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 46 毫秒
1.
The rate law ? d[O3]/ dt = k1[A][O3] + k3[A][O3]2/ (k4 + k5[O2]) has been found to obtain for the reaction of ozone with allene and with 1,2-butadiene. We now find that this rate law also applies to the reaction of ozone with ethylene and presumably with all lower alkenes. This generalizes the inhibiting effect of oxygen and accounts for the simplifed rate law found in the presence of excess oxygen. Oxygen itself is a product of the ozone–ethylene reaction, and we find that as [O3]0 increases, the (O2 formed)/(O3 used) ratio approaches 1.5. Values of k1, k3/k5 for ethylene are compared with those for allene, 1,3-butadiene, and propene. A generalized mechanism is postulated for the reaction of ozone with alkenes involving a chain sequence that produces oxygen and which accounts for the observed rate law. A specific mechanism is postulated for the reaction of O3 with ethylene, and the thermochemistry of the chain sequence is examined in detail.  相似文献   

2.
Summary Catalysis of the CeIV-allyl alcohol (AA) reaction in acid solution depends both on the of rate enhancement and product distribution on the catalyst used: OsVIII results mainly in acrolein, whereas PdII gives acrylic acid. The rate laws in the two cases also differ:viz., Equations 1 and 2K1 is the equilibrium constant of formation of the OsVIII-allyl alcohol complex and k1 is the rate constant of its oxidation by CeIV; K2 is the equilibrium constant for the formation of the CeIV-PdII-allyl alcohol complex and k2 is its rate constant of decomposition. Rate = K1k1[CeIV][AA][OsVIII]/(1+K1[AA]) (1) Rate = K1k1[CeIV][PdII]/(1+K2[CeIV]) (2)While OsVIII is effective in H2SO4 solution, aqueous HClO4 is needed for PdII. Both reactions proceed through formation of catalyst-allyl alcohol complexes with participation of free radicals. The details of these observations are discussed.  相似文献   

3.
The kinetics of oxidation of dimethyl sulfoxide (DMSO) by chloramine-T (CAT) is studied in HClO4 and NaOH media with OsO4 as a catalyst in the latter medium. In acid medium, the rate law is -d [CAT]/dt = k [CAT][DMSO][H+]. Alkali retards the reaction and the rate law takes the form -d [CAT]/dt = k [CAT][DMSO][OsO4]/[NaOH], but is reduced to -d [CAT]/dt = k [CAT][DMSO] at higher alkali concentrations. The reaction is subjected to changes in (a) ionic strength, (b) concentrations of added neutral salts, (c) concentrations of added reaction product, (d) dielectric constant, and (e) solvent isotope effect, and the subsequent effects on the reaction rate are studied. The reaction mechanism in acid medium assumes an electrophilic attack by the free acid RNHCl (CAT′) at the sulfur site in DMSO, forming a reaction intermediate which subsequently decomposes to dimethyl sulfone on hydrolysis. Formation of a cyclic complex between RNHCl and OsO4 which interacts with the substrate in a slow step explains the observed results in alkaline medium. The simplification of the rate equation at higher alkali concentrations is attributed to a direct reaction between chloramine-T and the substrate.  相似文献   

4.
The nucleophilic second-order rate constant (kOH) for the reaction of ōH with ionized N-hydroxyphthalimide (S?) appears to follow a reaction mechanism similar to that for reactions of ōH with neutral phthalimide and its N-substituted derivatives. Kinetically indistinguishable terms, kw[H2O][S?] and kōH[ōH][SH] (SH represents nonionized N-hydroxyphthalimide), which constitute the pH-independent rate region of the pH-rate profile, are resolved qualitatively. It is shown that the term kōH[ōH][SH], rather than kw[H2O][S?], is important in these reactions. The rates of ōH-catalyzed cleavage of S? were studied at 32° in the presence of micelles of sodium dodecyl sulphate (SDS). At a constant [ōH], the observed pseudo first-order rate constants (kobs) increase linearly with [SDS]T (total SDS concentration). These data are explained in terms of the pseudophase model of micellar effects on reactivity. The linear dependence of kobs with [SDS]T (within [SDS]T range of 0.0–0.2 or 0.3 M) is attributed to the occurrence of the reaction between the exterior boundary of Stern layer and Gouy-Chapman layer.  相似文献   

