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 312 毫秒
1.
Use was made of differential absorption in the near-infrared region to follow the rates of copolymerization of acrylonitrile (AN, M1) with ethylenesulfonic acid (ESA, M2) in aqueous zinc chloride solution. The concentrations of the monomers were followed separately and simultaneously. It was found experimentally that the ratios d log [M1]/dt and d log [M2]/dt were each constant. This was interpreted to mean that the product of the reactivity ratios of the two monomers (r1,r2) is unity and that the ratio of termination rate constants is equal to the propagation reactivity ratio. It was found that d log [M1]/d log [M2] = r1 = 4.52. This value is in fair agreement with polymer composition data obtained independently. In the Q—e system the equality r1r2 = 1 is equivalent to the monomers having equal e values. Thus, in the AN—ESA system, P1/P2 = k11/k21 = k12/k22 = k1T/k2T, where P1 is the resonance constant of polymer radicals ending in units of M1; and k11, k12, and k1T are the rate constants involving the reaction of this radical with M1, M2, and T (terminating agent), respectively. A gel effect was not observed even at M1 conversions as high as 88%.  相似文献   

2.
On the basis of the experimental data reported in literature, the contributions of cation mass (m) and molar volume (V) to lattice heat capacity (C) were analyzed. The volumetric-mass formula, Cx=(l —fC1+f·C2+Cm·(mxmx′), was presented for estimating the heat capacities of rare-earth compounds. In the formula C1 and C2 represent the lattice heat capacities of two reference substances respectively, f = VxV1/V2V1 and Cm represents the lattice heat capacity variation with the variation 1 g of cation mass. The equation relating the Cm with temperatures was derived as follows: Cm = 0.084 e ?0.0074T ?0.27 e ?0.045T, and mx and mx′ (= (1 - f) m1+f m2) represent the practical and “assumed” cation masses of the substance in question respectively.  相似文献   

3.
The glass transition temperature Tg of nylon 6 decreases monotonically toward a finite value Tgl upon increase of the moisture content. The mechanism of this decrease entails the reversible replacement of intercaternary hydrogen bonds in the accessible regions of the polyamide. The limiting glass transition temperature Tgl is approached when the moisture content approaches Wl, which corresponds to the amount of water required for complete interaction with all accessible amide groups. Denoting with Tg0 the glass transition temperature of the dry polymer, the effect of water on Tg is represented by the equation, Tg = (ΔTg)0 exp{?[ln(ΔTg)0]W/τWl} + Tgl, where (ΔTg)0 = Tg0 ?Tgl, and τ = W(Tgl+1)/Wl. This equation appears to be generally applicable to hydrophilic polymers, since correspondingly calculated data are also in very good agreement with experimental data for polymers such as nylon 66, poly(vinyl alcohol), and polyN-vinylpyrrolidone. The effect of water of Young's modulus E of nylon 6 is represented by an analogous relationship, and the quantity In[(E?El)/(Tg?Tgl)] is a linear function of the moisture content.  相似文献   

4.
A study was carried out on the rate constant ratio (k RH/k EtH) in reactions of alkanes C3H8, n-C4H10, n-C5H12, n-C6H14, i-C4H10, c-C5H10, and c-C6H12 with OH radicals in water at 5-55°C and the relative activation parameters A RH/A EtH and E EtHE RH. The values of E EtHE RH in water and the gas phase have opposite signs. The values of k RH/k EtH decrease with increasing temperature in the gas phase but increase in water. The behavior of these reactions in water may be attributed to a solvent cage effect.  相似文献   

