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.
The kinetic behavior of the thiyl radical in the solution containing polyisoprenes and polybutadienes has been studied by the flash photolysis method. For benzothiazole-2-thiyl radical, the addition rate constants toward these polymers and the model compounds of the polymers were evaluated. The relative reverse rate constants and equilibrium constants were also estimated. The addition rate constants decrease with an increase in the degree of polymerization; the ratio of the addition rate constant for polyisoprene (3.1 × 104 M?1 s?1 (in monomer unit); Mv = 674,000) to that for 2-methyl-2-butene (1.5 × 105 M?1 s?1) is about 1/5. This indicates that the polymer chain effect appears in the free-radical addition reaction. The relative reverse rate constants for the polymers are also smaller than those for 2-methyl-2-butene, suggesting a kind of polymer effects; i.e., it can be presumed that the bonded-thiyl radicals migrate very rapidly to the neighboring double bonds in the polymer. Significant differences in the rate parameters were observed between polyisoprene and polybutadiene, between cis- and trans-polyisoprenes, and between 1,4- and 1,2-polybutadienes.  相似文献   

2.
The 2,6‐spirodicyclohexyl substituted nitroxide, cyclohexane‐1‐spiro‐2′‐(3′,5′‐dioxo‐4′‐benzylpiperazine‐1′‐oxyl)‐6′‐spiro‐1″‐cyclohexane (BODAZ), was investigated as a mediator for controlled/living free‐radical polymerization of styrene. The values of the number‐average molecular weight increased linearly with conversion, but the polydispersities were higher than for the corresponding 2,2,6,6‐tetramethylpiperidinyl‐1‐oxy (TEMPO) and 2,5‐bis(spirocyclohexyl)‐3‐benzylimidazolidin‐4‐one‐1‐oxyl (NO88Bn) mediated systems at approximately 2.2 and 1.6 at 100 and 120 °C, respectively. These results were reflected in the rate coefficients obtained by electron spin resonance spectroscopy; at 120 °C, the values of the rate coefficients for polystyrene‐BODAZ alkoxyamine dissociation (kd), combination of BODAZ and propagating radicals (kc), and the equilibrium constant (K) were 1.60 × 10?5 s?1, 5.19 × 106 M?1 s?1, and 3.08 × 10?12 M, respectively. The value of kd was approximately one and two orders of magnitude lower, and that of K was approximately 20 and 7 times lower than for the NO88Bn and TEMPO adducts. These results are explained in terms of X‐ray crystal structures of BODAZ and NO88Bn; the six‐membered ring of BODAZ deviates significantly from planarity as compared to the planar five‐membered ring of NO88Bn and possesses a benzyl substituent oriented away from the nitroxyl group leading to a seemingly more exposed oxyl group, which resulted in a higher kc and a lower kd than NO88Bn. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 3892–3900, 2003  相似文献   

3.
The kinetics of reactions involving the ground-state azide radical, N3 (X2Πg, have been investigated in a discharge-flow system using mass spectrometric detection with molecular-beam sampling. The following rate constants have been determined at 295 K: Cl + N3Cl → Cl2 + N3,k295 = (1.78 ± 0.26) × 10?12 cm3 s?1 (1σ): N3 + NO → N2O + N2, k295 = (1.19 ± 0.31) × 10.?12 cm3 s?1 (1σ). A method for determining absolute N3 radical concentration is reported.  相似文献   

4.
Butylated hydroxytoluene (BHT) is a synthetic antioxidant widely used as a preservative in foods containing fat, the petroleum industry, and pharmaceuticals. Herein, a systematic investigation of reaction mechanisms and kinetics of BHT initiated by HO? and HOO? radicals in physiological environments has been reported for the first time. The overall rate coefficients were determined according to the QM-ORSA (quantum mechanics-based test for overall free radical scavenging activity) procedure at the M06-2X/6-311++G(d,p) level of theory. The radical adduct formation has been found to be the decisive mechanism for HO? scavenging in both polar and lipid media with koverall = 1.39 × 1012 and 1.13 × 1011 M–1 s–1, respectively. While this mechanism has no contribution to HOO? scavenging in physiological environments which mainly takes place via the hydrogen atom transfer mechanism (koverall = 2.51 × 105 and 1.70 × 104 M–1 s–1 in water and lipid media, respectively). The obtained findings are consistent with the available experimental data, which validates the accuracy of the calculations.  相似文献   

