首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Fiber spinning and mechanical properties of four rigid polyesters with alkoxy substituents of different length and placement were evaluated. Properties of oriented fibers from the polymer with dodecyloxy substituents on the terephthalate moiety, PTA12HQ, were significantly affected by the crystal modification. At room temperature the following properties (tensile modulus E, tensile strength σb, and strain at break εb) could be obtained: E = 9.5 GPa, σb = 85 MPa and εb = 1.1% for phase Lf (the “frozen in layered mesophase”); E = 10.4 GPa, σb = 59 MPa and εb = 0.6% for modification A; E = 17.3 GPa, σb = 158 MPa and εb = 1.2% for modification B. Because of the higher amount of main chains per cross sectional area the polymer with hexyloxy side chains, PTA6HQ, showed better properties at a comparable degree of molecular orientation: E = 24 GPa, σb = 270 MPa, εb = 1.4%. Fibers obtained from the polyester with dodecyloxy substituents on the hydroquinone moiety, PTAHQ12, were too brittle to handle. The polyester with dodecyloxy substituents on both moieties, PTA12HQ12, was spun from the isotropic melt. Because of the obtained low degree of orientation, properties (E = 1 GPa, σb = 40 MPa, and εb = 6.3%) were governed by interactions between the chains (the main chains are not load-bearing). © 1993 John Wiley & Sons, Inc.  相似文献   

2.
The hydrogen bond (H‐bond) energies are evaluated for 18 molecular crystals with 28 moderate and strong O? H···O bonds using the approaches based on the electron density properties, which are derived from the B3LYP/6‐311G** calculations with periodic boundary conditions. The approaches considered explore linear relationships between the local electronic kinetic Gb and potential Vb densities at the H···O bond critical point and the H‐bond energy EHB. Comparison of the computed EHB values with the experimental data and enthalpies evaluated using the empirical correlation of spectral and thermodynamic parameters (Iogansen, Spectrochim. Acta Part A 1999 , 55, 1585) enables to estimate the accuracy and applicability limits of the approaches used. The Vb?EHB approach overestimates the energy of moderate H‐bonds (EHB < 60 kJ/mol) by ~20% and gives unreliably high energies for crystals with strong H‐bonds. On the other hand, the Gb?EHB approach affords reliable results for the crystals under consideration. The linear relationship between Gb and EHB is basis set superposition error (BSSE) free and allows to estimate the H‐bond energy without computing it by means of the supramolecular approach. Therefore, for the evaluation of H‐bond energies in molecular crystals, the Gb value can be recommended to be obtained from both density functional theory (DFT) computations with periodic boundary conditions and precise X‐ray diffraction experiments. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
The main purpose of this paper is to prove the applicability of the mechanism of congruent dissociative vaporization (CDV) to the solid-state decomposition kinetics through the comparison of the fundamental theoretical relationship Ei/Ee=(a+b)/a resulted from this mechanism with experiment. It has been shown that the ratios of Ei and Ee parameters of the Arrhenius equation measured in the isobaric and equimolar modes (in the presence and absence of H2O vapour) for 22 reactants with the general formula aSalt⋅bH2O or aOxide⋅bH2O are in agreement with the values of (a+b)/a. The relative standard deviation is only 17% and the correlation coefficient is close to 0.99. A probability of accidental correlation for all set of the E parameters taken from the literature is lower than 4⋅10–16 . This strongly supports the validity of the CDV mechanism. The problem of stability of polyatomic molecules of inorganic salts in the gaseous state, which are the primary decomposition products of crystalline hydrates, was also discussed on the basis of recent mass spectroscopy studies. It was concluded that any doubts in the applicability of the CDV mechanism as a general mechanism of solid-state decomposition reactions are unsound.  相似文献   

4.
Previous theory for pressure effects on conduction in amorphous macromolecular solids is extended to include pressure effects on hyperelectronic polarization. A theoretical estimate of hopping activation energies is given. Experimental studies of the frequency effects of the pressure coefficients of conduction and permittivity permits resolution of the salient factors. Carrier mobility in ekaconjugated polymers is best represented by activated hopping, for conductivity increase with frequency, and permittivity drops. Conductivity and hyperelectronic polarization increase markedly with pressure. Their activation energies, Ea0, and the pressure coefficients of their activation energies (bE + bH) are nearly identical. Ea0 is frequency-dependent, but the (bE + bH) terms are not, showing the frequency dependence of Ea0 to lie in the hopping activation enthalpy. Similar arguments on the “total” pressure coefficients b and b for conduction and polarization show the latter coefficients to have a frequency dependence lying in the hopping activation entropy.  相似文献   