5.
6.
 The optical absorption, photoluminescence, and photoconductivity spectra of some compounds of the formulas [R(CH2) n NH3] x M y X z , [R(CH2) n NH(CH3)2] x M y X z , [R(CH2) n S(CH3)2] x M y X z , [R(CH2) n SC(NH2)2] x M y X z , and [R(CH2) n SeC(NH2)2] x M y X z (R = organic residue; M = Bi(III), Pb(II), Sn(II), Cu(I), Ag(I) etc; X = I, Br, Cl; n, x, y, z = 0, 1, 2, 3, …) are briefly reviewed, and some new results are reported. The position, intensity, and shape of the excitonic bands depend on the dimensionality and size of the inorganic network as well as on the nature of the M, X, R, and onium moieties.  相似文献   

7.
A modification of a variation principle due to Delves, is derived which permits the direct calculation of energy differences between states of two different Hamiltonians: [Δ ??] = 〈X0| ??xWx|X1〉 – 〈Y0|??yWy|y1〉 + 〈X0| Δ ??|Y0〉 · 〈X0| Y0?1. Δ ?? = ??y – ??x, |X0〉 and |Y0〉 are the wave functions for the X and Y states and |X1〉 and |Y1〉 are functions defined in the text. The principle is applied to a few simple examples.  相似文献   

8.
Interaction of [Ru(-C 6 H 6 )Cl 2 ] 2 with indenyl- or fluorenyllithium in THF gives, together with cationic benzene complexes [Ru( 5 -C 9 H 7 )(-C 6 H 6 )]+ and [Ru( 5 -C 13 H 9 )(-C 6 H 6 )]+, the neutral cyclohexadienyl derivatives Ru( 5 -C 9 H 6 -C 9 H 7 ) and Ru( 5 -C 13 H 9 )( 5 -C 6 H 6 -C 13 H 9 ), respectively. Interaction of the cyclohexadienyl complexes with Al 2 O 3 , Ph 3 C+, and CF 3 CO 2 H has been studied. Reaction of Ru( 5 -C 13 H 9 )( 5 -C 6 H 7 ) with CF 3 CO 2 H in the presence of an arene yields cationic cyclohexadienylarene complexes: [Ru( 5 -C 13 H 9 )( 6 -arene)]+ (arene=C 6 H 6 or 1,3,5-Me 3 C 6 H 3 ).A. N. Nesmeyanov Institute of Organoelemental Compounds, Russian Academy of Sciences, 117813 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 3, pp. 699–706, March, 1992.  相似文献   

9.
Poly-S-vinyl-O-tert-butylthiocarbonate is an excellent precursor to poly(vinyl mercaptan) because the tert-butyloxycarbonyl blocking group can be removed by either acid hydrolysis or thermolysis under conditions which minimize the oxidation of the liberated mercaptan to disulfide. Dilatometric studies of the homopolymerization of S-vinyl-O-tert-butylthiocarbonate demonstrated that the polymerization rate was directly proportional to the concentration of free-radical initiator; no thermal initiation was observed. The molecular weight of the homopolymers and copolymers ranged from 30,000 to 50,000 (GPC). Copolymerization of S-vinyl-O-tert-butylthiocarbonate (M2) with styrene, (r1 = 3.0, r2 = 0.2), methyl methacrylate (r1 = 1.40, r2 = 0.17) and vinyl acetate (r1 = 0.04, r2 = 11.0) indicated that a sulfur atom adjacent to the vinyl group increases the resonance stability (Q2 = 0.5) and the electron density (e2 = ?1.4) of the double bond and the corresponding radical. Water-soluble copolymers could be prépared by incorporating either N-vinylpyrrolidone (r1 = 0.12, r2 = 3.94) or N-isopropylacrylamide (r1 = 1.17, r2 = 0.3) with M2. The water solubility of the copolymers decreased markedly when the tert-butyloxycarbonyl group was removed. Copolymers of M2 with N-vinyl-O-tert-butylcarbamate (r1 = 0.13, r2 = 5.10) were utilized to prepare crosslinked poly(vinyl amine–vinyl mercaptan); the crosslinking resulted from urea linkages formed during thermolysis of the copolymer.  相似文献   