5.
The flash photolysis of aqueous solutions of rhodizonic and croconic acids has been studied in the presence and absence of electron acceptors. No transient absorption which could be identified with an excited state was observed with either anion. The rate of recovery of the ground state in the absence of additives was a first-order process with both acids and gave rate constants for deactivation of the excited state, kD, of 2.4 × 105 s?1 for rhodizonate and 2.8 × 105 s?1 for croconate. With croconate dianion in the presence of three acceptors, 4-nitrobenzylbromide, methylviologen, and biacetyl, a transient absorption was detected, with a maximum absorbance at 500 nm, and was tentatively identified with the monoanion radical, formed following electron transfer to the acceptor. From the rate of growth of the transient, rate constants for the rate of electron transfer to the acceptor were measured as follows: 4-nitrobenzylbromide: 2.8 × 109 M?1 s?1; methyl viologen: 3.7 × 1010 M?1 s?1; and biacetyl: 2.0 × 108 M?1 s?1. The significance of the measurements is discussed in relation to the mechanism proposed for the photochemical reactions of these dianions. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
Laser flash photolysis combined with competition kinetics with SCN? as the reference substance has been used to determine the rate constants of OH radicals with three fluorinated and three chlorinated ethanols in water as a function of temperature. The following Arrhenius expressions have been obtained for the reactions of OH radicals with (1) 2‐fluoroethanol, k1(T) = (5.7 ± 0.8) × 1011 exp((?2047 ± 1202)/T) M?1 s?1, (2) 2,2‐difluoroethanol, k2(T) = (4.5 ± 0.5) × 109 exp((?855 ± 796)/T) M?1 s?1, (3) 2,2,2‐trifluoroethanol, k3(T) = (2.0 ± 0.1) × 1011 exp((?2400 ± 790)/T) M?1 s?1, (4) 2‐chloroethanol, k4(T) = (3.0 ± 0.2) × 1010 exp((?1067 ± 440)/T) M?1 s?1, (5) 2, 2‐dichloroethanol, k5(T) = (2.1 ± 0.2) × 1010 exp((?1179 ± 517)/T) M?1 s?1, and (6) 2,2,2‐trichloroethanol, k6(T) = (1.6 ± 0.1) × 1010 exp((?1237 ± 550)/T) M?1 s?1. All experiments were carried out at temperatures between 288 and 328 K and at pH = 5.5–6.5. This set of compounds has been chosen for a detailed study because of their possible environmental impact as alternatives to chlorofluorocarbon and hydrogen‐containing chlorofluorocarbon compounds in the case of the fluorinated alcohols and due to the demonstrated toxicity when chlorinated alcohols are considered. The observed rate constants and derived activation energies of the reactions are correlated with the corresponding bond dissociation energy (BDE) and ionization potential (IP), where the BDEs and IPs of the chlorinated ethanols have been calculated using quantum mechanical calculations. The errors stated in this study are statistical errors for a confidence interval of 95%. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 174–188, 2008  相似文献   

7.
The pressure-jump method has been used to determine the rate constants for the formation and dissociation of nickel(II) and cobalt(II) complexes with cinchomeronate in aqueous solution at zero ionic strength. The forward and reverse rate constants obtained are kf = 2.27 × 106 M?1 s?1 and kr = 3.81 × 101 s?1 for the nickel(II) complex and kf = 1.23 × 107 M?1 s?1 and kr = 2.66 × 102 s?1 for the cobalt(II) complex at 25°C. The activation parameters of the reactions have also been obtained from the temperature variation study. The results indicate that the rate determining step of the reaction is a loss of a water molecule from the inner coordination sphere of the cation for the nickel(II) complex and the chelate ring closure for the cobalt(II) complex. The influence of the pyridine ring nitrogen atom of the cinchomeronate ligand on the complexation of cobalt(II) ion is also discussed.  相似文献   

