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 500 毫秒
1.
We have used nuclear reaction analysis to measure diffusion coefficients D in couples consisting of hydrogenated polybutadienes of structure (C2H3(C2H5))x(C4H8)1?x and their partly deuterated counterparts. The 1,2- and 1,4-olefinic isomers are randomly distributed along the chains and the mean vinyl fraction x varies between 0.38 and 0.94. We find that the effective monomeric mobility D0 [defined by D = D0(Ne/N2) for each copolymer, where N is the backbone length and Ne the entanglement spacing] decreases monotonically with increasing vinyl content x. Over the range of microstructures and temperatures T (?14?40°C) investigated we find log(D0/T) varies smoothly with (T ? Tg), where Tg is the glass transition temperature of the respective melts. An analysis of our data in terms of a simple activated rate process model suggests that D0 is controlled by thermally activated hopping of segments whose effective volume is close to that of the respective statistical segment lengths of the copolymeric chains. ©1995 John Wiley & Sons, Inc.  相似文献   

2.
Considering one long chain (N monomeric units) in a homodisperse melt of chemically identical, but shorter, “solvent” chains (P monomers per chain), we propose some tentative scaling laws for the self-diffusion constant D(N) and the relaxation time T(N) of the solute chain. We also discuss the viscosity increment δη due to a small volume fraction Φ of the long chains. We find three regimes of behavior, depending on N and P, and on the distance between entanglement points (assumed smaller than N and P): (A) reptation of the N chain; (B) Stokes–Einstein regime; the solute moves like a usual polymer coil in a viscous fluid of P chains; (C) mixed regime, where D(N) is controlled by reptation, while δη is of type B. Contrary to our earlier belief, we find no significant regime where the process of “tube renewal” could be dominant.  相似文献   

3.
Using crossed beams of ground state alkali atoms A (A = Li, Na, K, Rb, Cs) and metastable He(23 S), He(21 S) atoms, we have measured the energy spectra of electrons resulting in the respective Penning ionization processes at: thermal collision energies. The data are interpreted to yield the well depthD e * of the2Σ interaction potentials as follows: He(23 S)+A:D e * (A=Li)=868(20) meV;D e * (Na)=740(25) meV;D e * (K)=591(24) meV;D e * (Rb)=546(18) meV;D e * (Cs)=533(18) meV. He(21 S)+A:D e * (Li)=330(17) meV;D e * (Na)=277(24) meV;D e * (K)=202(23) meV;D e * (Rb)=219(18) meV;D e * (Cs)=277(18) meV. The well depth for He(23 S)+A(2Σ) is always close to 80% of the well depth for Li(2s)+A(X 1Σ). The ionization cross sections for He(21 S)+A are about 3 to 4 times larger than those for He(23 S)+A.  相似文献   

4.
Gas‐phase FTIR spectra of the ν6 (B‐type) and the ν4 (C‐type) fundamental bands of S2N2 (D2h) were recorded with a resolution of ≤0.004 cm?1 and the vibrational spectrum of S2N2 (D2h) in solid Ar has been revisited. All IR‐active fundamentals and four combination bands were assigned in excellent agreement with calculated values from anharmonic VPT2 and VCI theory based on (explicitly correlated) coupled‐cluster surfaces. Accurate experimental vibrational ground‐ and excited‐state rotational constants of 32S214N2 are obtained from a rovibrational analysis of the ν6 and ν4 fundamental bands, and precise zero‐point average rz (Rz(SN)=1.647694(95) Å, αz(NSN)=91.1125(33)°) and semi‐experimental equilibrium structures (Re(SN)=1.64182(33) Å, αe(NSN)=91.0716(93)°) of S2N2 have been established. These are compared to the solid‐state structure of S2N2 and structural properties of related sulfur nitrogen compounds and to results of ab initio structure calculations.  相似文献   

5.
Several predictions for a recently proposed mesoscopic model for polymer melts and concentrated solutions is presented. It is a single Kramers chain model in which elementary motions of the Orwoll-Stockmayer type are allowed. However, for this model, the bead jumps are no longer given by a Markovian probability, but rather are described by a fractal “waiting-time” distribution function, with a single adjustable parameter β, which describes the long-time behavior of the distribution: ∼ 1/t1+β. We find that the model predicts D ∼ 1/N2 and η0N3.4 for β ≈︁ 1.4, where N is the degree of polymerization. The generalized model predicts that the relaxation spectrum has a plateau regime whose height is independent of N, but whose width is strongly N dependent, in agreement with experiment. The model also predicts that rings will diffuse somewhat more slowly than linear chains of the same molecular weight (about 80% as fast), with the same scaling dependence on N as linear chains, also in agreement with preliminary data.  相似文献   

