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 15 毫秒
1.
 本文用激光光散射方法研究了具有特殊相行为[(低临界溶解温度(LCST),高临界溶解温度(UCST)共存]共混体系羧化聚苯醚和聚苯乙烯在UCST域内的不稳相分离初期分子量对动力学参数的影响。结果表明:相分离初期动力学过程与Cahn理论吻合;随着分子量增加,表现扩散系数Dapp明显减小;该体系的表现扩散系数为10-14 cm2s-1数量级。deGennes管子模型可很好地描述不稳相分离初期大分子扩散行为。  相似文献   

2.
The effect of initial concentration of products in a general reaction of reversible process with complex stoichiometry expressed as mO + ne? ai qR was evaluated using numerical simulations with various complex stoichiometric coefficients m and q of oxidized (O) and reduced (R) species, respectively, and with various initial concentrations. In the simulations, the transform of voltammograms according to changes in the initial concentrations were characterized by the parameter ??? (= c 0*(D O)1/2/c R* (D R)1/2) for any stoichiometric system. Based on the simulation results, empirical relations were obtained for (1) peak potential involving parameters for stoichiometry (m and q) and for the initial conditions (???), and (2) peak current.  相似文献   

3.
The pressure dependences of the self-diffusion coefficients of deuterium oxide in 4.5m solutions of LiCl–D2O and CsCl–D2O (also 7m) and 3.06m CaCl2–D2O have been measured by the NMR spin-echo method at 30°C, 60°C, and 90°C. Shear viscosities and densities of these solutions have also been determined over the same range of experimental conditions. The experimental data show that the diffusion constantD decreases with the increasing structure-making ability of the electrolyte cation Ca+2>Li+. In contrast, the diffusion coefficient for D2O in the 4.5 and 7m CsCl solutions is equal to that for pure D2O at 30°C but lower at 60°C and 90°C. It has been found that the Stokes-Einstein equation relates well the diffusion coefficients to shear viscosity in these concentrated electrolyte solutions.  相似文献   

4.
Abstract

Optically active (R,R)-(-)-trans-1,2-dichlorocyclohexane (DCC) was isolated as an inclusion crystal with the optically active host, (R,R)-(-)-trans-2,3-(hydroxydiphenylmethyl)-1,4-dioxaspiro[4.4]-nonane, and the structure of the 2:1 inclusion crystal has been determined by X-ray analysis. Crystal data: C72H74O8Cl2, orthorhombic, P21212 (No. 18), a = 17.465(6) Å, b = 20.095(6) Å, c = 8.664(5) Å, V = 3040(2) Å3, Z = 2, Dc = 1.24g cm?3, Dm = 1.23g cm?3, T = 293 K and final R 1 = 0.050 for 2766 observed data (I > 2σ(I)). The conformation of DCC in the inclusion crystal has been found to be equatorial and the absolute configuration was definitely determined to be (R,R) on the basis of the known configuration of the host.  相似文献   

5.
The quotients for the ionization of D2O and the neutralization of D2PO 4 have been determined potentiometrically in 0.2m KCl from 50 to 300°C at the saturation pressure. By combination with the other data, analytical expressions for the dependence on temperature and ionic strength have been derived. Rounded values for the thermodynamic quantities for the ionization of D2O and the neutralization of D2PO 4 are given along with standard errors. The magnitude of the isotope effects is discussed in terms of the zero-point-energy approximation and the acid strength in light water.ORAU Summer Trainee Program 1973.  相似文献   

6.
Length scale hierarchy in gelatin sol, gel, and coacervate (induced by ethanol) phases, having same concentration of gelatin in aqueous medium (13% w/v), has been investigated through small angle neutron scattering and rheology measurements. The static structure factor profile, I(q) versus wave vector q, was found to be remarkably similar for all these samples. This data could be split into three distinct q‐regimes: the low‐q regime, Iex(q) = Iex(0)/(1+q2ζ2)2 valid for q < 3Rg?1; the intermediate q‐regime, I(q) = I(0)/(1+q2ξ2) for 3Rg?1 < q < ξ?1; and the asymptotic regime, I(q) = (c/q) exp(?Rc2q2/2) for q > ξ?1. Consequently, three distinct length scales could be deduced from structure factor data: (a) inhomogeneity of size, ζ = 20 ± 1 nm for all the three phases; (b) average mesh size, ξ0 = 2.6 ± 0.2 nm for sol and gel, and smaller mesh size, ξos = 1.2 ± 0.2 nm for coacervate; and (c) cross section of gelatin chains, Rc = 0.35 ± 0.04 nm. In addition, the structure factor data obtained from coacervating solution analyzed in the Guinier region, I(q) = exp(?q2Rg2/3), yielded value of typical radius of gyration of clusters, Rg ≈ 69 nm that indicated existence of triple‐helices of length, L ≈ 239 nm; (d) Frequency and temperature sweep measurements conducted on coacervate samples revealed two other length scales: (e) viscoelastic length, ξve = 14 ± 2 nm and (f) correlation length at melting, ξT = 500 ± 70 nm. Thus, existence of six distinct length scales, (a–f), ranging from 1.2 to 500 nm has been established in the coacervate phase of gelatin–ethanol–water system. Results are discussed within the framework of Landau‐Ginzburg treatment of dynamically asymmetric systems (Prog Theor Phys 1977, 57, 826; Phys Rev A 1991, 44, R817; J Phys II (France) 1992, 2, 1631). © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 1653–1667, 2006  相似文献   