8.
Rate constants for a series of alcohols, ethers, and esters toward the sulfate radical (SO4?) have been directly determined using a laser photolysis set‐up in which the radical was produced by the photodissociation of peroxodisulfate anions. The sulfate radical concentration was monitored by following its optical absorption by means of time resolved spectroscopy techniques. At room temperature the following rate constants were derived: methanol ((1.6 ± 0.2) × 107 M?1 s?1); ethanol ((7.8 ± 1.2) × 107 M?1 s?1); tert‐butanol ((8.9 ± 0.3) × 105 M?1 s?1); diethyl ether ((1.8 ± 0.1) × 108 M?1 s?1); MTBE ((3.13 ± 0.02) × 107 M?1 s?1); tetrahydrofuran (THF) ((2.3 ± 0.2) × 108 M?1 s?1); hydrated formaldehyde ((1.4 ± 0.2) × 107 M?1 s?1); hydrated glyoxal ((2.4 ± 0.2) × 107 M?1 s?1); dimethyl malonate (CH3OC(O)CH2C(O)OCH3) ((1.28 ± 0.02) × 106 M?1 s?1); and dimethyl succinate (CH3OC(O)CH2CH2C(O)OCH3) ((1.37 ± 0.08) × 106 M?1 s?1) where the errors represent 2σ. For the two latter species, we also measured the temperature dependence of the corresponding rate constants. A correlation of these kinetics with the bond dissociation energy is also presented and discussed. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 539–547, 2001  相似文献   

9.
For the rate constant of addition of tert-butyl radicals to acrylonitrile at T = 300 K in solution modulated ESR spectroscopy and muon spin rotation yield 106 M?1 s?1 and 2.4 × 106 M?1 s?1. The addition of pivaloyl radical to acrylonitrile proceeds with Arrhenius parameters log A/M?1 s?1 = 7.7 and Ea = 11.5 kJ/ mol. The results are discussed in terms of polar effects in radical addition reactions.  相似文献   

10.
The complex [Fe(imox) (Nmim)2], where imox is a planar bis(iminooxime) macrocyclic ligand and Nmim=N-methyl imidazole, undergoes substitution reactions in the presence of 2,2′bipyridine (bipy), leading to a series of intermediate mixed complexes. The forward and reverse rate constants for the substitution of the Nmim ligand by water are k1 = 0.23 s?1 and k?1=62 M?1s?1, respectively. The binding of a bipy to the [Fe(imox) (Nmim) (H2O)]+ complex proceeds according to k2 = 4.7×10?2 M?1 s?1 and k?2 = 3.6 × 10?4 s?1, yielding [Fe(imox) (bipy) (Nmim)]+, which eliminates Nmim according to k3 = 1.6 × 10?4 s?1. Further substitution in the [Fe(imox) (bipy)]+ complex with bipy takes place very slowly leading to the [Fe(imox) (bipy)2]+ and [Fe(bipy)3]2+ complexes. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
Kinetics for the reactions of OBrO with NO, O3, OClO, and ClO at 240–350 K were investigated using the technique of discharge flow coupled with mass spectrometry. The Arrhenius expression for the OBrO reaction with NO was determined to be k1 = (2.37 ± 0.96) × 10?13 exp[(607 ± 63)/T] cm3 molecule?1 s?1. The reactions of OBrO with O3, OClO, and ClO are slow chemical processes at 240–350 K. Upper limit rate constants for the OBrO reactions with O3, OClO, and ClO at 240–350 K were estimated to be k2 < 5.0 × 10?15 cm3 molecule?1 s?1, k3 < 6.0 × 10?14 cm3 molecule?1 s?1, and k4 < 1.5 × 10?13 cm3 molecule?1 s?1, respectively. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 430–437, 2002  相似文献   