5.
Summary The acid-catalysed decarboxylation of thecis-- andcis--[CoL(CO3)]+ complexes (L = 3,6-dimethyl-1,8-diamino-3,6-diazaoctane) have been studied over a range of HClO4 concentrations and the temperatures 25, 35 and 45° at I = 1.0 mol dm–3 (NaClO4). The rate expression takes the form kobs = k0 + k1[H+] where kobs is the observed first order rate constant at constant hydrogen ion concentration. The k0 term makes only a minor contribution to the overall reaction. Both complexes display solvent deuterium isotope effects ofca. 2.6 for the acid-catalysed decarboxylation, consistent with a rapid proton pre-equilibrium mechanism. Activation parameters have been determined and the mechanism of the reaction discussed. The magnitude of the solvent isotope effect is consistent with an A-1 type mechanism involving formation of a 5-coordinate intermediate.  相似文献   

6.
The kinetics of oxidation of N,N‐dimethylformamide by chromium(VI) has been studied spectrophotometrically in aqueous perchloric acid media at 20°C. The rate showed a first‐order dependence on both [Cr(VI)] and [DMF], and increased markedly with increasing [H+]. The order with respect to [HClO4] was found to lie between 1 and 2. The rate was found to be independent of ionic strength as well as of any inhibition effect of Mn(II). The formation of superoxochromium(III) ion was detected in an aerated solution of chromium(VI), DMF and HClO4. The proposed mechanism, involving two reaction pathways, leads to the rate law, rate = Ka1 [HCrO4] [DMF] (kI Ka2 [H+]²+kII[H+]). The first pathway, with rate constant kI, involves the formation of chromium(V) and a free radical. The second pathway, with rate constant kII, involves the formation of Cr(IV), CO2 and dimethylamine. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 409–415, 1999  相似文献   

7.
The kinetics of oxidation of the aliphatic primary amines, n-propylamine, n-butylamine, and isoamylamine, by N-sodio-N-bromobenznesulfonamide or bromamine-B (BAB), in the presence of osmium(VIII), has been studied in alkaline medium at 35°C. In the presence of the catalyst, the experimental rate law for the oxidation of the amine substrate (S) takes the form, rate=k[BAB][OsO4][OH]x, which in the absence of the catalyst changes to the form, rate=k[BAB][S][OH]y, where x and y are less than unity. Additions of halide ions and the reduction product of BAB (benzenesulfonamide), and the variation of ionic strength of the solvent medium have no effect on the reaction rate. Activation parameters have been evaluated. The proposed mechanism assumes the formation of a complex intermediate between the active oxidant species, PhSO2NBr, and the catalyst, OsO4, in the rate determining step. This complex then interacts with the substrate amine in fast steps to yield the end products. The average value for the deprotonation constant of monobromamine-B, forming PhSO2NBr, is evaluated for the Os(VIII) catalyzed reactions of the three amines in alkaline medium as 9.80×103 at 35°C. The average value for the same constant for the uncatalyzed reactions is 1.02×104 at 35°C. © 1997 John Wiley & Sons, Inc.  相似文献   

8.
A kinetic study of oxidation of metronidazole (Met) with sodium N‐bromo‐p‐toluenesulfonamide or bromamine‐T (BAT) has been carried out in HClO4 (30°C) and NaOH (40°C) media. The experimental rate laws obtained are –d[BAT]/dt=k[BAT][Met]x [H+]y in acid medium and –d[BAT]/dt=k[BAT][Met]x [OH?]y/[PTS]z in alkaline medium, where x, y, and z are less than unity and PTS is p‐toluenesulfonamide. The reaction was subjected to changes in (a) ionic strength, (b) concentration of added reduction product PTS, (c) concentration of added neutral salts, (d) dielectric permittivity, and (e) solvent isotope effect. In both media, the stoichiometry of the reaction was found to be 1:1, and the oxidation product of metronidazole was identified as its aldehyde. The reaction was studied at different temperatures, and the activation parameters have been evaluated. The reaction constants involved in the proposed schemes were deduced. The reaction was found to be faster in acid medium in comparison with alkaline medium, which is attributed to the involvement of different oxidizing species. Mechanisms proposed and the rate laws derived are consistent with the observed kinetics. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 700–709, 2005  相似文献   