7.
Melting-point and spherulite growth rate measurements for a sample of syndiotactic polypropylene (S = 0.716 and η = 0.356) were analyzed for the parameters characterizing crystal formation and growth: Tm = 159 ± 2°C, σe = 47 erg cm ?2, σ = 4.4 erg cm?2, and q = 5.6 kcal per mole of folds. The q and σe values place syndiotactic polypropylene in the group of “unhindered” polymers. Failure of the isotactic-polypropylene spherulite growth rate data to follow current theories of crystal growth precluded a comparison of crystal parameters of the two stereoisomers. At comparable degrees of supercooling, the absolute growth rates for the two forms are of the same order of magnitude and exhibit one or more crossover(s) in relative position.  相似文献   

8.
The heterogeneous isotopic exchange reactions in strontium polymolybdates of Sr2+ and MoO4 2- ions in the strontium nitrate and sodium molybdate solutions have been studied using 90Sr and 99Mo as tracers. Electrometric methods have been used to study the compositions of strontium molybdates obtained by adding strontium chloride to a progressively acidified solution of sodium molybdate. It has been found that the exchange fraction increases with increasing chain length of strontium polymolybdate. The exchange equilibrium constant (K ex) has been calculated between 298 and 348 K as well as DG°, DH° and DS°. The results indicate that Sr2+ cations have a much higher affinity for exchangers than MoO4 2- anions. By fitting the data to the Dubinin-Radushkevich (D-R) isotherm it has been shown that the exchange capacity (X m ) for both ions is affected by the ion adsorption process at low temperatures and by the ion exchange process at high temperatures. At high concentrations, the recrystallization process contributes to on the cation exchange but is ineffective on the anion exchange mechanism.  相似文献   

9.
A oxo‐bridged binuclear iron complex with chiral salen ligand [Fe2L2]O (H2L = R, R'‐N, N'‐bis (3,5‐di‐tert‐butylsalicylaldene)‐1, 2‐cyclohexanediamine) has been synthesized and the structure has been determined by X‐ray diffraction analysis. The title complex crystallizes in triclinic system with space group P‐1. Crystal data: a = 1.42555(14) nm, b = 1.54889 (15) nm, c = 1.83662 (17) run, = 103.873(2)°, β = 100.506(2)°, 7=101.840(2)°, V = 3.7371 nm3, Dc = 1.082 g/cm3 and Z = 2. In the complex, each iron center is penta‐coordination and in distorted square‐pyramidal environment. Two Fem atoms are intramolecularity bridged by an oxygen atom with Fe‐Fe nonbond distance of 0.3517(3) nm.  相似文献   

10.
The diffusion coefficients of polystyrene latex spheres and hematite particles in both Newtonian and elastic liquids have been measured using dynamic light scattering. The diffusion coefficients of the latex particles measured in glycerol/water (Newtonian) solutions obey Stokes–Einstein behaviour over a range of solvent viscosities and temperatures. Two apparent diffusion coefficients for the particles are measured in visco-elastic polyacrylamide and polyacrylate solutions and are designated Dfast and Dslow. The apparent fast diffusion coefficients measured in the elastic solutions show an increase to a maximum, above that measured in the solvent water, with increasing polyelectrolyte concentration. At higher polyelectrolyte concentrations the observed Dfast values decrease below the value obtained in the solvent water. Dfast increases with the scattering vector squared (q2) while Dslow, is independent of q2.  相似文献   

