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.
The spectral characteristics and the quantum yield of the fluorescence from the second excited singlet state S2 of the aromatic thioketone molecules xanthione (XS) and thioxanthione (TXS) have been determined in solution at room temperature and 77 K. In 3-methylpentane, the measured quantum yields are φf (295 K) = 5.1 × 10?3 and φf(77 K) = 1.0 × 10?2 for XS, and φf (295 K) = 1.5 × 10?3 and φf (77 K) = 2.5 × 10?3 for TXS. Using the Strickler-Berg expression for the radiative lifetime, the decay rate of S2 is derived. It is concluded that internal conversion S2 ? S1 is the dominating deactivation channel of S2 with k77 Knr(S2 ? S1) = 1.0 × 1010 s?1 for XS and k77 Knr (S2→S1) = 2.2 × 1010 s?1 for TXS. Between 295 and 77 K, φf increases by a factor of about 2 following an Arrhenius type expression. This temperature dependence of φf is considered to be intramolecular in nature and is attributed to a temperature sensitive rate constant knr(S2?S1) with an activation energy of 190 ± 20 cm?1 and a frequency factor knr = 3 × 1010 s?1 for the XS molecule in 3-methylpentane.  相似文献   

2.
It is shown under very general conditions that the intermediate scattering function for the generalized Rouse—Zimm model always takes the simple form G(K, t) α exp[?K2(kBT/f)t], when the scattering vector K becomes sufficiently large. (Here kB is Boltzmann's constant, T is the absolute temperature and f is the individual bead friction factor.) A microscopic formulation for the bulk modulus and friction factor density of a gel network is incorporated into the viscoelastic continuum model of Tanaka et al. The resulting expression for the apparent long-wavelength diffusion coefficient of the gel is DG = (kBT/f)2(1 - 2/Φ), where Φ is the network functionality.  相似文献   

3.
4.
5.
The low-temperature (5 to 310 K) heat capacity of cesium fluoroxysulfate, CsSO4F, has been measured by adiabatic calorimetry. At T = 298.15 K, the heat capacity Cpo(T) and standard entropy So(T) are (163.46±0.82) and (201.89±1.01) J · K?1 · mol?1, respectively. Based on an earlier measurement of the standard enthalpy of formation ΔHfo the Gibbs energy of formation ΔGfo(CsSO4F, c, 298.15 K) is calculated to be ?(877.6±1.6) kJ · mol?1. For the half-reaction: SO4F?(aq)+2H+(aq)+2e? = HSO4?(aq)+HF(aq), the standard electrode potential E at 298.15 K, is (2.47±0.01) V.  相似文献   

6.
A new method for determination of the conversion dependence of substantial initiation rate constants k i = f(C) in free-radical polymerization processes has been developed. On the basis of the known data on k i1 = f(C) dependences for initiator I1 and the kinetic analysis of a single trivial and simple experiment, this method allows one to calculate k i2 = f(C) function for any other initiator I2 under the same conditions (monomer, temperature). The reference experiment includes measurements of polymerization rates in the presence of initiator I1 in a wide conversion range from 0 to 100% and in the presence of I2, on the condition that the rates of initiation are equal w i1 = w i2, thus ensuring equal initial rates of polymerization. The above-described approach has been approved for the polymerization of styrene, methyl methacrylate, and vinyl acetate initiated with AIBN and benzoyl peroxide.  相似文献   

