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.
Statistical analysis of quantitative structure—property relationships (QSPR) for the phase transition gel—liquid crystal (main phase transition) temperatures (T m) in hydrated vesicle bilayers of diacyl phosphatidylcholines (PC) and diacyl phosphatidylethanolamines (PE) with saturated acyl chains was carried out. Multiple regression equations relating the T m values to the volumes of the hydrocarbon fragments of the sn-1 and sn-2 acyl chains show that the acyl groups influence the transiton temperatures T m in different manner, the effect of the sn-2 group being predominant. The transition temperatures T m of saturated phosphatidylethanolamines and diacyl phosphatidylglycerols can be predicted using the T m values calculated for the corresponding diacyl phosphatidylcholines.  相似文献   

2.
We studied the formation and stability of vesicles consisting of 1,2-dioleoyl-3-trimethylammonium propane (DOTAP) and phosphatidylcholines by electron spin resonance (ESR) analysis and observation of their hemolytic activities. In contrast with previous findings on dimethyldialkylammoniums, DOTAP formed vesicles at 37 degrees C with phosphatidylcholines containing either saturated acyl chains such as dimyristoylphosphatidylcholine (DMPC) or unsaturated acyl chains such as dilinoleoylphosphatidylcholine (DLPC). Phosphatidylcholines made the bilayer more rigid and significantly reduced the hemolytic activity of DOTAP. In the presence of equimolar concentration of DOTAP and phosphatidylcholines, formation of tightly aggregated structures of several erythrocytes was observed, as previously reported for the vesicles containing dimethyldipalmitylammonium. These findings indicate that DOTAP vesicles were stabilized by phosphatidylcholines with either saturated acyl chains or unsaturated acyl chains, and the interaction with the lipid bilayer of biological membranes as cationic vesicles became prominent with minimal membrane damage by DOTAP monomers.  相似文献   

3.
An extension to the rotating-sector method, which is usually applied to determine propagation and termination rate constants, is presented. The analytical treatment developed accounts for the simultaneous presence of a thermal initiation and of a first-order termination process. The applicability of the rotating-sector method is thus extended to situations where the rate in dark is higher than 5% of the rate in the presence of light, and more accurate estimates of the rate constants are obtained than before for any values of the “dark” rate. A previously published experiment on the application of the rotating-sector method to the autoxidation of styrene was reanalyzed. The estimates obtained for the propagation and the termination rate constants were 11% and 19% higher than the previous estimates, respectively. Finally, the improved rotating-sector method was also applied to the experimental determination of propagation (kp) and termination rate constants (2×kt) for both 1-palmitoyl-2-linoleoyl-sn-glycero-3-phosphocholine (PLPC) and 1,2-dilinoleoyl-sn-glycero-3-phosphocholine (DLPC) liposomes. The following results were obtained at 37°C: for PLPC kp =16.6 M−1s−1, and 2×kt=1.27×105 M−1s−1; for DLPC kp(intermolecular)=(13.3–13.9) M−1s−1, kp(intramolecular)=(4.7–5.4) s−1, and 2×kt=(0.99–1.05)×105 M−1s−1. The separation of the intermolecular and intramolecular propagation rate constants for DLPC was made possible both by a special adaptation of the rotating-sector equations to substrates with two oxidizable moieties, and by the experimental determination of the ratio between partially oxidized DLPC molecules (only one acyl is oxidized) and fully oxidized DLPC molecules (both acyls are oxidized). © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 753–767, 1998  相似文献   