12.
Styrene radical polymerizations mediated by the imidazolidinone nitroxides 2,5‐bis(spirocyclohexyl)‐3‐methylimidazolidin‐4‐one‐1‐oxyl (NO88Me) and 2,5‐bis(spirocyclohexyl)‐3‐benzylimidazolidin‐4‐one‐1‐oxyl (NO88Bn) were investigated. Polymeric alkoxyamine (PS‐NO88Bn)‐initiated systems exhibited controlled/living characteristics at 100–120 °C but not at 80 °C. All systems exhibited rates of polymerization similar to those of thermal polymerization, with the exception of the PS‐NO88Bn system at 80 °C, which polymerized twice as quickly. The dissociation rate constants (kd) for the PS‐NO88Me and PS‐NO88Bn coupling products were determined by electron spin resonance at 50–100 °C. The equilibrium constants were estimated to be 9.01 × 10?11 and 6.47 × 10?11 mol L?1 at 120 °C for NO88Me and NO88Bn, respectively, resulting in the combination rate constants (kc) 2.77 × 106 (NO88Me) and 2.07 × 106 L mol?1 s?1 (NO88Bn). The similar polymerization results and kinetic parameters for NO88Me and NO88Bn indicated the absence of any 3‐N‐transannular effect by the benzyl substituent relative to the methyl substituent. The values of kd and kc were 4–8 and 25–33 times lower, respectively, than the reported values for PS‐TEMPO at 120 °C, indicating that the 2,5‐spirodicyclohexyl rings have a more profound effect on the combination reaction rather than the dissociation reaction. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 327–334, 2003  相似文献   

13.
The reduction of Fe(CN)5L2? (L = pyridine, isonicotinamide, 4,4′‐bipyridine) complexes by ascorbic acid has been subjected to a detailed kinetic study in the range of pH 1–7.5. The rate law of the reaction is interpreted as a rate determining reaction between Fe(III) complexes and the ascorbic acid in the form of H2A(k0), HA?(k1), and A2? (k2), depending on the pH of the solution, followed by a rapid scavenge of the ascorbic acid radicals by Fe(III) complex. With given Ka1 and Ka2, the rate constants are k0 = 1.8, 7.0, and 4.4 M?1 s?1; k1 = 2.4 × 103, 5.8 × 103, and 5.3 × 103 M?1 s?1; k2 = 6.5 × 108, 8.8 × 108, and 7.9 × 108 M?1 s?1 for L = py, isn, and bpy, respectively, at μ = 0.10 M HClO4/LiClO4, T = 25°C. The kinetic results are compatible with the Marcus prediction. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 126–133, 2005  相似文献   

14.
The kinetics of ethylene/propylene copolymerization catalyzed by (ethylene bis (indeyl)-ZrCI2/methylaluminoxane) has been investigated. Radiolabeling found about 80% of the Zr to be catalytically active. The estimates for rate constants at 50°C are k11 = 1104 (Ms)?1, k12 = 430 (Ms)?1, k22 = 396 (Ms)?1,k21 = 1020 (Ms)?1, and kAtr,1 + kAtr.2 = 1.9 × 10?3 s?1. Substitution of trimethylaluminum for methylaluminoxane resulted in proportionate decrease in polymerization rate. The molecular weight of the copolymer is slightly increased by loweing the [Al]/[Zr] ratio, or addition of Lewis base modifier but at the expense of lowered catalytic activity and increase in ethylene content in the copolymer. Lowering of the polymerization temperature to 0°C resulted in a doubling of molecular weight but suffered 10-fold reduction in polymerization activity and increase of ethylene in copolymer.  相似文献   