11.
With the low permeability and high swelling property, Gaomiaozi (GMZ) bentonite is regarded as the favorable candidate backfilling material for a potential repository. The diffusion behaviors of HTO in GMZ bentonite were studied to obtain effective diffusion coefficient (D e) and accessible porosity (ε) by through- and out-diffusion experiments. A computer code named Fitting for diffusion coefficient (FDP) was used for the experimental data processing and theoretical modeling. The D e and ε values were (5.2–11.2) × 10−11 m2/s and 0.35–0.50 at dry density from 1,800 to 2,000 kg/m3, respectively. The D e values at 1,800 kg/m3 was a little higher than that of at 2,000 kg/m3, whereas the D e value at 1,600 kg/m3 was significantly higher (approximately twice) than that of at 1,800 and 2,000 kg/m3. It may be explained that the diffusion of HTO mainly occurred in the interlayer space for the highly compacted clay (dry density exceeding 1,300 kg/m3). 1,800 and 2,000 kg/m3 probably had similar interlayer space, whereas 1,600 kg/m3 had more. Both D e and ε values decreased with increasing dry density. For compacted bentonite, the relationship of D e and ε could be described by Archie’s law with exponent n = 4.5 ± 1.0.  相似文献   

12.
Summary The crystal structure of (PPh4)2[ReO(OH)(CN)4]·5H2O has been determined from three-dimensional x-ray diffraction data. The light brown crystals are monoclinic, space group P21/n, with cell dimensionsa=16.753(2),b=19.928(2),c=15.338(2) Å and =101,894(1)°,z=4, Dm=1.45(1) g cm–3. The anisotropic refinement of the 6088 observed reflections converged to R=0.077.The [ReO(OH)(CN)4]2– ion has a distorted octahedral geometry. Bond distances: Re =1.70(1), Re–OH=1.90(1) and ReCav=2.12(2) Å. The Re atom is displaced by 0.08 Å out of the plane formed by the four carbon atoms towards the terminal oxo ligand.  相似文献   

13.
The polymeric surfactant with quaternary ammonium salt (PQ) was synthesized by cationic ring-open polymerization using boron trifluoride diethyletherate as cationic catalyst. The chemical structure and aggregation behavior of PQ were studied by 1H NMR, surface tension, static light scattering, dynamic laser light scattering, electrical conductivity, and fluorescence measurement. The results show the surface tension (γcmc) and critical micelle concentration (cmc) of PQ decrease with increasing of sodium chloride concentration. The cmc and γcmc values of PQ measured by electrical conductivity and fluorescence measurements mainly identify with that obtained by surface tension measurements. The thermodynamic parameters (DGm0 \Delta G_m^0 ,DHm0 \Delta H_m^0 ,DSm0 \Delta S_m^0 ) from electrical conductivity indicated that the micellization of PQ was mainly the process of entropy-driven. In addition, the results from the viscosity stability between hydrolyzed polyacrylamide (HPAM) and PQ showed that the viscosities of mixed system for HPAM and PQ are higher than the viscosity of HPAM.  相似文献   

14.

The reaction of MX2 (M = Co(II), Ni(II); X = Cl, Br) with 2-aminopyrimidine in aqueous acid yields compounds [(2-apmH)2MX4], (2-apmH)2[MX4], or (2-apmH2) [MX2(H2O)4]X2 (2-apmH = 2-aminopyrimidinium; 2-apmH2 = 2-aminopyrimidinium(2+)). All compounds have been characterized by single crystal X-ray diffraction. The compounds [(2-apmH)2MX4] with M = Co, X = Cl (1); M = Ni, X = Cl (3); and M = Ni, X = Br (4) are isomorphous and crystallize as nearly square planar MX4 units with the 2-apmH cations coordinated in the axial sites through the unprotonated ring nitrogen. (2-ApmH)2[CoBr4] (2) crystallizes as the salt with a nearly tetrahedral CuBr4 2- anion. (2-ApmH2)[NiBr2(H2O)4]Br2 (5) forms as a cocrystal of the neutral, six-coordinate nickel complex and (2-ampH2)Br2, stabilized by extensive hydrogen bonding. Crystal data (1): monoclinic, P21/c, a = 7.540(4), b = 12.954(4), c = 7.277(3) Å, β = 110.09(6), V = 667.4(5) Å3, Z = 2, Dcalc = 1.955 Mg/m3, μ = 2.079 mm-1, R = 0.0501 for [|I|≥2(I)]. For (2): triclinic, P-1, a = 7.720(2), b = 7.916(2), c = 14.797(3) Å, α = 97.264(3), β = 104.788(3), γ = 105.171(3)°, V = 825.3(3) Å3, Z = 2, Dcalc = 2.296 Mg/m3, μ = 10.715 mm-1, R = 0.0308 for [|I|≥2(I)]. For (3): monoclinic, P21/c, a = 7.595(3), b = 12.891(4), c = 7.204(3) Å, β = 111.07(3)°, V = 658.2 Å3, Z = 2, Dcalc = 1.982 Mg/m3, μ = 2.279 mm-1, R = 0.0552 for [|I|≥2(I)]. For (4): monoclinic, P21/c, a = 7.840(2), b = 13.358(4), c = 7.518(2) Å, β = 110.923(3)°, V = 938.6(3) Å3, Z = 2, Dcalc = 2.577 Mg/m3, μ = 12.18 mm-1, R = 0.0280 for [|I|≥2(I)]. For (5): orthorhombic, Pnma, a = 16.776(6), b = 11.943(4), c = 7.079(3) Å, V = 1418.2(9) Å3, Z = 4, Dcalc = 2.564 Mg/m3, μ = 12.639 mm-1, R = 0.0381 for [|I|≥2σ(I)].  相似文献   