4.
Br-atom atomic resonance absorption spectrometry (ARAS) has been developed and applied to measure thermal decomposition rate constants for CF3Br (+ Kr)→CF3+Br (+ Kr) over the temperature range, 1222–1624 K. The Br-atom curve-of-growth (145<λ<163 nm) was determined using this reaction. For [Br]≤1×1012 molecules cm−3, absorbance, (ABS)=1.410×10−13 [Br], yielding σ=1.419×10−14 cm2. The curve-of-growth was then used to convert (ABS) to Br-atom profiles which were then analyzed to give measured rate constants. These can be expressed in second-order by k1=8.147×10−9 exp(−24488 K/T) cm3 molecule−1 s−1 (±33%, 1222≤T≤1624 K). A unimolecular theoretical approach was used to rationalize the data. Theory indicates that the dissociation rates are closer to second- than to first-order, i.e., the magnitudes are 30–53% of the low-pressure-limit rate constants over 1222–1624 K and 123–757 torr. With the known, E0=ΔH00=70.1 kcal mole−1, the optimized theoretical fit to the ARAS data requires 〈ΔEdown=550 cm−1. These conclusions are consistent with recently published data and theory from Kiefer and Sathyanarayana. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 859–867, 1998  相似文献   

5.
3‐Methyl‐3‐(3‐pentyl)‐1,2‐dioxetane 1 and 3‐methyl‐3‐(2,2‐dimethyl‐1‐propyl)‐1,2‐dioxetane 2 were synthesized in low yield by the α‐bromohydroperoxide method. The activation parameters were determined by the chemiluminescence method (for 1 ΔH‡ = 25.0 ± 0.3 kcal/mol, ΔS‡ = −1.0 entropy unit (e.u.), ΔG‡ = 25.3 kcal/mol, k1 (60°C) = 4.6 × 10−4s−1; for 2 ΔH‡ = 24.2 ± 0.2 kcal/mol, ΔS‡ = −2.0 e.u., ΔG‡ = 24.9 kcal/mol, k1 (60°C) = 9.2 × 10−4s−1. Thermolysis of 1–2 produced excited carbonyl fragments (direct production of high yields of triplets relative to excited singlets) (chemiexcitation yields for 1: ϕT = 0.02, ϕS ≤ 0.0005; for 2: ϕT = 0.02, ϕS ≤ 0.0004). The results are discussed in relation to a diradical‐like mechanism. © 2001 John Wiley & Sons, Inc. Heteroatom Chem 12:176–179, 2001  相似文献   

6.
The reaction of NO with the peroxy radical CFCl2CH2O2, and with CH3CFClO2 was investigated at 8(SINGLEBOND)20 torr and 263(SINGLEBOND)321 K by UV flash photolysis of CFCl2CH3/O2/NO gas mixtures. The kinetics were determined from observations of the growth rate of the CFCl2CH2O radical and the decay rate of NO by time-resolved mass spectrometry. The temperature dependence of the bimolecular rate coefficients, with their statistical uncertainties, can be expressed as (2.9 ± 0.7) e(435±96)/T × 10−12 cm3 molecule −1s−1, or (1.3 ± 0.2) (T/300)&minus(1.5±0.2) × 10−11 cm3 molecule−1 s−1 for NO + CFCl2CH2O2, and (3.3 ± 0.6)e(516±73)/T × 10−12 cm3 molecule−1 s−1, or (2.0 ± 0.3) (T/300)&minus(1.8±0.3) × 10−11 cm3 molecule−1 s−1 for NO + CH3CFClO2. No pressure dependence of the rate coefficients could be detected over the 8(SINGLEBOND)20 torr range investigated. © 1996 John Wiley & Sons, Inc.  相似文献   