7.
马晓明  林国栋  张鸿斌 《催化学报》2006,27(11):1019-1027
 以自行制备的多壁碳纳米管(CNT)作为添加剂,制备共沉淀型CNT促进的Co-Mo-K硫化物基催化剂. 实验发现,与未添加CNT的催化剂相比,添加少量CNT可显著提高CO的加氢转化活性和生成低碳醇的选择性. 在5.0 MPa, 623 K, V(CO)∶V(H2)∶V(N2)=45∶45∶10, GHSV=3600 ml/(g·h)的反应条件下, Co1Mo1K0.3-10%CNT催化剂上CO的转化率达21.6%, 相应的总醇(C1~4醇)时空产率为241.5 mg/(g·h), 产物中C2+醇/C1醇=1.39 (C基选择性比). 添加少量CNT并不会导致Co1Mo1K0.3硫化物基催化剂上CO加氢反应表观活化能发生明显变化,但却导致工作态催化剂表面催化活性Mo物种(Mo4+)的摩尔百分率有所提高; 另一方面, CNT促进的催化剂对H2有更强的吸附活化能力,并能在相当大程度上抑制水煤气变换副反应的发生. 这些因素有利于提高催化剂的活性和选择性.  相似文献   

8.
The nucleation of the cationic miniemulsion polymerization of octamethylcyclotetrasiloxane (D4) has been investigated in detail in this work. The particle size was traced by dynamic laser scattering in the polymerization process using different concentration of dodecylbenzenesulfonic acid (DBSA) as both the initiator and surfactant. The results reveal the progressively decreasing particle sizes with the steadily increasing monomer conversion. The descending particle sizes can be explained by the existence of homogeneous nucleation of hydroxyl-terminated polydimethylsiloxane (PDMS) oligomers in addition to the primary droplet nucleation. A linear relationship between the natural logarithm of the average particle diameter (lnd(t)) and the natural logarithm of the monomer conversion (lnf(t)) is then deduced based on the theory, which has been validated by the experimental data. The slope of lnd(t)~lnf(t) line relies on the parameter of K which is the ratio of PDMS oligomers entering water phase to all PDMS oligomers produced. The influence of DBSA concentration on the K and the fraction of homogeneous nucleation (F(H)) has been then obtained. Both of the values of K and F(H) increase with the increasing DBSA concentrations.  相似文献   

9.
BiCl3 reacts with sodium 2,4,6-tris(trifluoromethyl)phenoxide (NaOR4f) in ether solution to produce an unusual condensation product in which three ORf functions have been coupled with the elimination of three fluorine atoms. The product is RfOC6H2(CF3)2C(O)ORf, which has been characterized spectroscopically and by X-ray crystallography (triclinic space group P 1; a = 8.958(1), b = 12.652(2), c = 13.722(2)Å, α = 89.596(8)°, β = 75.92(1)°, γ = 71.412(7)°, V = 1425.6(3)Å3, Z = 2). Bi(ORf)3 is believed to be an intermediate in this process. The carbonfluorine bond activation is not observed in the absence of BiCl3.  相似文献   

10.
The results of the X-ray structural study for the K4LiH3(SO4)4 single crystal are presented at a wide temperature range. The thermal expansion of the crystal using the X-ray dilatometry and the capacitance dilatometry from 8 to 500 K was carried out. The crystal structures data collection, solution and refinement at 125, 295, 443 and 480 K were performed. The K4LiH3(SO4)4 crystal has tetragonal symmetry with the P41 space group (Z=4) at room temperature as well as at the considered temperature range. The existence of a low-temperature, para-ferroelastic phase transition at about 120 K is excluded. The layered structure of the crystal reflects a cleavage plane parallel to (001) and an anisotropy of the protonic conductivity. The superionic high-temperature phase transition at TS=425 K is isostructural. Nevertheless, taking into account an increase of the SO4 tetrahedra libration above TS, a mechanism of the Grotthus type could be applied for the proton transport explanation.  相似文献   

11.
Azoethane was irradiated in the presence of carbon monoxide in the temperature range of 238 to 378 K. Kinetic parameters for the addition of ethyl radicals to carbon monoxide and for the decomposition of propionyl radicals were determined. The rate constants were found to be log k(cm3 mol?1 sec?1) = 11.19 - 4.8/θ and log k(sec?1) = 12.77 - 14.4/θ, respectively. Estimated thermochemical properties of the propionyl radical are ΔHf0 = -10.6 ± 1.0 kcal mol?1, S0 = 77.3 ± 1.0 cal K?1 mol?1, and D(C2H5CO? H) = 87.4 kcal mol?1.  相似文献   