15.
Optimal design and operation of bioreactors for insect cell culture is facilitated by functional relations providing quantitative information on cellular metabolite consumption kinetics, as well as on the specific cell growth rates (μG). Initial specific consumption rates of glucose, malate, and oxygen, and associated changes in μG, were measured forSpodoptera frugiperda clone 9 (Sf9) cells grown in batch suspension culture in medium containing 7–35 mM glucose, 0–16 mM malate, and 4–16 mM glutamine. The initial specific glucose consumption rate (q G ) could be described by a modified Michaelis-Menten equation treating malate as a “competitive” inhibitorK 1 = 6.5 mM) and glutamine as a “noncompetitive” inhibitorK I = 14 mM) ofq G , with aK m of 7.1 mM for glucose. All three carbon sources were found to increase μG in a saturable manner, and a modified Monod equation was employed to describe this relationship (μGmax = 0.047 h-1). The initial specific oxygen consumption rate (qO2) in Sf9 cells could be related to μG by the maintenance energy model, and it was calculated that, under typical culture conditions, about 15–20% of the cellular energy demand comes from functions not related to growth. Fitted parameters in mathematical expression for μg: K4, Monod constant for glucose (mM); K5, modified Monod constant for malate (mM); K6, Monod constant for glutamine (mM); mo2, specific consumption rate of oxygen by the cells under zero-growth conditions (nmol/cell/h); qF, initial specific fumarate production rate (nmol/cell/ h);q G , initial specific glucose consumption rate (nmol/cell/h); qGmax, maximum initial specific glucose consumption rate (nmol/cell/h);q M , initial specific malate consumption rate (nmol/cell/h); qo2, initial specific oxygen consumption rate (nmol/cell/h); Yo2, cell yield on oxygen (cells/nmol); μ, initial specific cell growth rate (h-1); μg, initial specific cell growth rate (h-1); μGmax, maximum initial specific cell growth rate (h-1).  相似文献   

16.
The processes of H3O+ production from alcohols (ethanol, 2‐propanol, 1‐propanol, 2‐butanol) and ethers (diethyl ether and ethyl methyl ether), and their deuterium‐substituted species, by intense laser fields (800 nm, 100 fs, ~1 × 1014 W/cm) were investigated through time‐of‐flight (TOF) mass spectrometry. H3O+ formation was observed for all these compounds except for ethyl methyl ether. From the analysis of TOF signals of H(3?n)DnO+ (n = 0, 1, 2, and 3) that have expanding tails with increasing flight time, it has been confirmed that the reaction proceeds through metastable dissociation from the intermediate species C2H(5?m)DmO+(m = 0–5). The common shape of the H(3?n)DnO+ signal profiles contains two major distributions in the time constant, i.e., fast and slow components of <50 ns and ~500 ns, respectively. The H(3?n)DnO+ branching ratio is interpreted to be the result of complete scrambling of four hydrogen atoms at the C? C site in C2H4‐OH+, and partial exchange (18–38%) of a hydrogen atom in the OH group with four other hydrogen atoms within 1 ns prior to H(3?n)DnO+ production. Ab initio calculations for the isomers and transition states of C2H5O+ were also performed, and the observed H(3?n)DnO+ production mechanism has been discussed. In addition, a stable isomer having a complex structure and two isomerization pathways were discovered to contribute to the H3O+ formation process. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