7.
Ligand substitution kinetics for the reaction [PtIVMe3(X)(NN)]+NaY=[PtIVMe3(Y)(NN)]+NaX, where NN=bipy or phen, X=MeO, CH3COO, or HCOO, and Y=SCN or N3, has been studied in methanol at various temperatures. The kinetic parameters for the reaction are as follows. The reaction of [PtMe3(OMe)(phen)] with NaSCN: k1=36.1±10.0 s−1; ΔH1=65.9±14.2 kJ mol−1; ΔS1=6±47 J mol−1 K−1; k−2=0.0355±0.0034 s−1; ΔH−2=63.8±1.1 kJ mol−1; ΔS−2=−58.8±3.6 J mol−1 K−1; and k−1/k2=148±19. The reaction of [PtMe3(OAc)(bipy)] with NaN3: k1=26.2±0.1 s−1; ΔH1=60.5±6.6 kJ mol−1; ΔS1=−14±22 J mol−1K−1; k−2=0.134±0.081 s−1; ΔH−2=74.1±24.3 kJ mol−1; ΔS−2=−10±82 J mol−1K−1; and k−1/k2=0.479±0.012. The reaction of [PtMe3(OAc)(bipy)] with NaSCN: k1=26.4±0.3 s−1; ΔH1=59.6±6.7 kJ mol−1; ΔS1=−17±23 J mol−1K−1; k−2=0.174±0.200 s−1; ΔH−2=62.7±10.3 kJ mol−1; ΔS−2=−48±35 J mol−1K−1; and k−1/k2=1.01±0.08. The reaction of [PtMe3(OOCH)(bipy)] with NaN3: k1=36.8±0.3 s−1; ΔH1=66.4±4.7 kJ mol−1; ΔS1=7±16 J mol−1K−1; k−2=0.164±0.076 s−1; ΔH−2=47.0±18.1 kJ mol−1; ΔS−2=−101±61 J mol−1 K−1; and k−1/k2=5.90±0.18. The reaction of [PtMe3(OOCH)(bipy)] with NaSCN: k1 =33.5±0.2 s−1; ΔH1=58.0±0.4 kJ mol−1; ΔS1=−20.5±1.6 J mol−1 K−1; k−2=0.222±0.083 s−1; ΔH−2=54.9±6.3 kJ mol−1; ΔS−2=−73.0±21.3 J mol−1 K−1; and k−1/k2=12.0±0.3. Conditional pseudo-first-order rate constant k0 increased linearly with the concentration of NaY, while it decreased drastically with the concentration of NaX. Some plausible mechanisms were examined, and the following mechanism was proposed. [Note to reader: Please see article pdf to view this scheme.] © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 523–532, 1998  相似文献   

8.
In this work, we discussed the stochastic thermodynamics of mesoscopic electron transfer reactions between ions and electrodes. With a relationship between the reaction rate constant and the electrode potential, we find that the heat dissipation βq equals to the dynamic irreversibility of the reaction system minus an internal entropy change term. The total entropy change Δst is defined as the summation of the system entropy change Δs and the heat dissipation βq such that Δsts+βq. Even though the heat dissipation depends linearly on the electrode potential, the total entropy change is found to satisfy the fluctuation theorem < (e)st>=1, and hence a second law-like inequality reads <Δst>≥0. Our study provides a practical methodology for the stochastic thermodynamics of electrochemical reactions, which may find applications in biochemical and electrochemical reaction systems.  相似文献   

9.
trans-3-Methyl-4-(p-anisyl)-1,2-dioxetane 1, trans-3-methyl-4-(o-anisyl)-1,2-dioxetane 2 , 3-methyl-3-benzyl-1,2-dioxetane 3 , and 3-methyl-3-p-methoxybenzyl-1,2-dioxetane 4 were synthesized in low yield by the β-bromo hydroperoxide method. The activation parameters were determined by the chemiluminescence method (for 1 ΔG≠ = 22.8 ± 0.3 kcal/mol, Δ≠ = 22.2, ΔS≠ = −1.7 e.u., k60 = 7.6 × 10−3s−1; for 2 ΔG≠ + 23.6 ± 0.3 kcal/mol, ΔH≠ = 22.8, ΔS≠ = −2.2 e.u., k60 = 2.5 × 10−3S−1; for 3 ΔG≠ = 24.0 ± 0.4 kcal/mol, ΔH≠ = 23.1, ΔS≠ = −2.7 e.u., k60 = 1.2 × 10−3S−1; for 4 ΔG≠ = 24.0 ± 0.2 kcal/mol, ΔH≠, = 23.2, ΔS≠, = −2.4 e.u., k60 = 1.2 × 10−3s−1). Thermolysis of 1–4 produced excited carbonyl fragments (direct production of high yields of triplets relative to excited singlets) [chemiexcitation yields ϕT, ϕS, respectively: for 1 0.02, 0.0001; for 2 0.02, 0.0001; for 3 0.03, 0.0002; for 4 0.02, 0.0001]. The effect of paramethoxyaryl substitution was consistent with electronic effects. The ortho substitution in 2 resulted in an increase in stability of the dioxetane, opposite that observed for an electronic effect. The results are discussed in relation to a diradical-like mechanism.  相似文献   