9.
The reaction between ozone and thiophene was studied from 30 to 125°C over a pressure range of 0.005-0.3 torr ozone and 0.1-1 torr thiophene. The most abundant product was O2 with smaller amounts of CO2 and SO2. The mass balance was 100% for oxygen and approached 100% for sulfur at the higher values of [O3]0. The carbon balance, however, was only 25% and no H-containing products were found, suggesting that the missing product is a hydrocarbon which may be a polymer. The rate law found was -d[O3]/dt = k1[Th] [O3] + k2 [Th] [O3]2 where log k1(M?1 · sec?1) = 7.8 ± 0.5 - (8400 ± 700)/2.3RT, and log k2(M?2 · sec?1) = 12.4 ± 0.4 - (4700 ± 400)/2.3RT. Added O2 had no effect on k1 but reduced k2 to a limiting value. It is thus not possible to measure the primary rate constant in this system by measuring the overall rate in the presence of oxygen, and this restriction may also apply to other ozone systems. A mechanism is postulated involving two chain sequences, one of which is inhibited by added O2. A comparison with other ozone systems is made, and the chain lengths are far greater for ozone + thiophene than other systems, under the conditions employed. Possible intermediates in the mechanism are discussed.  相似文献   

10.
The kinetics of the reaction between ozone and allene (A) were studied in the range of 226 to 325°K in the gas phase. Initial O3 pressures varied from 0.01 to 0.7 torr and allene pressures varied from 0.05 to 6 torr. At the higher initial O3 pressures the most important product was O2 followed by CO, H2O, CO2, and C2H4. Oxygen balances averaging about 110% were obtained, which implies that no important oxygenated products were missed. However, carbon balances were only about 50% and hydrogen balances were even less, so that unidentified hydrocarbons were presumably formed. The rate law found was ? d[O3]/dt = k1[O3][A] + k2a[O3]2[A]/[O3]0 where log k1(M?1sec?1) = 6.0 ± 0.7 ? (5500±1000)/2.30RT and log k2a(M?1sec?1) = 6.9 ± 0.7 ? (6200 ± 800/2.30RT). A mechanism is proposed which accounts for the rate law and the observed stoichiometry of O2 formed–O3 used. This involves a heterogeneous catalyzed decomposition of O3. The rate constant k1 is identified with the primary addition reaction A + O3 → AO3, and this rate constant is compared with those from other O3 addition reactions.  相似文献   

11.
The kinetics of the reaction between 1,2,3-trihydroxybenzene (pyrogallol) and O2 (autoxidation) have been determined by monitoring the concentration of dissolved dioxygen with a polarographic oxygen electrode. The reaction is carried out in pseudo-first-order excess pyrogallol, 25°C, 0.08 M NaCl, and 0.04 M phosphate buffer in the pH range 6.9–10.5. Data collection precedes reaction initiation, but only the data recorded after the estimated 3.2 s dead time are used in kinetics calculations. Observed rate constants are corrected for incomplete mixing, which is treated as a first-order process that has an experimentally determined mixing rate constant of 4.0 s?1. The rate law for the reaction is ?d[O2]/dt=kapp[PYR]tot[O2], in which [PYR]tot is the total stoichiometric pyrogallol concentration. A mechanism is presented which explains the increase in rate with increasing [OH?] by postulating that H2PYR? (k2) has greater reactivity with dissolved dioxygen than does H3PYR (k1). The data best fit the equation kapp=(k1 + k2KH[OH?])/(1 + KH[OH?]) when the value of the hydrolysis constant KH (the quotient of the pyrogallol acid dissociation and water autoprotolysis constants) for this medium equals 3.1×104 M?1. The resulting values of k1 and k2, respectively, equal (0.13 + 0.01) M?1 s?1 and (3.5 plusmn; 0.1) M?1 s?1. This reaction is recommended as a test reaction for calibrating the dynamic response of an O2-electrode. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Chromium(III)-isonicotinate complexes, cis-[Cr(C2O4)2(N-inic)(H2O)]- and [Cr(C2O4)(H2O)3-OH-Cr(C2O4)2(O-inic)]-(N-inic)(H2 (N-inic = N-bonded and O-inic = O-bonded isonicotinic acid) were obtained and characterized in solution. Kinetics of acid-catalyzed isonicotinate ligand liberation were studied spectrophotometrically in the 0.1–1.0 m HClO4 range, at I=1.0 m. The dependencies of the pseudo-first order rate constant on [H+] were established: kobs = k0+kHQH[H+] and kobs = kHQH[H+] for the N-inic and O-inic complex, respectively, where k0 and kH are the rate constants of the spontaneous and the acid-catalyzed reaction paths, and QH is the protonation constant of the carboxylic group in isonicotinic ligand. The obtained results indicate that N-bonded isonicotinic acid liberation occurs mainly via a spontaneous reaction path and is much slower than O-bonded inic liberation. The mechanisms for these processes are proposed.  相似文献   