5.
A combinatorial chemistry approach has been used to synthesize an array of Schiff bases, five of which, namely N‐[(E,2E)‐3‐(4‐methoxy­phenyl)‐2‐propenyl­idene]‐3‐nitro­aniline, C16H14N2O3, (1a), N‐[(E,2E)‐3‐(4‐methoxy­phenyl)‐2‐propenyl­idene]‐4‐nitro­aniline, C16H14N2O3, (2a), N‐{(E,2E)‐3‐[4‐(di­methyl­amino)­phenyl]‐2‐propenyl­idene}‐3‐nitro­aniline, C17H17N3O2, (1b), N‐{(E,2E)‐3‐[4‐(di­methyl­amino)­phenyl]‐2‐propenyl­idene}‐4‐nitro­aniline, C17H17N3O2, (2b), and N‐{(E,2E)‐3‐[4‐(di­methyl­amino)­phenyl]‐2‐propenyl­idene}‐2‐methyl‐4‐nitro­aniline, C18H19N3O2, (3b), have been structurally characterized. A stack structure is observed for (1a) and (1b) in the crystal phase. Experimental and calculated molecular structures are discussed for these compounds which belong to a chemical class having potential applications as non‐linear optical materials.  相似文献   

6.
To examine the roles of competing intermolecular interactions in differentiating the molecular packing arrangements of some isomeric phenylhydrazones from each other, the crystal structures of five nitrile–halogen substituted phenylhydrazones and two nitro–halogen substituted phenylhydrazones have been determined and are described here: (E)‐4‐cyanobenzaldehyde 4‐chlorophenylhydrazone, C14H10ClN3, (Ia); (E)‐4‐cyanobenzaldehyde 4‐bromophenylhydrazone, C14H10BrN3, (Ib); (E)‐4‐cyanobenzaldehyde 4‐iodophenylhydrazone, C14H10IN3, (Ic); (E)‐4‐bromobenzaldehyde 4‐cyanophenylhydrazone, C14H10BrN3, (IIb); (E)‐4‐iodobenzaldehyde 4‐cyanophenylhydrazone, C14H10IN3, (IIc); (E)‐4‐chlorobenzaldehyde 4‐nitrophenylhydrazone, C13H10ClN3O2, (III); and (E)‐4‐nitrobenzaldehyde 4‐chlorophenylhydrazone, C13H10ClN3O2, (IV). Both (Ia) and (Ib) are disordered (less than 7% of the molecules have the minor orientation in each structure). Pairs (Ia)/(Ib) and (IIb)/(IIc), related by a halogen exchange, are isomorphous, but none of the `bridge‐flipped' isomeric pairs, viz. (Ib)/(IIb), (Ic)/(IIc) or (III)/(IV), is isomorphous. In the nitrile–halogen structures (Ia)–(Ic) and (IIb)–(IIc), only the bridge N—H group and not the bridge C—H group acts as a hydrogen‐bond donor to the nitrile group, but in the nitro–halogen structures (III) (with Z′ = 2) and (IV), both the bridge N—H group and the bridge C—H group interact with the nitro group as hydrogen‐bond donors, albeit via different motifs. The occurrence here of the bridge C—H contact with a hydrogen‐bond acceptor suggests the possibility that other pairs of `bridge‐flipped' isomeric phenylhydrazones may prove to be isomorphous, regardless of the change from isomer to isomer in the position of the N—H group within the bridge.  相似文献   

7.
Organometallic linkage of heteroaromatic compounds provided a means of synthesizing complicated combinations of heteroaromatic compounds; not only nucleophilic aromatic substitution, but also “Ar? Cu/Ar? Hal linkage”, “organometallic oxidative linkage”, and “metal amide linkage” have been employed. The heterocyclopolyaromatic compounds are made of one, two, or three kinds of heteroaromatic species as ring members. These syntheses illustrate the construction of heterocycles from large, performed structural units. Competition experiments showed that the reactivity typical of the individual species is enhanced in open-chain combinations (ArNu)n and (ArE)n (ArNu, ArE: nucleo- and electrophilic heteroaromatic systems, respectively); the opposite situation is mostly encountered in the case of ArNu? AE combinations.—Cycloocta[1,2-b:4,3-b′5,6-b″:8,7-b?]tetrathiophene, the only heterocyclopolyaromatic system yet to have been studied in detail, proved surprisingly inclined to undergo monosubstitution reactions.  相似文献   