10.
Polymerization of lipid assemblies may be usefully employed to alter the properties of the assemblies. The possible locations of the reactive group in the lipids include (1) the chain terminus, (2) the head group, and (3) near the lipid backbone. The third strategy yields polymerized assemblies which retain their head group functionality and lipid chain motion. We have designed and synthesized new members of this later category by the use of 2-methylene-substituted acyl chains. The main transition temperature (Tm) from gel to liquid crystalline phase of hydrated bilayers of 1-palmitoyl-2-(2-methylene)palmitoyl-sn-glycero-3-phosphocholine ( 1 ) and the disubstituted 1,2-bis(2-methylenepalmitoyl)-sn-glycero-3-phosphocholine ( 2 ) were 33.6 and 25.3°C, respectively. The Tm of the mono-substituted 1-oleoyl-2-(2-methylene)palmitoyl-sn-glycero-3-phosphocholine ( 3 ) bilayers was detected in a range from ?15 to ?10°C by x-ray diffraction. Hydrated bilayers of each individual lipid were successfully polymerized with a water-soluble initiator, azobis(2-amidinopropane) dihydrochloride (AAPD). These results indicate the lipid 2-methylene groups are accessible to the water interface. Thermal polymerization of the mono-substituted lipids in aqueous suspensions with AAPD, yielded oligomers. However the bis-2-methylene PC ( 2 ) was successfully polymerized to yield stabilized crosslinked bilayers. © 1994 John Wiley & Sons, Inc.  相似文献   

11.
Poly(3-hydroxyoctanoate-co-3-hydroxy-10-undecenoate)s (PHOUs) with controlled amounts of unsaturated repeating units were epoxidized to various extents with m-chloroperbenzoic acid (MCPBA) in homogeneous solution. The epoxidation reaction was second order, with an initial rate constant of 1.1 × 10−3Lmol−1.s−1 at 20°C, regardless of the unsaturated unit content in PHOU. No substantial change in either molecular weight or molecular weight distribution occurred as a result of epoxidation, but the melt transition temperature and enthalpy of melting both decreased as the unsaturated groups were increasingly converted into epoxide groups. In contrast, the glass transition temperature (Tg) increased by approximately 0.25°C for each 1 mol % of epoxidation, irrespective of the composition of the PHOU. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. A Polym. Chem. 36: 2381–2387, 1998  相似文献   

12.
Three-membered ring (3MR) forming processes of X(SINGLE BOND)CH2(SINGLE BOND)CH2(SINGLE BOND)F and CH2(SINGLE BOND)C((SINGLE BOND)Y)(SINGLE BOND)CH2(SINGLE BOND)F (X(DOUBLE BOND)CH2, O, or S and Y(DOUBLE BOND)0 or S) through a gas phase neighboring group mechanism (SNi) are studied theoretically using the ab initio molecular orbital method with the 6–31+G* basis set. When electron correlation effects are considered, the activation (ΔG) and reaction energies (ΔG0) are lowered by ca. 10 kcal mol−1, indicating the importance of the electron correlation effect in these reactions. The contribution of entropy of activation (−TΔS) at 298 K to ΔG is very small, and the reactions are enthalpy controlled. The ΔG and ΔG0 values for these ring closure processes largely depend on the stabilities of the reactants and the heteroatom acting as a nucleophilic center. The Bell–Evans–Polanyi principle applies well to all these reaction series. © 1997 John Wiley & Sons, Inc. J Comput Chem 18 : 1773–1784, 1997  相似文献   