15.
The photochemical decomposition of peroxomonosulfate (PMS) in the presence and absence of 2-propanol at 25°C was found to obey an overall first-order rate – d[PMS]/dt = k?[PMS]. In the absence of 2-propanol, the quantum yield ≤ for the decomposition of PMS was found to depend upon the concentration of PMS at [PMS] > 2 × 10?M, and is independent of concentration at [PMS] > 2 × 10?2M. The quantum yield in the presence of 2-propanol was found to be 3.03 at [PMS] = 1 × 10?2M and 4.45 at higher concentrations of PMS. In the pH range of 2–9.0 the quantum yield was found to be independent of pH, and the overall rate constant k? was found to be 6.49 × 10?3 s?1 and 1.68 × 10?3 s?1, respectively, in the presence and absence of isopropanol. A suitable chain mechanism is proposed and explained.  相似文献   

16.
An analysis of the former works devoted to the reactions of I(III) in acidic nonbuffered solutions gives new thermodynamic and kinetic information. At low iodide concentrations, the rate law of the reaction IO + I? + 2H+ ? IO2H + IOH is k+B [IO][I?][H+]2k?B [IO2H][IOH] with k+B = 4.5 × 103 M?3s?1 and k?B = 240 M?1s?1 at 25°C and zero ionic strength. The rate law of the reaction IO2H + I? + H+ ? 2IOH is k+C [IO2H][I?][H+] – k?C [IOH]2 with k+C = 1.9 × 1010 M?2s?1 and k?C = 25 M?1s?1. These values lead to a Gibbs free energy of IO2H formation of ?95 kJ mol?1. The pKa of iodous acid should be about 6, leading to a Gibbs free energy of IO formation of about ?61 kJ mol?1. Estimations of the four rate constants at 50°C give, respectively, 1.2 × 104 M?3s?1, 590 M?1s?1, 2 × 109 M?2s?1, and 20 M?1 s?1. Mechanisms of these reactions involving the protonation IO2H + H+ ? IO2H and an explanation of the decrease of the last two rate constants when the temperature increases, are proposed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 647–652, 2008  相似文献   

17.
The kinetics of oxidation of tartaric acid (TAR) by peroxomonosulfate (PMS) in the presence of Cu(II) and Ni(II) ions was studied in the pH range 4.05–5.20 and also in alkaline medium (pH ~12.7). The rate was calculated by measuring the [PMS] at various time intervals. The metal ions concentration range used in the kinetic studies was 2.50 × 10?5 to 1.00 × 10?4 M [Cu(II)], 2.50 × 10?4 to 2.00 × 10?3M [Ni(II)], 0.05 to 0.10 M [TAR], and µ = 0.15 M. The metal(II) tartarates, not TAR/tartarate, are oxidized by PMS. The oxidation of copper(II) tartarate at the acidic pH shows an appreciable induction period, usually 30–60 min, as in classical autocatalysis reaction. The induction period in nickel(II) tartarate is small. Analysis of the [PMS]–time profile shows that the reactions proceed through autocatalysis. In alkaline medium, the Cu(II) tartarate–PMS reaction involves autocatalysis whereas Ni(II) tartarate obeys simple first‐order kinetics with respect to [PMS]. The calculated rate constants for the initial oxidation (k1) and catalyzed oxidation (k2) at [TAR] = 0.05 M, pH 4.05, and 31°C are Cu(II) (1.00 × 10?4 M): k1 = 4.12 × 10?6 s?1, k2 = 7.76 × 10?1 M?1s?1 and Ni(II) (1.00 × 10?3 M): k1 = 5.80 × 10?5 s?1, k2 = 8.11 × 10?2 M?1 s?1. The results suggest that the initial reaction is the oxidative decarboxylation of the tartarate to an aldehyde. The aldehyde intermediate may react with the alpha hydroxyl group of the tartarate to give a hemi acetal, which may be responsible for the autocatalysis. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 620–630, 2011  相似文献   