8.
詹传郎  王夺元 《中国化学》2000,18(3):418-424
We analyzed statistically the linear correlation of the solva-tochromic shifts of the stilbazolium-like dyes in the nonselected solvents with the reaction field function, L(εr) - bL( n2), and the solvent polarity parameter, ETN, respectively, and observed that there were not perfectly linearity relationships between them, so we introduced ETN into L(εr) - bL(n2) to form a new reaction field function, L(εr) - bL(n2) g ETN, called as the modified reaction field function, which can be perfectly linearly correlated with the solvatochromic shifts of the stilbazolium-like dyes in the nonselected solvents.  相似文献   

9.
The asymptotic behavior of a system's ground-state energy from the t expansion of Horn and Weinstein has been suggested to have the form E 1(t)=E 1+exp(–a n t+b n ). In the limit of very large t, this becomes E 1(t)=E 1+exp(–a 1 t+b 1). A simple analysis shows that the parameters are a 1=E 2E 1 and b 1=ln[(E 2E 1)|c 2|2/|c 1|2]. Functions are introduced which allow determination of a 1, b 1 and lower bounds to E 1.  相似文献   

10.
Rayleigh–Schrouml;dinger (RS ) perturbation expansions for the eigenvalues E(λ) of a hydrogen atom in the general polynomial perturbation V(r) = aλr + b>2r2, a, b > 0, are studied. When a2 = 2b, the ground state energy is exactly E(λ) = -(1/2) + (3/2)a>, i.e., the RS series is truncated. In the case a2 > 2b, the RS series is negative Stieltjes. In general, when λ < 0, a well of depth ω ≈ -a2/(4b2) (note the λ independence) is situated at rω = a/(2b|λ|). When a2 > 2b/N2, and interaction between this well and the hydrogenic state ψNLM(λ) is possible, thus creating a pair of asymptotically degenerate eigenstates separated by a “gap” δE(λ). The large order behavior of the RS coefficients E may be computed from the asymptotics of δE(λ), which is, in turn, related to the tunnelling integral. For excited states, stricter inequalities must be obeyed for Stieltjes behavior. The E(n)NLM may be calculated either numerically or in closed form via the “so(4, 2) Lie algebra technology” for such hydrogenic problems.  相似文献   

11.
Hooke numbers He ≡ σb/(E) are calculated from published ultimate tensile strengths σb, tensile moduli E, and ultimate elongations ?b. Data for common thermoplastics and natural fibers each follow a function He = [1 + (?b/?crit)ab]?1/b with a Hookean region I (He = 1) at ?b ? ?crit, a non-Hookean region III at ?b ? crit, and a transition region II for ?b ≈ ?crit. Only non-Hookean regions III were found for semisimultaneous interpenetrating networks from polyisobutylene-polymethyl methacrylate, thermoplastic elastomers from segmented polyamide-polyethers, molecular composites from poly(p-phenylene benzobisthiazole) and poly[2,5(6)-benzimidazole], and three groups of various synthetic fibers. The Hooke numbers of lyotropic and thermotropic liquid-crystalline polymers vary with the heat treatment and depend on orientation angles for orientation angles greater than ca. 10°. Hooke numbers much greater than 1 are observed for highly stressed polymers. ©1995 John Wiley & Sons, Inc.  相似文献   

12.
The elemental and ionic compositions of the surface of NbCl2(C n H n ) (n = 10–12), an active catalyst for acetylene cyclotrimerization into benzene, have been determined by X-ray photoelectron spectroscopy. Binding energy data for the sample sputtered with argon ions E b (Nb3d 5/2) = 203.8–204.2 eV) suggest that the oxidation state of niobium in the active catalyst is +2 or +3. The narrow C1s line indicates the equivalence of all carbon atoms, and the corresponding binding energy, E b(C1s) = 284.0 eV, is close to the BE value for cyclic unsaturated hydrocarbons with conjugate double bonds. Interacting with atmospheric oxygen and moisture during sample preparation, niobium ions on the catalyst surface oxidize to their highest oxidation state, +5, characterized by E b (Nb3d 5/2) = 207.3–207.7 eV. These data suggest that niobium oxychlorides or oxides form ion the sample surface. The catalyst is stable in a high vacuum and undergoes slight charging under the action of an X-ray beam, showing poor dielectric properties.  相似文献   