6.
K2Fe[P2S6] was synthesized from the elements at 1173 K in sealed quartz tubes. The compound forms transparent orange crystals, stable against air and moisture. K2Fe[P2S6] crystallizes in the monoclinic system, space group P21/n (No. 14), with cell dimensions (T = 298.5 K) a = 6.0622(4), b = 12.172(1) and c = 7.3787(8) Å, β = 101.113(7)°, Z = 2. The novel structure type (mP22) is characterized by columns of alternating face-sharing S6 octahedra and trigonal antiprisms (both distorted) parallel to the a axis, which are interconnected by inserted K+ (CN 10; {2,6,2}-polyhedra; d(K? S) = 3.231 ? 3.845 Å). The S6 polyhedra of the columns are centered alternately by Fe (d?(Fe? S) = 2.577 Å) and P2 pairs which are inclined to the a axis by 73.4°. The bond lengths in the hexathiodiphosphate(IV) anions, [P2S6]4?, with approximate 3 2/m – D3d symmetry, are d?(P? P) = 2.20 and d?(P? S) = 2.02 Å. The compound is paramagnetic above TN = 28 K with μ = 4.69 B.M. and orders antiferromagnetically below TN. The internal modes of the observed Raman and FIR spectra of K2Fe[P2S6] are in accord with the factor group analysis, and the spectra are assigned on the basis of [P2S6]4? units, taking into account the deviation from D3d symmetry.  相似文献   

7.
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   

8.
The highly stable nitrosyl iron(II) mononuclear complex [Fe(bztpen)(NO)](PF6)2 (bztpen=N‐benzyl‐N,N′,N′‐tris(2‐pyridylmethyl)ethylenediamine) displays an S=1/2?S=3/2 spin crossover (SCO) behavior (T1/2=370 K, ΔH=12.48 kJ mol?1, ΔS=33 J K?1 mol?1) stemming from strong magnetic coupling between the NO radical (S=1/2) and thermally interconverted (S=0?S=2) ferrous spin states. The crystal structure of this robust complex has been investigated in the temperature range 120–420 K affording a detailed picture of how the electronic distribution of the t2g–eg orbitals modulates the structure of the {FeNO}7 bond, providing valuable magneto–structural and spectroscopic correlations and DFT analysis.  相似文献   

9.
The spatial distributions of electron temperature and density in a dc glow discharge that is created by a pair of planar electrodes were obtained by using double Langmuir probes. The contribution of double Langmuir probes measurement is to provide a relatively quantitative tool to identify the electron distribution behavior. Electrons gain energy from the imposed electric field, and electron temperature (Te) rises very sharply from the cathode to the leading edge of the negative glow where Te reaches the maximum. In this region, the number of electrons (Ne) is relatively small and does not increase much. The accelerated electrons lose energy by ionizing gas atoms, and Te decreases rapidly from the trailing edge of the negative glow and extends to the anode. Ne was observed to increase from the cathode to the anode, which is due to the electron impact ionization and electron movement. The electron density was observed to increase with increasing discharge voltage while the electron temperature remained approximately. At 800 V and 50 mTorr argon glow discharge, Te ranged from 15 to 52 eV and Ne ranged from 6.3×106/cm3 to 3.1×108/cm3 in the DC glow discharge, and Te and Ne were dependent on the axial position.  相似文献   

10.
《Chemical physics》1987,115(1):23-32
Using double-zeta plus polarization (DZP) basis sets systematically augmented with a variety of bond functions, the term dissociation energies are calculated for the A3Σ+u, B3Πg and W3Δu states of N2. It is found that the best agreement with literature values is generally with a basis set composition of DZP augmented by a set of s, p and d orbitals at the bond midpoint. The excited state potential energy curves and spectroscopic constants for the B3Πg state are calculated from this basis and compared with experimental values. Good agreement was obtained, considering the small basis set size, with the spectroscopic constants ωe, ωeχe, ωeye, Be and αe and the dissociation energy De (e.g., De = 3.38 (3.681, exp.), 4.75 (4.897) and 4.77(4.873) eV for the A, B and W stages, respectively). Poorer agreement was obtained for the term energy T0 (7.92 versus 7.35 eV, exp., for the B state). The error in term energy arises largely from an error in the calculated 4S → 2D splitting (2.705 versus 2.383 eV, exp.), and shifting the potential curve for the B state by a constant amount leads to much improved agreement relative to the ground state. The counterpoise correction applied to the potential curve of the B state causes a drastic deterioration of the results and shows qualitatively incorrect behaviour, and is therefore not recommended for calculations of this type.  相似文献   