18.
Ultrasound‐mediated atom transfer radical polymerization (sono‐ATRP) in miniemulsion media is used for the first time for the preparation of complex macromolecular architectures by a facile two‐step synthetic route. Initially, esterification reaction of sucrose or lactulose with α‐bromoisobutyryl bromide (BriBBr) is conducted to receive multifunctional ATRP macroinitiators with 8 initiation sites, followed by polymerization of n‐butyl acrylate (BA) forming arms of the star‐like polymers. The brominated lactulose‐based molecule was examined as an ATRP initiator by determining the activation rate constant (ka) of the catalytic process in the presence of a copper(II) bromide/tris(2‐pyridylmethyl)amine (CuIIBr2/TPMA) catalyst in both organic solvent and for the first time in miniemulsion media, resulting in ka = (1.03 ± 0.01) × 104 M?1 s?1 and ka = (1.16 ± 0.56) × 103 M?1 s?1, respectively. Star‐like macromolecules with a sucrose or lactulose core and poly(n‐butyl acrylate) (PBA) arms were successfully received using different catalyst concentration. Linear kinetics and a well‐defined structure of synthesized polymers reflected by narrow molecular weight distribution (Mw/Mn = 1.46) indicated 105 ppm wt of catalyst loading as concentration to maintain controlled manner of polymerization process. 1H NMR analysis confirms the formation of new sugar‐inspired star‐shaped polymers.  相似文献   

19.
The interaction of cisplatin with guanine DNA bases has been investigated using ab initio Hartree–Fock (HF) and density functional levels of theory in gas phase and aqueous solution. The overall process was divided into three steps: the reaction of the monoaqua [Pt(NH3)2Cl(H2O)]+ species with guanine (G) (reaction 1), the hydrolysis process yielding the adduct [Pt(NH3)2(G) (H2O)]2+ (reaction 2) and the reaction with a second guanine leading to the product [Pt(NH3)2(G)2]2+ (reaction 3). The functionals B3LYP, BHandH, and mPW1PW91 were used in the present study, to develop an understanding of the role of the distinct models. The geometries presented for the intermediate structures were obtained by IRC calculations from the transition state structure for each reaction. The structural analysis for the intermediates and transition states showed that hydrogen bonds with the guanine O6 atom play an important role in stabilizing these species. The geometries were not very sensitive to the level of theory applied with the HF level, giving a satisfactory overall performance. However, the energy barriers and the rate constants were found to be strongly dependent on the level of calculation and basis set, with the DFT calculations giving more accurate results. For reaction 1 the rate constant calculated in aqueous solution at PCM‐BHandH/6‐311G* (k1 = 7.55 × 10?1 M?1 s?1) was in good agreement with the experiment (5.4 × 10?1 M?1 s?1). The BHandH/6‐31G* calculated gas phase rate constants for reactions 2 and 3 were: k2 = 0.9 × 10?6 M?1 s?1 and k3 = 2.99 × 10?4 M?1 s?1 in fairly good accordance with the experimental findings for reaction 2 (1.0 × 10?6 M?1 s?1) and reaction 3 (3.0 × 10?4 M?1 s?1). © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

20.
The kinetics of reactions of OH radical with n‐heptane and n‐hexane over a temperature range of 240–340K has been investigated using the relative rate combined with discharge flow/mass spectrometry (RR/DF/MS) technique. The rate constant for the reaction of OH radical with n‐heptane was measured with both n‐octane and n‐nonane as references. At 298K, these rate constants were determined to be k1, octane = (6.68 ± 0.48) × 10?12 cm3 molecule?1 s?1 and k1, nonane = (6.64 ± 1.36) × 10?12 cm3 molecule?1 s?1, respectively, which are in very good agreement with the literature values. The rate constant for reaction of n‐hexane with the OH radical was determined to be k2 = (4.95 ± 0.40) × 10?12 cm3 molecule?1 s?1 at 298K using n‐heptane as a reference. The Arrhenius expression for these chemical reactions have been determined to be k1, octane = (2.25 ± 0.21) × 10?11 exp[(?293 ± 37)/T] and k2 = (2.43 ± 0.52) × 10?11 exp[(?481.2 ± 60)/T], respectively. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 489–497, 2011  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