12.
The thermal unimolecular decomposition of three vinylethers has been studied in a VLPP apparatus. The high-pressure rate constant for the retro-ene reaction of ethylvinylether was fit by log k (sec?1) = (11.47 + 0.25) - (43.4 ± 1.0)/2.303 RT at <T> = 900 K and that of t - butylvinylether by log k (sec?1) = (12.00 ± 0.27) - (38.4 ± 1.0)/2.303 RT at <T> = 800 K. No evidence for the competition of the higher energy homolytic bond-fission process could be obtained from the experimental data. The rate constant compatible with the C? O bond scission reaction in the case of benzylvinylether was log k (sec?1) = (16.63 ± 0.30) - (53.74 ± 1.0)/2.303 RT at <T> = 750 K. Together with ΔHf,3000(benzyl·) = 47.0 kcal/mol, the activation energy for this reaction results in ΔHf,3000(CH2CHO) = +3.0 ± 2.0 kcal/mol and a corresponding resonance stabilization energy of 3.2 ± 2.0 kcal/mol for 2-ethanalyl radical.  相似文献   

13.
Cuprocobaltites RBaCuCoO5 + gd(R = Nd, Sm, Gd) were prepared. Their unit cell parameters were determined, and thermal expansion, electrical conductivity (σ), and Seebeck coefficient (S) were studied in air in the range 300–1100 K. The compounds have tetragonal structures (space group P4/mmm, Z = 1). Their unit cell parameters are a = 0.3906(2) nm, c= 0.7648(7) nm for NdBaCuCoO5.21; a = 0.3904(2) nm, c = 0.7609(6) nm for SmBaCuCoO5.06; and a = 0.3891(2), c = 0.7592(6) nm for GdBaCuCoO5.02. The RBaCuCoO5 + δ cuprocobaltites at room temperature are p-type semiconductors, whose electrical conductivity and linear thermal expansion coefficient (LTEC) increase with increasing R3+ ionic radius, whereas the Seebeck coefficient decreases. The LTECs of RBaCuCoO5 + δ phases in the range 500–575 K increase by a factor of 1.2–1.5 because of the elimination of weakly bound oxygen. S = f(T) curves for RBaCuCoO5 + δ (R = Nd, Sm, Gd) feature maxima at 510 K for R = Sm and 365 K for R = Gd, probably, due to the change in the spin state of cobalt cations in these phases.  相似文献   

14.
Wenbin Zeng  Dieter Hoppe 《Tetrahedron》2005,61(13):3281-3287
(E)/(Z)-Isomeric allylic carbamate esters were deprotonated by n-butyllithium/(−)-sparteine in toluene. Trapping experiments with chlorotrimethylsilane afforded the α-substitution products, with (R)-configuration, revealing that the pro-S proton is removed predominantly to form the corresponding (S)-lithium·(−)-sparteine derivatives; kS/kR>15:1 and >7:1, respectively. A slow (S)→(R)-epimerization occurs at −78 °C (T1/2>60 min). The allylic double bond is stable to (Z)-(E) isomerization under these conditions.  相似文献   

15.
Hydrolysis reactions of silylurethanes Me3Si(p-XC6H4)NCOOEt (I) with X = Cl, H or Me in aqueous buffer solutions, with pH values from 1.94 to 10.00 were studied.The catalytic rate constants for the acid and base catalysed reactions and for the “non-catalysed” reaction k(H3O+), k(CH3COO?), k(H2PO4?), k(HPO42?), k(NH3), k(OH?) and k0 were evaluated from the pseudo first-order rate constants kexp determined by UV spectroscopy.The Brönsted coefficients for the base-catalysed reactions were obtained from the catalytic rate constants found and the known constants of dissociation K(HB+).The ρ values of the reactions could be derived from the σ constants given by Jaffé.The kientical results obtained are interpreted mechanistically and are believed to also have model character for other nucleophilic substitution reactions with silicon compounds.  相似文献   