10.
The AM1 calculation was done for ortho-substituted toluenes (o-X-C6H4-CH3) and ortho-substituted tert-butylbenzenes (o-X-C6H4-t-Bu). The difference in the calculated heat of formation between o-X-C6H4-CH3 and o-X-C6H4-t-Bu was used as a theoretical steric index for ortho-X. The correlation of this theoretical steric index with the empirical steric parameter sets such as our recently defined Es(AMD) and the Taft–Kutter–Hansch (TKH) Es was examined. In spite of the simplicity of the model system, the theoretical index was linear with the Es(AMD) constant with a correlation coefficient of r = 0.972 for 17 substituents of various structures. Including the phenyl group, the correlation with the TKH Es constant was r = 0.948. The theoretically calculated index was shown to serve as a measure of the ortho steric effect.  相似文献   

11.
On the basis of an isoviscosity criterion for the glass transition (ηg ? 1013 poise) in liquids of low molecular weight, theoretical Tg values were calculated for the n-alkane series by the equation log η = log A + B/(T ? T0), with the use of values reported by Lewis for the parameters. The Tg/T0 ratio reaches a limiting value of 1.25 and ?g = (Tg ? T0)/2.3B = 0.027, a constant. Extrapolation to (CH2) gives Tg = 200°K., T0 = 160°K., and B = 640°K. This Tg is consistent with other estimates for poly-ethylene, and T0 coincides with the temperature at which the “excess” liquid entropy for (CH2) becomes zero from thermodynamic data. For polymer liquids it is proposed that E0 = 2.3RB is determined by the internal barriers to rotation for the “isolated” polymer chains. Thus, E0 = 2.9 kcal./mole for polyethylene, 3.0 kcal./mole for polystyrene, 5.7 kcal./mole for polyisobutylene, and 1.9 kcal./mole for polydimethylsiloxane.  相似文献   

12.
Summary It was proved that the partial adsorbed quantities can be calculated from the individual adsorption isotherms. This calculation can be carried out by equations which take into' account the energetical heterogeneity of the surface of adsorbent as well. Thus, the partial isotherms equations which can be applied are as followsV A (p A ,p B ) =p A /[b a + (p A +b AB p B ) m ] 1/m andV B (p A ,p B ) =p B /[b B + (p B +b BA p A ) m ] 1/m wherep A andp B are the partial equilibrium pressures, whileb A ,b B andm are constants which can be determined on the basis of the individual isotherms.
Zusammenfassung Es wurde nachgewiesen, daß die partialen Adsorbatmengen aus den individuellen Isothermen berechnet werden können. Die Berechnung kann mit Hilfe von Gleichungen durchgeführt werden, die auch die energetische Inhomogenität der adsorbierenden Oberfläche berücksichtigen. Benutzt werden die Gleichungen: VA(pA,pB) = pA/ [bA + (PA + bABpB)m]1/m und VB(pA,pB) pB/[bB + (pB + bBApA)m]1/m , wop A undp B die Gleichgewichtspartialdrucke undb A ,b B undm die aus den individuellen Isothermen zu bestimmenden Konstanten darstellen.
  相似文献   

13.
Crosslinks are introduced by γ irradiation into 1,2-polybutadiene while strained in uniaxial extension near Tg with stretch ratio λ0, thereby trapping a proportion of the entanglements originally present. The stress at any subsequent strain λ is accurately given by the sum σN + σx, where σN is the stress contributed by a trapped entanglement network with λ = 1 as reference and a Mooney–Rivlin stress-strain relation, and σx is that contributed by a crosslink network with λ = λ0 as reference and neo-Hookean stress-strain relation. The birefringence is accurately given as δn = ?NσN + ?xσx, where the ?'s are the respective stress-optical coefficients. From measurements at λ = λ0 where σx = 0, ?N can be determined separately. For polymer with 88% 1,2 microstructure, ?N and ?x are nearly equal and independent of irradiation dose, though strongly dependent on temperature. For polymer with (95–96)% 1,2, ?N and ?x are different (even opposite in sign) and dependent on dose. This behavior is associated with a side reaction of cyclization by the γ irradiation, which is inhibited by the 1,4 moiety in the polymer with lesser 1,2 content. It is responsible for residual birefringence in the state of ease (λ = λs) where σN = –σx and the stress is zero.  相似文献   

