiments where only a speaker is active at any given time are carried out, as it often happens in a polite conversation between two or more people. Speech segments using a voice activity detector (VAD) [41] are further extracted and processed using a GCC-PHAT step, for the signal to be more robust to reverberations. Thus, the measure vector obtained za (see Section 2.1) can now be rewritten as za={τm(t)}, where each component τm is the TDOA collected at the m-th microphone pair at each time step t. Since TDOAs are not linear in the speaker position, they must be input into an extended Kalman filter (EKF), as in [10] to get an audio position estimation.
[ { "id": "T1", "start": 211, "end": 256, "type": 1, "type_": "Process" }, { "id": "T2", "start": 258, "end": 266, "type": 1, "type_": "Process" }, { "id": "T3", "start": 641, "end": 648, "type": 1, "type_": "Process" }, { "id": "T4", "start": 791, "end": 814, "type": 1, "type_": "Process" }, { "id": "T5", "start": 816, "end": 819, "type": 1, "type_": "Process" }, { "id": "T6", "start": 870, "end": 883, "type": 1, "type_": "Process" }, { "id": "T7", "start": 1081, "end": 1101, "type": 1, "type_": "Process" }, { "id": "T8", "start": 1211, "end": 1224, "type": 1, "type_": "Process" }, { "id": "T9", "start": 1226, "end": 1229, "type": 1, "type_": "Process" }, { "id": "T10", "start": 1253, "end": 1278, "type": 2, "type_": "Task" }, { "id": "T11", "start": 767, "end": 782, "type": 0, "type_": "Material" }, { "id": "T12", "start": 22, "end": 28, "type": 1, "type_": "Process" }, { "id": "T13", "start": 390, "end": 403, "type": 1, "type_": "Process" } ]
[ { "arg1": "T1", "arg2": "T2", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T2", "arg2": "T1", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T4", "arg2": "T5", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T5", "arg2": "T4", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T8", "arg2": "T9", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T9", "arg2": "T8", "relation": 1, "relation_": "Synonym-of" } ]
S088523081530036X
Note that the presented architecture works at the frame level, meaning that each single frame (plus its corresponding context) is fed-forward through the network, obtaining a class posterior probability for all of the target languages. This fact makes the DNNs particularly suitable for real-time applications because, unlike other approaches (i.e. i-vectors), we can potentially make a decision about the language at each new frame. Indeed, at each frame, we can combine the evidence from past frames to get a single similarity score between the test utterance and the targetlanguages. A simple way of doing this combination is to assume that frames are independent and multiply the posterior estimates of the last layer. The score sl for language l of a given test utterance is computed by multiplying the output probabilities pl obtained for all of its frames; or equivalently, accumulating the logs as:(6)sl=1N∑t=1Nlogp(Ll|xt​, θ)where p(Ll|xt​, θ) represents the class probability output for the language l corresponding to the input example at time t, xt by using the DNN defined by parameters θ.
[ { "id": "T1", "start": 256, "end": 260, "type": 1, "type_": "Process" }, { "id": "T2", "start": 326, "end": 342, "type": 1, "type_": "Process" }, { "id": "T3", "start": 349, "end": 358, "type": 1, "type_": "Process" }, { "id": "T4", "start": 218, "end": 234, "type": 0, "type_": "Material" }, { "id": "T5", "start": 476, "end": 501, "type": 0, "type_": "Material" }, { "id": "T6", "start": 762, "end": 776, "type": 0, "type_": "Material" }, { "id": "T7", "start": 1074, "end": 1077, "type": 1, "type_": "Process" }, { "id": "T8", "start": 464, "end": 501, "type": 1, "type_": "Process" }, { "id": "T9", "start": 632, "end": 721, "type": 1, "type_": "Process" }, { "id": "T10", "start": 792, "end": 862, "type": 1, "type_": "Process" }, { "id": "T11", "start": 881, "end": 934, "type": 1, "type_": "Process" } ]
[ { "arg1": "T10", "arg2": "T11", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T11", "arg2": "T10", "relation": 1, "relation_": "Synonym-of" }, { "arg1": "T3", "arg2": "T2", "relation": 2, "relation_": "Hyponym-of" } ]
S2214509515300103
Another important reason for the damages incurred by the RC buildings is workmanship defects. It is understood that granulometry of the handmade concretes was not in compliance with the standards since the aggregate utilized in them was not sieved. Also the compaction process was not properly implemented in general in the installment of concrete in RC buildings. This situation resulted in the concrete to exhibit an excessively porous structure. The most fundamental rules of thumb of construction, namely concrete cover, was not taken care of in formwork workmanship. Faults in the connections of stirrups to the longitudinal bars, unstaggered formation of stirrup hooks in beams and columns, the perpendicular angles of the hooks, inadequately anchorage lengths of the stirrup hooks and longitudinal bars, and the use of cold joints were the other frequently encountered workmanship defects (Figs. 19–22).