17.
After exploring the potential energy surfaces of MmCE2p (E=S−Te, M=Li−Cs, m=2, 3 and p=m-2) and MnCE3q (E=S−Te, M=Li−Cs, n=1, 2, q=n-2) combinations, we introduce 38 new global minima containing a planar hypercoordinate carbon atom (24 with a planar tetracoordinate carbon and 14 with a planar pentacoordinate carbon). These exotic clusters result from the decoration of V-shaped CE22− and Y-shaped CE32− dianions, respectively, with alkali counterions. All these 38 systems fulfill the geometrical and electronic criteria to be considered as true planar hypercoordinate carbon systems. Chemical bonding analyses indicate that carbon is covalently bonded to chalcogens and ionically connected to alkali metals.  相似文献   

18.
In this work, the g factors, dd transition band, local distortion, and their concentration dependences for impurity V4+ in 20Li2O–20PbO–45B2O3–(15 − x)P2O5:V2O5 (0 ≤ x ≤ 2.5 mol%) glasses are theoretically investigated by using perturbation formulas of g factors for a tetragonally compressed octahedral 3d1 cluster. In the light of the cubic polynomial concentration functions for cubic field parameter Dq, covalency factor N, and relative tetragonal compression ratio ρ, the calculated concentration dependences of dd transition band and g factors for V4+ show good agreement with the experimental data. With increasing x, N (≈0.7682–0.8165) displays the monotonously increasing trend, whereas ρ (≈6.5–4.2%) and Dq (≈1504.9–1481.1 cm−1) exhibit the decreasing tendencies. The above concentration dependences can be ascribed to the modifications of the V4+–O2− bonding and orbital admixtures around the impurity V4+ due to the effects of V2O5 doping on the stability of the glass network, the strength of local crystal fields, and the electron cloud distribution.  相似文献   

19.
A ternary binuclear complex of dysprosium chloride hexahydrate with m-nitrobenzoic acid and 1,10-phenanthroline, [Dy(m-NBA)3phen]2·4H2O (m-NBA: m-nitrobenzoate; phen: 1,10-phenanthroline) was synthesized. The dissolution enthalpies of [2phen·H2O(s)], [6m-HNBA(s)], [2DyCl3·6H2O(s)], and [Dy(m-NBA)3phen]2·4H2O(s) in the calorimetric solvent (VDMSO:VMeOH = 3:2) were determined by the solution–reaction isoperibol calorimeter at 298.15 K to be \Updelta\texts H\textmq \Updelta_{\text{s}} H_{\text{m}}^{\theta } [2phen·H2O(s), 298.15 K] = 21.7367 ± 0.3150 kJ·mol−1, \Updelta\texts H\textmq \Updelta_{\text{s}} H_{\text{m}}^{\theta } [6m-HNBA(s), 298.15 K] = 15.3635 ± 0.2235 kJ·mol−1, \Updelta\texts H\textmq \Updelta_{\text{s}} H_{\text{m}}^{\theta } [2DyCl3·6H2O(s), 298.15 K] = −203.5331 ± 0.2200 kJ·mol−1, and \Updelta\texts H\textmq \Updelta_{\text{s}} H_{\text{m}}^{\theta } [[Dy(m-NBA)3phen]2·4H2O(s), 298.15 K] = 53.5965 ± 0.2367 kJ·mol−1, respectively. The enthalpy change of the reaction was determined to be \Updelta\textr H\textmq = 3 6 9. 4 9 ±0. 5 6   \textkJ·\textmol - 1 . \Updelta_{\text{r}} H_{\text{m}}^{\theta } = 3 6 9. 4 9 \pm 0. 5 6 \;{\text{kJ}}\cdot {\text{mol}}^{ - 1} . According to the above results and the relevant data in the literature, through Hess’ law, the standard molar enthalpy of formation of [Dy(m-NBA)3phen]2·4H2O(s) was estimated to be \Updelta\textf H\textmq \Updelta_{\text{f}} H_{\text{m}}^{\theta } [[Dy(m-NBA)3phen]2·4H2O(s), 298.15 K] = −5525 ± 6 kJ·mol−1.  相似文献   

20.
Technetium(VII) extraction has been investigated to obtain useful information concerming the back-extraction of Tc(VII). Radioactive technetium-95m was used to determine the distribution ratio (D Tc) of95mTcO 4 for Tc(VII) extraction using Primene JMT (RNH2) in heptane solution. An emulsion formation did not occur in the ammonium carbonate system but occurred in the sodium hydroxide solutions. The extraction mechanism has also determined by using the slope analysis method to study the relationships between logD Tc and log [RNH2], and between logD Tc and pH.  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