13.
CH3NH2 thermal decomposition is shown to provide a suitable NH2 radical source for spectroscopic and kinetic shock tube studies. Using this precursor, the absorption coefficient of the NH2 radical at a detection wavelength of 16739.90 cm−1 has been determined. In the temperature range 1600–2000K the low‐pressure absorption coefficient is described by the polynominal equation: kNH2=3.953×1010/T 3+7.295×105/T 2−1.549×103/T [atm−1 cm−1] The uncertainty of the determined absorption coefficient is estimated to be ±10%. The rate of the thermal decomposition reaction CH3NH2+M → CH3+NH2+M is determined over the temperature range 1550–1900 K and at pressures near 1.6 atm. The rate coefficient was found to be: k1=2.51×1016 exp(−28430/T) [cm3 mol−1 s−1] The uncertainty of the determined rate coefficients is estimated to be ±20%. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 323–330, 1999  相似文献   

14.
Densities (ρ) at different temperatures from 303.15 to 318.15 K, speeds of sound (u) and viscosities (η) at 303.15 K were measured for the binary mixtures of cyclohexanol with 2-chlorotoluene, 3-chlorotoluene and 4-chlorotoluene over the entire range of composition. The excess volumes (VE) for the mixtures have been computed from the experimental density data. Further, the deviation in isentropic compressibilities (Δκs) and deviation in viscosities (Δη) for the binary mixtures have been calculated from the speed of sound and viscosity data, respectively. The VE values and Δκs values were positive and Δη data were negative for all the mixtures over the whole range of composition at the measured temperatures. The calculated excess functions VE, Δκs and Δη were fitted to Redlich–Kister equation. The excess functions have been discussed in terms of molecular interactions between component molecules of the binary mixtures.  相似文献   

15.
The temperature dependence of the rate coefficients for the OH radical reactions with iso-propyl acetate (k1), iso-butyl acetate (k2), sec-butyl acetate (k3), and tert-butyl acetate (k4) have been determined over the temperature range 253–372 K. The Arrhenius expressions obtained are: k1=(0.30±0.03)×10−12 exp[(770±52)/T]; k2=(109±0.14)×10−12 exp[(534±79)/T]; k3=(0.73±0.08)×10−12 exp[(640±62)/T]; and k4=(22.2±0.34)×10−12 exp[−(395±92)/T] (in units of cm3 molecule−1 s−1). At room temperature, the rate constants obtained (in units of 10−12 cm3 molecule−1 s−1) were as follows: iso-propyl acetate (3.77±0.29); iso-butyl acetate (6.33±0.52); sec-butyl acetate (6.04±0.58); and tert-butyl acetate (0.56±0.05). Our results are compared with the previous determinations and discussed in terms of structure-activity relationships. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet: 29: 683–688, 1997.  相似文献   

16.
3‐Methyl‐3‐(3‐pentyl)‐1,2‐dioxetane 1 and 3‐methyl‐3‐(2,2‐dimethyl‐1‐propyl)‐1,2‐dioxetane 2 were synthesized in low yield by the α‐bromohydroperoxide method. The activation parameters were determined by the chemiluminescence method (for 1 ΔH‡ = 25.0 ± 0.3 kcal/mol, ΔS‡ = −1.0 entropy unit (e.u.), ΔG‡ = 25.3 kcal/mol, k1 (60°C) = 4.6 × 10−4s−1; for 2 ΔH‡ = 24.2 ± 0.2 kcal/mol, ΔS‡ = −2.0 e.u., ΔG‡ = 24.9 kcal/mol, k1 (60°C) = 9.2 × 10−4s−1. Thermolysis of 1–2 produced excited carbonyl fragments (direct production of high yields of triplets relative to excited singlets) (chemiexcitation yields for 1: ϕT = 0.02, ϕ ≤ 0.0005; for 2: ϕT = 0.02, ϕS ≤ 0.0004). The results are discussed in relation to a diradical‐like mechanism. © 2001 John Wiley & Sons, Inc. Heteroatom Chem 12:459–462, 2001  相似文献   