[ { "id": "T1", "start": 73, "end": 92, "type": 1, "type_": "Process" }, { "id": "T2", "start": 116, "end": 154, "type": 1, "type_": "Process" }, { "id": "T3", "start": 258, "end": 276, "type": 1, "type_": "Process" }, { "id": "T4", "start": 324, "end": 347, "type": 1, "type_": "Process" }, { "id": "T5", "start": 509, "end": 523, "type": 0, "type_": "Material" }, { "id": "T6", "start": 572, "end": 609, "type": 1, "type_": "Process" }, { "id": "T7", "start": 636, "end": 668, "type": 1, "type_": "Process" }, { "id": "T8", "start": 701, "end": 721, "type": 1, "type_": "Process" }, { "id": "T9", "start": 736, "end": 766, "type": 1, "type_": "Process" }, { "id": "T10", "start": 819, "end": 837, "type": 1, "type_": "Process" }, { "id": "T11", "start": 876, "end": 895, "type": 1, "type_": "Process" }, { "id": "T12", "start": 396, "end": 404, "type": 0, "type_": "Material" }, { "id": "T13", "start": 774, "end": 787, "type": 0, "type_": "Material" }, { "id": "T14", "start": 792, "end": 809, "type": 0, "type_": "Material" }, { "id": "T15", "start": 669, "end": 674, "type": 0, "type_": "Material" }, { "id": "T16", "start": 678, "end": 683, "type": 0, "type_": "Material" }, { "id": "T17", "start": 688, "end": 695, "type": 0, "type_": "Material" }, { "id": "T18", "start": 729, "end": 734, "type": 0, "type_": "Material" } ]
[ { "arg1": "T6", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T7", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T8", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T9", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" }, { "arg1": "T10", "arg2": "T11", "relation": 2, "relation_": "Hyponym-of" } ]
S0375960113004568
The length effect is always important in nanodevices. So we investigate the length dependence of electronic transport properties in M3 by increasing the number of carbon unit cells in the scattering region. Here we present the transport results when the numbers of carbon unit cells in the scattering region are 10 and 12, which are called M4 and M5, respectively. The current–voltage characteristics shown in Fig. 8. We can see that the large rectifying ratio still can be observed irrespective of the length of heterojunctions. This is due to the fact that the electronic transport properties for M3 are mainly determined by the parity of the π and π⁎ subbands of left and right electrodes. Thees results indicate that the lengths of the two parts in the scattering regions have no affects on the qualitative charge transport in M3.
[ { "id": "T1", "start": 41, "end": 52, "type": 0, "type_": "Material" }, { "id": "T2", "start": 4, "end": 17, "type": 2, "type_": "Task" }, { "id": "T3", "start": 76, "end": 128, "type": 2, "type_": "Task" }, { "id": "T4", "start": 138, "end": 180, "type": 1, "type_": "Process" }, { "id": "T5", "start": 163, "end": 180, "type": 0, "type_": "Material" }, { "id": "T6", "start": 265, "end": 282, "type": 0, "type_": "Material" }, { "id": "T7", "start": 369, "end": 400, "type": 1, "type_": "Process" }, { "id": "T8", "start": 438, "end": 460, "type": 1, "type_": "Process" }, { "id": "T9", "start": 563, "end": 594, "type": 2, "type_": "Task" }, { "id": "T10", "start": 799, "end": 827, "type": 2, "type_": "Task" }, { "id": "T11", "start": 645, "end": 662, "type": 1, "type_": "Process" } ]
[]