16.
The rate coefficient, k1, for the gas‐phase reaction OH + CH3CHO (acetaldehyde) → products, was measured over the temperature range 204–373 K using pulsed laser photolytic production of OH coupled with its detection via laser‐induced fluorescence. The CH3CHO concentration was measured using Fourier transform infrared spectroscopy, UV absorption at 184.9 nm and gas flow rates. The room temperature rate coefficient and Arrhenius expression obtained are k1(296 K) = (1.52 ± 0.15) × 10?11 cm3 molecule?1 s?1 and k1(T) = (5.32 ± 0.55) × 10?12 exp[(315 ± 40)/T] cm3 molecule?1 s?1. The rate coefficient for the reaction OH (ν = 1) + CH3CHO, k7(T) (where k7 is the rate coefficient for the overall removal of OH (ν = 1)), was determined over the temperature range 204–296 K and is given by k7(T) = (3.5 ± 1.4) × 10?12 exp[(500 ± 90)/T], where k7(296 K) = (1.9 ± 0.6) × 10?11 cm3 molecule?1 s?1. The quoted uncertainties are 2σ (95% confidence level). The preexponential term and the room temperature rate coefficient include estimated systematic errors. k7 is slightly larger than k1 over the range of temperatures included in this study. The results from this study were found to be in good agreement with previously reported values of k1(T) for temperatures <298 K. An expression for k1(T), suitable for use in atmospheric models, in the NASA/JPL and IUPAC format, was determined by combining the present results with previously reported values and was found to be k1(298 K) = 1.5 × 10?11 cm3 molecule?1 s?1, f(298 K) = 1.1, E/R = 340 K, and Δ E/R (or g) = 20 K over the temperature range relevant to the atmosphere. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 635–646, 2008  相似文献   

17.
18.
The kinetic and the exchange energy functionals are expressed in the form T[ρ] = CTF∫ drρ5/3(r)ft(s) and K[ρ] = CD∫ drρ4/3(r)fK(s), where CTF = (3/10)(3π2)2/3 and CD = −(3/4)(3/π)4/3 are the Thomas-Fermi and the Dirac coefficients, respectively, and s = |∇ρ(r)|/Csρ4/3(r), with Cs = 2(3π2)1/3. These expressions are used to perform a comparison of fT(s) and fK(s) in terms of their generalized gradient expansion approximations. It is shown that fκ(s) and is congruent to fT(s) in the range characteristic of the interior regions of atoms and many solids and that the second-order gradient expansion of the kinetic energy provides a rather reasonable approximation to the generalized gradient expansion approximation of both the kinetic and the exchange energy functionals. © 1996 John Wiley & Sons, Inc.  相似文献   

19.
It is noted that the pK a values of organic acids can be calculated using the unique recurrence relation pK a(n + 1) = apK a(n) + b from the pK a values of other (usually the simplest and, consequently, better characterized) homologues of the same series. It is shown that this relation is valid within two taxonomic groups: insertion homologues of the ω-substituted acids X(CH2) n CO2H (n ≥ 1) and isomers that differ in the position of substituents X in their alkyl fragments, k-X(C n H2n )CO2H (n ≥ 1, 1 ≤ kn + 1). It is concluded that this algorithm is a consequence of the unique mathematical properties of recurrence relations.  相似文献   

20.
In most solid state reactions the reaction velocity can be described as a product of two functionsK(T) andf(1?α) whereT is the temperature and α the degree of conversion of the solid reactant. The physical interpretation of these functions is discussed, and a systematic method is described by whichf(1?α) of a reaction is identified from its kinetic data.K(T) and the reaction mechanism are then determined. This method has been successfully applied to analyse the kinetics of the thermal decomposition of silver azide.  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