17.
The chlorine transfer reaction between 3-azabicyclo[3,3,0]octane “AZA” and chloramine was studied over pH 8–13 in order to follow both the amination and halogenation properties of NH2Cl. The results show the existence of two competitive reactions which lead to the simultaneous formation of N-amino- and N-chloro- 3-azabicyclo[3,3,0]octane by bimolecular kinetics. The halogenation reaction is reversible and the chlorine derivative obtained, which is thermolabile and unstable in the pure state, was identified by electrospray mass spectrometry. These phenomena were quantified by a reaction between neutral species according to an apparent SN2-type mechanism for the amination process and a ionic mechanism involving a reaction between chloramine and protonated amine for the halogenation process. Amination occurs only in strongly basic solutions (pH ≥ 13) while chlorination occurs at lower pH's (pH ≤ 8). At intermediate pH's, a mixture of these two compounds is obtained. The relative proportions of the products are a function of intrinsic rate constants, pH and pKa of the reactants. The rate constants and thermodynamic activation parameters are the following: k1 = 45.5 × 10−3 M−1 s−1; ΔH10# = 59.8 kJ mol−1; ΔS10# = − 86.5 J mol−1 K−1 for amination; k2 = 114 × 10−3 M−1 s−1; ΔH20# = 63.9 kJ mol−1; and ΔS20# = − 48.3 J mol−1 K−1 for chlorination. The ability of an interaction corresponding to a specific (NH3Cl+/RR′NH) or general (NH2Cl/RR′NH) acid catalysis has been also discussed. © 1997 John Wiley & Sons, Inc.  相似文献   

18.
After an exhaustive study of the system ammonia–dimethylchloramine in liquid ammonia, it was interesting to compare the reactivity of this system in liquid ammonia with the same system in an aqueous medium. Dimethylchloramine prepared in a pure state undergoes dehydrohalogenation in an alkaline medium: the principal products formed are N-methylmethanimine, 1,3,5-trimethylhexahydrotriazine, formaldehyde, and methylamine. The kinetics of this reaction was studied by UV, GC, and HPLC as a function of temperature, initial concentrations of sodium hydroxide, and chlorinated derivative. The reaction is of the second order and obeys an E2 mechanism (k1 = 4.2 × 10−5 M−1 s−1, ΔH○# = 82 kJ mol−1, ΔS○# = −59 J mol−1 K−1). The oxidation of unsymmetrical dimethylhydrazine by dimethylchloramine involves two consecutive processes. The first step follows a first-order law with respect to haloamine and hydrazine, leading to the formation of an aminonitrene intermediate (k2 = 150 × 10−5 M−1 s−1). The second step corresponds to the conversion of aminonitrene into formaldehyde dimethylhydrazone at pH 13). This reaction follows a first-order law (k3 = 23.5 × 10−5 s−1). The dimethylchloramine–ammonia interaction corresponds to a SN2 bimolecular mechanism (k4 = 0.9 × 10−5 M−1 s−1, pH 13, and T = 25°C). The kinetic model formulated on the basis of the above reactions shows that the formation of the hydrazine in an aqueous medium comes under strong competition from the dehydrohalogenation of dimethylchloramine and the oxidation of the hydrazine formed by the original chlorinated derivative. A global model that explains the mechanisms both in an anhydrous and in an aqueous medium was elaborated. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 340–351, 2008  相似文献   