13.
The glow curve deconvolution (GCD) analysis of a composite thermoluminescence (TL) glow curve into its individual glow-peaks needs appropriate equations describing a single glow peak. In the present work, new single glow peak equations are presented, which are produced by transformation of the I(n 0,E,s,T) and I(n 0,E,s,b,T) single glow-peak equations into I(I m,T m,E,T) and I(I m,T m,E,b,T), respectively. Moreover, equations of the forms I(I m,T m,w,b,T) are also introduced. The proposed equations have two basic advantages: (1) they use parameters, which are directly obtained from the experimental glow peaks and (2) their accuracy is equal to that of the original thermoluminescence single glow-peak equations.  相似文献   

14.
Combined use of X-ray photoelectron spectroscopy (XPS) and in situ mass spectrometry made it possible to simultaneously obtain the O1s spectra and the mass spectrometric signal of formaldehyde (m/z = 30) in the course of heating (420–670 K) of polycrystalline foil in a flow of the reaction mixture of methanol and oxygen (with a total pressure of 0.1 mbar and a ratio of 3/1). It is shown that the O1s spectra contain two lines with E b = 530.1 and 531.2 eV, whose relative intensities depend on the sample temperature. At a low temperature (420 K) the line with a lower binding energy dominates, whereas sample heating leads to a drastic decrease in its intensity and its replacement by a line with a higher value of E b. A decrease in the intensity of the latter line occurs at T > 550 K, in the same temperature range as a drastic increase in the intensity of the formaldehyde signal. These lines were assigned on the basis of literature data and data obtained by the authors for the known forms of oxygen on copper and for the intermediate species of the reaction, such as methoxy and formate. The O1s line with E b = 530.1 eV was assigned to methoxy groups, and the line with E b = 531.2 eV was assigned to suboxide oxygen. The correlation of the intensity of the XPS signal of suboxide oxygen with the yield of formaldehyde was supported by stationary experiments using in situ XPS that prove its participation in the key step of the selective oxidation of methanol to formaldehyde.  相似文献   

15.
We performed a comprehensive study of the size‐, shape‐, and composition‐dependent polarizabilities of SimCn (m, n = 1–4) clusters on the basis of the density‐functional‐based coupled perturbed Hartree–Fock calculations. We found better correlations between the polarizabilities and both the binding energies (Eb) and change in charge distribution (Δq) than the energy gaps. The α values exhibit overall decreasing and increasing trends with increases in the Eb and Δq values, respectively. For isomers with the same Eb values and different polarizabilities, Δq can well explain the difference in polarizabilities. The π‐electron delocalization effect is the best factor for understanding the shape‐dependence. For a given m/n value, the linear clusters have an obviously larger polarizability than both the prolate and compact clusters, irrespective of the cluster size. We fit a quantitative expression [α = A ? (A ? B) × exp(?k(m/n))] to describe the composition‐dependent polarizabilities. © 2012 Wiley Periodicals, Inc.  相似文献   

16.
The large N expansion of the restricted Hartree–Fock (RHF) exchange energy per atom E(N) of the Pariser–Parr–Pople (PPP) model of cyclic polyenes (annulenes) CNHN is derived in detail. We explicitly derive the coefficients E0 and E1 of the asymptotic expansion: E(N)=E0+E1 ln N/N2+O(N−2), N→∞, in the very simple case of half-filling and no bond alternation. The exchange energy per atom in the infinite chain can be written as Eex=(2/π2)∑j=1{[γ(2j−1)]/[(2j−1)2]}, where γ is the two-electron repulsion integral in the infinite chain. On the other hand, the second coefficient E1 giving a finite-size correction is found to be 1/2b, where b is the bond length. This value of E1 differs slightly from that of a linear chain with periodic boundary conditions because the distance between sites depends upon the radius of the ring, i.e., upon N. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 66 : 397–407, 1998  相似文献   