14.
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   

15.
The reactivity of κ2-N,S-chelated ruthenium borate complexes, [Ru(PPh3)(κ2-N,S-(NC7H4S2)Ru{κ3-H,S,S′-H2B (NC7H4S2)2}] ( 1 a ) and [Ru(PPh3)(κ2-N,S-(NC5H4S){κ3-H,S,S′-H2B(NC5H4S)2}] ( 1 b ) with monoboranes have been explored. The prolonged room temperature reaction of [Ru(PPh3){κ2-N,S-(NC7H4S2)}{κ3-H,S,S′-H2B(NC7H4S2)2}], 1 a with an excess of [BH3 ⋅ THF] led to the formation of hydrogen-rich complex, arachno-[PPh3(κ2-B3H8)Ru{κ3-H,S,S′-H2B(NC7H4S2)2] ( 2 ). In a similar fashion, the isomeric ruthenatetraborane complexes of [Ru(PPh3)(κ2-N,S-(B3H8){κ3-H,S,S′-H2B(NC5H4S)2}], 4 and 5 were isolated from the room temperature reaction of 1 b with [BH3 ⋅ THF]. In complex 2 , the borate ligand, [H2B(NC5H4S)2] and the PPh3 occupy the endo and exo sites of the butterfly-shaped RuB3 core, respectively. In contrast, the borate unit [H2B(NC5H4S)2] in 4 sits on the exo position of the butterfly core, while the phosphine ligand occupies the endo site. Multinuclear spectroscopic analyses were done to characterize all new complexes and the structures were further confirmed by single-crystal X-ray diffraction analysis. Density functional theory (DFT) calculations were performed to probe into the bonding modes of these complexes.  相似文献   

16.
The reaction of bis(N-salicylidene)dicarboxylic acid dihydrazides (H4 Lig) with cobalt(II) salts was investigated. Chelates of the types Co(H3 Lig)X ·nH2O, Co2(H2 Lig)X 2, Co3(H3 Lig)2 X 4, Co(H2 Lig) ·nH2O and Co(H3 Lig)2 ·nH2O (X=Cl, Br, I or SCN) were isolated and characterized by their infrared and electronic spectra as well as their magnetic properties.
Koordinationsverbindungen von Hydrazinderivaten mit Übergangsmetallen, 23. Mitt.: Kobalt(II)-Chelate von Bis(N-salicyliden)dicarbonsäuredihydraziden
Zusammenfassung Die Reaktion von Bis(N-salicyliden)dicarbonsäuredihydraziden (H4 Lig) mit Kobalt(II)-Salzen ergab Chelate vom Typ Co(H3 Lig)X ·nH2O, Co2(H2 Lig)X 2, Co3(H3 Lig)2 X 4, Co(H2 Lig) ·nH2O and Co(H3 Lig)2 ·nH2O (X=Cl, Br, I oder SCN). Die Charakterisierung erfolgte mittels IR, Elektronenspektren und magnetischer Eigenschaften.
  相似文献   

17.
Single crystals of Rh(Si2O)(PO4)3 and In4(Si2O) · (PO4)6 were prepared by chemical transport reactions in silica tubes and their structures were determined. Crystal data of Rh(Si2O)(PO4)3: trigonal, space group P 3 c1, a = 8.088(3) Å, c = 8.740(2) Å, Z = 2, R(F2) = 0.0379, Rw(F2) = 0.0518 for 601 unique reflections. In4(Si2O)(PO4)6: hexagonal, space group P63/m, a = 8.5149(10) Å, c = 7.7481(12) Å, Z = 1, R(F2) = 0.0436, Rw(F2) = 0.0522 for 509 unique reflections. Both of the compounds have hexagonal close packed array of phosphate groups with metal atoms and SiOSi units in the octahedral interstices, where the SiOSi units show occupational disorder. The structure of the indium compound is considered to be a disordered structure of the reported Mo4Si2P6O13 structure, and contains confacial bioctahedral units.  相似文献   