13.
Summary The oxidation of MeCHO by chromium(VI) has been studied in HClO4 medium over a wide range of experimental conditions and has been found to obey the rate law;v=k[MeCHO][HCrO 4 ][H+]. The calculated H and-S values for the reaction are 30±2kJ mol–1 and 171±7J mol–1deg–1, respectively. The mechanism is discussed in terms of carbon-hydrogen bond cleavage.  相似文献   

14.
Summary The kinetics of oxidation of 1,4-butanediol by chromium(VI) was studied in acid perchlorate medium and the oxidation product of the diol was identified as 4-hydroxybutanal. The kinetic rate law observed accounted for the complex dependence of the hydrogen ion k=(k2K1[H+]+k3K1K2[H+]2)/(1+K1[H]+) where k is the observed second-order rate constant.  相似文献   

15.
In the stirred batch experiment, the Mn(II)-catalyzed bromate-saccharide reaction in aqueous H2SO4 or HClO4 solution exhibits damped oscillations in the concentrations of bromide and Mn(II) ions. Peculiar multiple oscillations are observed in the system with arabinose or ribose. The apparent second-order rate constants of the Mn(III)-saccharide reactions at 25°C are (0.659, 1.03, 1.76, 2.32, and 6.95) M−1 s−1 in 1.00 M H2SO4 and (4.69, 7.51, 10.2, 13.5, and 36.2) M−1 s−1 in (2.00–4.00) M HClO4 for (glucose, galactose, xylose, arabinose, and ribose), respectively. At 25°C, the observed pseudo-first-order rate constant of the Mn(III)-Br reaction is kobs = (0.2 ± 0.1) [Br] + (130 ± 5)[Br]2 + (2.6 ± 0.1) × 103[Br]3 + (1.2 ± 0.2) × 104[Br]4 s−1 and the rate constant of the Br2 Mn(II) reaction is less than 1 × 10−4 M−1 s−1. The second-order rate constants of the Br2-saccharide reactions are (3.65 ± 0.15, 11.0 ± 0.5, 4.05, 12.5 ± 0.7, and 2.62) × 10−4 M−1 s−1 at 25°C for glucose, galactose, xylose, arabinose, and ribose, respectively.  相似文献   

16.
A simple computerized method using APL is described which gives the rate constants for systems obeying the rate law –d[A]/dt = k1[A][B] + k2[A] m [B]l from sets of concentrations and times. Applications are discussed.  相似文献   

17.
Chromium(III)-lutidinato complexes of general formula [Cr(lutH) n (H2O)6−2n ]3−n (where lutH is N,O-bonded lutidinic acid anion) were obtained and characterized in solution. Acid-catalysed aquation of [Cr(lutH)3]0 leads to only one ligand dissociation, whereas base hydrolysis produces chromates(III) as a result of subsequent ligand liberation steps. The kinetics of the first ligand dissociation were studied spectrophotometrically, within the 0.1–1.0 M HClO4 and 0.4–1.0 M NaOH range. In acidic media, two reaction stages, the chelate-ring opening and the ligand dissociation, were characterized. The dependencies of pseudo-first-order rate constants on [H+] are as follows: k obs1 = k 1 + k −1/K 1[H+] and k obs2 = k 2 K 2[H+]/(1 + K 2[H+]), where k 1 and k 2 are the rate constants for the chelate-ring opening and the ligand dissociation, respectively, k −1 is the rate constant for the chelate-ring closure, and K 1 and K 2 are the protonation constants of the pyridine nitrogen atom and coordinated 2-carboxylate group in the one-end bonded intermediate, respectively. In alkaline media, the rate constant for the first ligand dissociation depends on [OH]: k obs1 = k OH(1) + k O[OH], where k OH(1) and k O are rate constants of the first ligand liberation from the hydroxo- and oxo-forms of the intermediate, respectively, and K 2 is an equilibrium constant between these two protolytic forms. Kinetic parameters were determined and a mechanism for the first ligand dissociation is proposed. The kinetics of the ligand liberation from [Cr(lut)(OH)4]3− were also studied and the values of the pseudo-first-order rate constants are [OH] independent.  相似文献   