17.
Abstract

The solid-state polymerization of diacetylenes (MDA-PBT-PDA) is studied with a concerted reaction model and the calculation method of EHMO-ASED and EHCO-ASED, where MDA = crystalline molecular diacetylenes, PBT = polybutatrienes, and PDA = polydiacetylenes. As the reaction goes on, the symmetry of frontier orbitals inverts at state PBT, HOCO from C 2-antisymmetry to C 2-symmetry and LUCO from C 2-symmetry to C 2-antisymmetry, which means completion of the 1,4-addition. Two necessary conditions must be satisfied for the reaction to take place: 1) the geometric parameters must undergo a series of concerted changes to make the conformation suitable for the intermolecular 1,4-addition, which should overcome an energy barrier Eb ; 2) the symmetry match between the frontier crystal orbitals of the reactant and the product must be satisfied-electrons of the reactant should be excited from HOCO (C 2-antisymmetry) into LUCO (C 2-symmetry), which faces an energy gap E g. At state MDA, there is E g(MDA) ≈ 5.6 eV. If MDA and PDA are analyzed according to Woodward-Hoffmann's rules, this reaction would be considered photochemically allowed but thermochemically forbidden. It has been shown that the E g gradually decreases along the reaction coordinate from state MDA to PBT. At state PBT there is E g(PBT) ≤ 0.1 eV, and the electrons of the reactant can be easily excited there. Since Eb ≤ 1.0 eV is not very large and Eg (PBT) ≤ 0.1 eV is very small, the two necessary conditions mentioned above can be satisfied thermally. Therefore, thermal polymerization can take place smoothly. By this pathway the apparent activation energy of the reaction will be Ea ≤ 1.0 eV, which is consistent with the activation energies of the polymerizations of diacetylenes in the literature.  相似文献   

18.
Photon transmission technique was used to monitor the evolution of transparency during film formation from poly(vinyl acetate) (PVAc) latex particles. The latex films were prepared below the glass transition temperature (T g) of PVAc. These films were annealed at elevated temperatures in various time intervals above the T g of PVAc. It is observed that transmitted photon intensity (I tr) from these films increased as the annealing temperature is increased. It is seen from I tr curves that there are two film formation stages. These successive stages are named void closure (viscous flow) and interdiffusion. The activation energies for viscous flow (ΔH) and backbone motion (ΔE b) were obtained by using well-defined models. The averaged values of the backbone (ΔE b) and the viscous flow activation energies (ΔH) were found to be 188.6 and 5.6 kcal/mol, respectively. The minimum film formation (τ M,T M) and healing points (τ H,T H) were determined. Minimum film formation (ΔE M) and healing activation energies (ΔE H) were measured using these time–temperature pairs. ΔE M and ΔE H were found to be 32.5 and 28.3 kcal/mol, respectively.  相似文献   

19.
A series of poly(cyclohexylethylene‐b‐ethylene‐co‐ethylethylene) (C‐E/EE) diblock copolymers containing approximately 50% by volume glassy C blocks and varying fraction (x) of EE repeat units, 0.07 ≤ x ≤ 0.90, was synthesized by anionic polymerization and catalytic hydrogenation. The effects of ethyl branch content on the melt state segment–segment (χ) interaction parameter and soft (E/EE) block crystallinity were studied. The percent crystallinity ranged from approximately 30% at x = 0.07 to 0% at about x ≥ 0.30, while the melting temperature changed from 101 °C at x = 0.07 to 44 °C at x = 0.28. Dynamic mechanical spectroscopy was employed to determine the order–disorder transition (ODT) temperatures, from which χ was calculated assuming the mean‐field prediction (χNn)ODT = 10.5. Previously published results for the temperature dependent binary interaction parameters for C‐E (x = 0.07), C‐EE (x = 0.90), and E‐EE (x = 0.07 and x = 0.90) fail to account for the quantitative x dependence of χ, based on a simple binary interaction model. © 2010 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 48: 566–574, 2010  相似文献   

20.
Arguably, the simplest way for the preparation of benzo[b]tellurophenes has been elaborated. Cyclization of o-haloarylalkynes in the presence of Te-NaOH-DMSO triad occurs by simple mixing of starting materials and heating overnight without the need for dry and inert atmosphere and provides the corresponding benzo[b]tellurophenes in up to 90 % GC-yield. Catalytic telluration of aryl iodide substrates extends the applicability of the developed cyclization cascade to substrates that are not suitable for the direct SNAr/5-endo-dig cyclization approach. Successful SEAr in the third position of benzo[b]tellurophene provides a synthetic intermediate for the preparation of Te analogues of selective estrogen receptor modulator (SERM) raloxifene.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号