18.
t-Butyl derivatives play a significant role in the organometallic chemistry of group 13 metals. It was shown on the basis of reactions of t-Bu3Al·OEt2 with [p-RC6H4S(O)2C(H)2C(Ph) = O] (where R = CH3, Cl) β-keto sulfones that the structure of the reaction products depends on the purity of the aluminum compound used. In the reactions, in addition to the expected complexes [p-RC6H4S(O2)C(H) = C (Ph)-OAl(t-Bu)2] [where R = CH3 ( 2 ); R = Cl ( 4 )] possessing β-keto sulfone ligands, complexes with β-hydroxy sulfone ligands [p-RC6H4S(O2)C(H)2-C (Ph)-OAl(t-Bu)2] [where R = CH3 ( 1 ); R = Cl ( 3 )] were formed. Compounds 1 and 3 were the result of the hydroalumination reaction of the β-keto sulfone ligands with t-Bu2AlH, which is an impurity of t-Bu3Al. These compounds are obtained, for the first time, as intermediate products in the hydrogenation reaction of β-keto sulfones. In this work, during t-Bu3Al·OEt2 production t-Bu2AlH·OEt2 formed as a by-product. Re-examination of reaction conditions of AlCl3 with t-BuMgCl resulted in a control of the t-Bu2AlH·OEt2 by-product content in t-Bu3Al·OEt2.  相似文献   

19.
The problem of pure-state N-representability of the two-particle spin-dependent density function ρ(x1, x2) is considered for an N-electron system, and a procedure for finding an N-representable ρ(x1, x2) is advanced. The problem is formulated in the framework of a family of N × N matrices formed from integrals of auxiliary two-particle functions θn(x1, x2) converging at n → ∞ to ρ(x1, x2)/[N(N−1)]. The simple requirement of positive definiteness of these matrices is shown to play a decisive role in finding an N-representable ρ(x1, x2). The results obtained may open new possibilities for using ρ(x1, x2) in the density-functional theory. © 1997 John Wiley & Sons, Inc. Int J Quant Chem 65 : 127–142, 1997  相似文献   

20.
A series of potentially useful lithium amidinates and guanidinates were prepared and fully characterized. Treatment of N,N′‐diisopropylcarbodiimide with phenyllithium in diethyl ether afforded the lithium amidinate [PhC(NiPr)2Li(OEt2)]2 ( 1 ). Similar treatment of N,N′‐diorganocarbodiimides R′–N=C=N–R′ [R′ = iPr, cyclohexyl (Cy)] with secondary lithium amides LiNR2 [R2 = Et2, iPr2, (CH2)4] followed by crystallization from THF or 1,4‐dioxane gave the lithium guanidinates [R2NC(NR′)2Li(S)]2 [ 2 : R = Et, R′ = iPr, S = THF; 3 : R2 = (CH2)4, R′ = iPr, S = THF; 4 : R = R′ = iPr, S = ½ 1,4‐dioxane; 5 : R2 = (CH2)4, R′ = Cy, S = 1,4‐dioxane] as crystalline solids. Reaction of N‐lithioaziridine with the corresponding carbodiimides afforded solvent‐deficient [{C2H4NC(NiPr2)2}2Li2(THF)]2 ( 6 ), and [C2H4NC(NEt)(NtBu)Li(THF)]2 ( 7 ). Crystal structure determination revealed the presence of common ladder‐type dimeric structures for 1 – 5 . Compound 6 exists as a dimer of two ladder‐type dimers in the crystal, and 7 exhibits an unusual dimeric structure comprising an eight‐membered C2N4Li2 ring.  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