11.
Two inclusion compounds of dithiobiurea and tetrapropylammonium and tetrabutylammonium are characterized and reported, namely tetrapropylammonium carbamothioyl(carbamothioylamino)azanide, C12H28N+·C2H5N4S2, (1), and tetrabutylammonium carbamothioyl(carbamothioylamino)azanide, C16H36N+·C2H5N4S2, (2). The results show that in (1), the dithiobiurea anion forms a dimer via N—H...N hydrogen bonds and the dimers are connected into wide hydrogen‐bonded ribbons. The guest tetrapropylammonium cation changes its character to become the host molecule, generating pseudo‐channels containing the aforementioned ribbons by C—H...S contacts, yielding the three‐dimensional network structure. In comparison, in (2), the dithiobiurea anions are linked via N—H...S interactions, producing one‐dimensional chains which pack to generate two‐dimensional hydrogen‐bonded layers. These layers accommodate the guest tetrabutylammonium cations, resulting in a sandwich‐like layer structure with host–guest C—H...S contacts.  相似文献   

12.
Introduction So far, considerable attention has been paid to mag-netic interaction between two different metal ions.1-3 As a potential bridging ligand, thiocyanate can coordinate to a harder metal center with N atom and softer ones with S atom at the same time, resulting in the formation of small ferromagnetic coupling.2 On the other hand, the Fe(III) atom is a good candidate as a hard acid and Ag(I) is a good candidate as a soft acid, so that the Fe(III) centers could be expected to conn…  相似文献   

13.
We have performed ab initio fourth-order Møller–Plesset perturbation theory calculations in the framework of the supermolecule approach on the vertical excitation spectra of the weakly bound van der Waals N2–He dimer. They indicate a ``T-shaped' stablest ground N2(X1g+)–He(1S) electronic state with a well depth, De, of 21.63 cm–1 at a minimum distance, Re, of 3.44 Å and zero-point vibration correction, Do, of 7.07 cm–1. They also indicate a ``T-shaped' stablest excited conformer with Re=3.25 Å, De=36.85 cm–1 and Do=17.06 cm–1 for the N2(B3g)–He(1S) triplet electronic level. In order to investigate the use of less-demanding correlation methods, test density functional theory calculations using the mPW1PW exchange–correlation functional are also presented for comparison.  相似文献   

14.
The total energies of ZnO(1Σ), ZnO(3Π), ZnO?(2Σ), ZnO+(2Σ), ZnS(1Σ), ZnS(3Π), ZnS?(2Σ), and ZnS+(2Σ) were calculated ab initio by the CCSD(T) method with the use of atomic basis sets including 80, 84, and 93 functions for O, S, and Zn, respectively. Similar calculations were performed for the Zn atom [Zn(1 S), Zn(3 P), Zn+(2 S), Zn2+(1 S)] and several oxygen and sulfur states [O(3 P), O?(2 S), O(1 D), O2(3Σ), O 2 ? (2Π), O2(1Δ), S(3 P), S?(2 S), S(1 D), S2(3Σ), S 2 ? (2Π), and S2(1Δ)]. The ideology of engagement groups suggested by us is considered. According to this approach, data treatment can be performed on the assumption that the errors in all the 24 results obey the normal distribution law. As a result, we obtained D e(ZnO) = 1.70 ± 0.21 and D e(ZnS) = 1.57 ± 0.25 eV (at a 95% confidence level).  相似文献   

15.
The reaction of 1,1,2,3,3,3-hexafluoropropyldiethylamine (PPDA) with hydroxy esters having PhC(H)(OH)-groups gave their corresponding secondary fluorides. Fluorination of (R)-(-)-mandelic acid ethylester ([α]29D −106.46) gave ethyl (S)-(+)-2-fluoro-2-phenyl acetate ([α]29D +6.57). The value of the enantiomeric excess (% ee) was 74.1. The reaction mechanism of the fluorination by PPDA is suggested to be the SN2 type.  相似文献   

16.
Structure and dynamics of a free aquaporin (AQP1) are studied by a coarse-grained Monte Carlo simulation as a function of temperature using a phenomenological potential with the input of a knowledge-based residue–residue interaction. Response of the radius of gyration (R g) of the protein to the temperature (T) is found to be nonlinear: Decay of R g at T ≤ T c is followed by a continuous increase at T ≥ T c before reaching its saturation. In thermo-responsive regime, the protein exhibits segmental globularization with the persistence of three regions along its sequence involving residues 1M–25V and 250V–269K toward the beginning and end segments with a narrow intermediate region around 155A–163D. A detail analysis of the structure factor S(q) shows a global random coil conformation at high temperatures with an effective dimension D e ~ 1.74 and a globular structure (D e ~ 3) at low temperatures. In thermo-responsive regime, the variation of S(q) with the wave vector q reveals a systematic redistribution of self-organizing residues (in globular and fibrous sections) that depends on the length scale and the temperature.  相似文献   