18.
The rate and molecular weight profiles were obtained for the spontaneous alternating copolymerizations conducted with diethylaluminum chloride. The rate formally fitted an expression, R = kp[MMA][Sty], and the rate constant was established by two distinct methods: (1) from the yield versus time data and (2) from initial rate over a range of initial concentrations; it was determined as 5.4 × 10?6 l./mole-sec with Ea = 4.2 kcal/mole. Molecular weights were determined by gel-permeation chromatography. No increase in molecular weight was observed with increased reaction time. Thus living centers or diradicals are not involved in the process. The M?n shows a steady decrease with increase in monomer-diethylaluminum chloride concentration but the rate is maximum at equimolar monomer concentrations. The data are interpreted on a chain-transfer mechanism and show close agreement to a model in which the excess complexed acceptor monomer is the transfer agent. The chain-transfer constant of 7.1 × 10?4 l./mole-sec is several orders of magnitude greater than for uncomplexed systems.  相似文献   

19.
Oxidation of 3-pyridinemethanol (3-pyol), 4-pyridinemethanol (4-pyol), 3-pyridinecarboxaldehyde (3-pyal) and 4-pyridinecarboxaldehyde (4-pyal) by CrVI was studied under pseudo-first-order conditions in the presence of a large excess of reductant and at various Haq + concentrations; [CrVI] = 8 × 10–4 M, [reductant] = 0.025–0.20 M, [HClO4] = 1.0 and 2.0 M (I = 1.2 and 2.1 M) or 0.5–2.0 (I = 2.1 M). A linear dependence of the pseudo-first-order rate constant (k obs) on [reductant] and a parabolic function of k obs versus [H+] lead to the rate law: –d[CrVI]/dt = (a + b[H+]2)[reductant][CrVI], where a and b describe the reaction paths via HCrO4 and H3CrO4 + species respectively, and are composite values including rate constants and equilibrium constants. The apparent activation parameters were determined from second-order rate constants at 1.0 and 2.0 M HClO4, at three temperatures within the 293–323 K range. The presence of chromium species with intermediate oxidation states – CrV, CrIV and CrII, was deduced based on e.s.r. measurements and the kinetic effects of MnII or O2 (Ar), respectively. The alcohols were oxidized to the aldehydes, and carboxylic acids and the aldehydes to the carboxylic acids. Chromium(III) was in the form of the [Cr(H2O)6]3+ complex.  相似文献   

20.
The chromium(III) complexes with a new potential chromium transporting ligand—2,5-pyridinedicarboxylic acid (isocinchomeronic acid, icaH2):[Cr(icaH)3]0, [Cr(icaH)2 (H2O)2]+ and [Cr(icaH)(H2O)4]2+ (where icaH = N,O-bonded isocinchomeronic acid anion), have been obtained and characterized in solution. The [Cr(icaH)3]0 complex undergoes aquation in acidic media to the diaqua-product. Kinetics of this process was studied spectrophotometrically in the 0.1–1.0 M HClO4 range, at I = 1.0 M. The first aquation stage, the chelate-ring opening at the Cr–N bond, is a much faster than the second one. The rate laws are of the form: k obs = k 1 + k −1/Q 1[H+] and k obs = k 2 Q 2[H+]/(1 + Q 2[H+]), where k 1 and k 2 are the rate constants for the chelate-ring opening and the ligand liberation, respectively, k −1 is the rate constant of the chelate-ring closure, Q 1 and Q 2 are the protonation constants of the pyridine nitrogen and 5-carboxylate group in the one-end bonded intermediate, respectively. The results are discussed in terms of potential pharmaceutical application of the complex.  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