19.
A laser photolysis–long path laser absorption (LP‐LPLA) experiment has been used to determine the rate constants for H‐atom abstraction reactions of the dichloride radical anion (Cl2) in aqueous solution. From direct measurements of the decay of Cl2 in the presence of different reactants at pH = 4 and I = 0.1 M the following rate constants at T = 298 K were derived: methanol, (5.1 ± 0.3)·104 M−1 s−1; ethanol, (1.2 ± 0.2)·105 M−1 s−1; 1‐propanol, (1.01 ± 0.07)·105 M−1 s−1; 2‐propanol, (1.9 ± 0.3)·105 M−1 s−1; tert.‐butanol, (2.6 ± 0.5)·104 M−1 s−1; formaldehyde, (3.6 ± 0.5)·104 M−1 s−1; diethylether, (4.0 ± 0.2)·105 M−1 s−1; methyl‐tert.‐butylether, (7 ± 1)·104 M−1 s−1; tetrahydrofuran, (4.8 ± 0.6)·105 M−1 s−1; acetone, (1.41 ± 0.09)·103 M−1 s−1. For the reactions of Cl2 with formic acid and acetic acid rate constants of (8.0 ± 1.4)·104 M−1 s−1 (pH = 0, I = 1.1 M and T = 298 K) and (1.5 ± 0.8) · 103 M−1 s−1 (pH = 0.42, I = 0.48 M and T = 298 K), respectively, were derived. A correlation between the rate constants at T = 298 K for all oxygenated hydrocarbons and the bond dissociation energy (BDE) of the weakest C‐H‐bond of log k2nd = (32.9 ± 8.9) − (0.073 ± 0.022)·BDE/kJ mol−1 is derived. From temperature‐dependent measurements the following Arrhenius expressions were derived: k (Cl2 + HCOOH) = (2.00 ± 0.05)·1010·exp(−(4500 ± 200) K/T) M−1 s−1, Ea = (37 ± 2) kJ mol−1 k (Cl2 + CH3COOH) = (2.7 ± 0.5)·1010·exp(−(4900 ± 1300) K/T) M−1 s−1, Ea = (41 ± 11) kJ mol−1 k (Cl2 + CH3OH) = (5.1 ± 0.9)·1012·exp(−(5500 ± 1500) K/T) M−1 s−1, Ea = (46 ± 13) kJ mol−1 k (Cl2 + CH2(OH)2) = (7.9 ± 0.7)·1010·exp(−(4400 ± 700) K/T) M−1 s−1, Ea = (36 ± 5) kJ mol−1 Finally, in measurements at different ionic strengths (I) a decrease of the rate constant with increasing I has been observed in the reactions of Cl2 with methanol and hydrated formaldehyde. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 169–181, 1999  相似文献   

20.
The reaction of Cl atoms with a series of C2–C5 unsaturated hydrocarbons has been investigated at atmospheric pressure of 760 Torr over the temperature range 283–323 K in air and N2 diluents. The decay of the hydrocarbons was followed using a gas chromatograph with a flame ionization detector (GC‐FID), and the kinetic constants were determined using a relative rate technique with n‐hexane as a reference compound. The Cl atoms were generated by UV photolysis (λ ≥ 300 nm) of Cl2 molecules. The following absolute rate constants (in units of 10−11 cm3 molecule−1 s−1, with errors representing ±2σ) for the reaction at 295 ± 2 K have been derived from the relative rate constants combined to the value 34.5 × 10−11 cm3 molecule−1 s−1 for the Cl + n‐hexane reaction: ethene (9.3 ± 0.6), propyne (22.1 ± 0.3), propene (27.6 ± 0.6), 1‐butene (35.2 ± 0.7), and 1‐pentene (48.3 ± 0.8). The temperature dependence of the reactions can be expressed as simple Arrhenius expressions (in units of 10−11 cm3 molecule−1 s−1): kethene = (0.39 ± 0.22) × 10−11 exp{(226 ± 42)/T}, kpropyne = (4.1 ± 2.5) × 10−11 exp{(118 ± 45)/T}, kpropene = (1.6 ± 1.8) × 10−11 exp{(203 ± 79)/T}, k1‐butene = (1.1 ± 1.3) × 10−11 exp{(245 ± 90)/T}, and k1‐pentene = (4.0 ± 2.2) × 10−11 exp{(423 ± 68)/T}. The applicability of our results to tropospheric chemistry is discussed. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 478–484, 2000  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