17.
The potential energy surfaces of N8 clusters were investigated by density functional theory (DFT) and a possible synthesis reaction pathway for N8 (CS) was suggested. The species involved were fully optimized up to the B3LYP/6‐311+G* level of theory. Relative energies were further calculated at the QCISD/6‐311+G*//B3LYP/6‐311+G* level. The reaction rate constants of these steps from the 1 (N5+?N3?, complex, CS) to 2 (N8, CS), 2 (N8, CS) to 3 (N8, CS), 3 (N8, CS) to 4 (N8, D2d), and 4 (N8, D2d) to 5 (N8, CS) reactions were predicted by the VTST theory. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 1334–1339, 2001  相似文献   

18.
A quantitative survey on the performance of multireference (MR), configuration interaction with all singles and doubles (CISD), MRCISD with the Davidson correction and MR-average quadratic coupled cluster (AQCC) methods for a wide range of excited states of the diatomic molecules B2, C2, N2 and O2 is presented. The spectroscopic constants r e, ωe, T e and D e for a total of 60 states have been evaluated and critically compared with available experimental data. Basis set extrapolations and size-extensivity corrections are essential for highly accurate results: MR-AQCC mean-errors of 0.001 ?, 10 cm−1, 300 cm−1 and 300 cm−1 have been obtained for r e, ωe, T e and D e, respectively. Owing to the very systematic behavior of the results depending on the basis set and the choice of method, shortcomings of the calculations, such as Rydberg state coupling or insufficient configuration spaces, can be identified independently of experimental data. On the other hand, significant discrepancies with experiment for states which indicate no shortcomings whatsoever in the theoretical treatment suggest the re-evaluation of experimental results. The broad variety of states included in our survey and the uniform quality of the results indicate that the observed systematics is a general feature of the methods and, hence, is molecule-independent. Received: 12 June 2000 / Accepted: 1 September 2000 / Published online: 21 December 2000  相似文献   

19.
Abstract

Radical homopolymerization of N-[4-N′-(α-methylbenzyl)-aminocarbonylphenyl]maleimide ((S)-MBCP) was carried out at 50 and 70°C for 24 h to give optically active polymers ([α]25 D = 159.8 to 163.4°). Radical copolymerizations of (S)-MBCP (M1) were performed with styrene (ST, M2, methyl methacrylate (MMA, M2) in THF at 50°C. The monomer reactivity ratios (r 1, r 2) and the Alfrey-Price Q, e values were determined as follows: r 1 = 0.32, r 2= 0.14, Q 1 = 1.74, e 1 = 0.96 in the (S)-MBCP-ST system; r 1 = 0.54, r 2 = 0.93, Q 1 = 1.11, e 1 = 1.23 in the (S)-MBCP-MMA system. Chiroptical properties of the polymers and the copolymers were also investigated, and asymmetric induction into the copolymer main chain is discussed.  相似文献   

20.
A novel perfluorinated liquid crystal 4′-(2,2,3,3,4,4,5,5,6,6,7,7,8,8,8-pentadecafluorooctanoyloxy)biphenyl-4-yl undec-10-enoate (PFOBU) was synthesized, which exhibited smectic C phase. Several liquid crystalline polymers (PI–PVI) were synthesized by use of poly(methylhydrogeno)siloxane, PFOBU, and cholesteryl 3-(4-allyloxy-phenyl)-acryloate. The chemical structures and liquid crystalline (LC) properties of the monomers and polymers, and some ferroelectric properties of the chiral smectic C (SC*) phase were characterized by use of various experimental techniques. The effect of perfluorocarbon chains on phase behaviors of the fluorinated LC polysiloxanes was studied as well. PI and PII showed single chiral nematic (N*) mesophase when they were heated and cooled, but PIII, PIV, PV, and PVI containing more perfluorocarbon chain units exhibited SC* phase besides N* mesophase. Introduction of perfluorocarbon chain containing mesogens to the chiral cholesteryl-containing polymer systems resulted in a SC* mesophases, indicating that the fluorophobic effect could lead to microphase segregation and modifications of smectic mesophases from the chiral nematic phase.  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