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 31 毫秒
1.
Observed pseudo-first-order rate constants (k(obs)) of the hydride-transfer reactions from isopropyl alcohol (i-PrOH) to two NAD(+) analogues, 9-phenylxanthylium ion (PhXn(+)) and 10-methylacridinium ion (MA(+)), were determined at temperatures ranging from 49 to 82 degrees C in i-PrOH containing various amounts of AN or water. Formations of the alcohol-cation ether adducts (ROPr-i) were observed as side equilibria. The equilibrium constants for the conversion of PhXn(+) to PhXnOPr-i in i-PrOH/AN (v/v = 1) were determined, and the equilibrium isotope effect (EIE = K(i-PrOH)/K(i-PrOD)) at 62 degrees C was calculated to be 2.67. The k(H) of the hydride-transfer step for both reactions were calculated on the basis of the k(obs) and K. The corresponding deuterium kinetic isotope effects (e.g., KIE(OD)(H) = k(H)(i-PrOH)/k(H)(i-PrOD) and KIE(beta-D6)(H) = k(obs)(i-PrOH)/k(obs)((CD3)2CHOH)), as well as the activation parameters, were derived. For the reaction of PhXn(+) (62 degrees C) and MA(+) (67 degrees C), primary KIE(alpha-D)(H) (4.4 and 2.1, respectively) as well as secondary KIE(OD)(H) (1.07 and 1.18) and KIE(beta-D6)(H) (1.1 and 1.5) were observed. The observed EIE and KIE(OD)(H) were explained in terms of the fractionation factors for deuterium between OH and OH(+)(OH(delta+)) sites. The observed inverse kinetic solvent isotope effect for the reaction of PhXn(+) (k(obs)(i-PrOH)/k(obs)(i-PrOD) = 0.39) is consistent with the intermolecular hydride-transfer mechanism. The dramatic reduction of the reaction rate for MA(+), when the water or i-PrOH cosolvent was replaced by AN, suggests that the hydride-transfer T.S. is stabilized by H-bonding between O of the solvent OH and the substrate alcohol OH(delta+). This result suggests an H-bonding stabilization effect on the T.S. of the alcohol dehydrogenase reactions.  相似文献   

2.
A panel of seven isotopically substituted sialoside natural substrate analogues based on the core structure 7-(5-acetamido-3,5-dideoxy-d-glycero-α-d-galacto-non-2-ulopyranosylonic acid)-(2→6)-β-D-galactopyranosyloxy)-8-fluoro-4-methylcoumarin (1, Neu5Acα2,6GalβFMU) have been synthesized and used to probe the rate-limiting step for turnover by the M. viridifaciens sialidase. The derived kinetic isotope effects (KIEs) on k(cat) for the ring oxygen ((18)V), leaving group oxygen ((18)V), anomeric carbon ((13)V), C3-carbon ((13)V), C3-R deuterium ((D)V(R)), C3-S deuterium ((D)V(S)), and C3-dideuterium ((D)(2)V) are 0.986 ± 0.003, 1.003 ± 0.005, 1.021 ± 0.006, 1.001 ± 0.008, 1.029 ± 0.007, 0.891 ± 0.008, and 0.890 ± 0.006, respectively. The solvent deuterium KIE ((D(2)O)V) for the sialidase-catalyzed hydrolysis of 1 is 1.585 ± 0.004. In addition, a linear proton inventory was measured for the rate of hydrolysis, under saturating condition, as a function of n, the fraction of deuterium in the solvent. These KIEs are compatible with rate-determining cleavage of the enzymatic tyrosinyl β-sialoside intermediate. Moreover, the secondary deuterium KIEs are consistent with the accumulating Michaelis complex in which the sialosyl ring of the carbohydrate substrate is in a (6)S(2) skew boat conformation. These KIE measurements are also consistent with the rate-determining deglycosylation reaction occurring via an exploded transition state in which synchronous charge delocalization is occurring onto the ring oxygen atom. Finally, the proton inventory and the magnitude of the solvent KIE are consistent with deglycosylation involving general acid-catalyzed protonation of the departing tyrosine residue rather than general base-assisted attack of the nucleophilic water.  相似文献   

3.
Choline oxidase catalyzes the four-electron oxidation of choline to glycine betaine, with betaine aldehyde as an intermediate. In this study, primary deuterium and solvent kinetic isotope effects have been used to elucidate the mechanism for substrate oxidation by choline oxidase using both steady-state kinetics and rapid kinetics techniques. The D(kcat/Km) value with 1,2-[2H4]-choline at saturating oxygen concentration was independent of pH in the range between 6.5 and 10, with a value of approximately 10.6, indicating that CH bond cleavage is not masked by other titratable kinetic steps belonging to the reductive half-reaction. In agreement with this conclusion, a Dkred value of approximately 8.9 was determined at pH 10 for the anaerobic reduction of the flavin by choline, irrespective of whether aqueous or deuterated solvent was used. At pH 10, both the D2(O)(kcat/Km) and the D2(O)kred values were not different from unity with choline or 1,2-[2H4]-choline, while the Dkcat and D2(O)kcat values were 7.3 and 1.1, respectively. The kcat and kred values were 133 s(-1) and 135 s(-1) with betaine aldehyde and 60 s(-1) and 93 s(-1) with choline. These data are consistent with a chemical mechanism in which the choline hydroxyl proton is not in flight in the transition state for CH bond cleavage and with chemical steps of flavin reduction by choline and betaine aldehyde being rate limiting for the overall turnover of the enzyme.  相似文献   

4.
Methanol solutions containing Cd(II), Mn(II), and a palladacycle, (dimethanol bis(N,N-dimethylbenzylamine-2C,N)palladium(II) (3), are shown to promote the methanolytic transesterification of O-methyl O-4-nitrophenyl phosphorothioate (2b) at 25 °C with impressive rate accelerations of 10(6)-10(11) over the background methoxide promoted reaction. A detailed mechanistic investigation of the methanolytic cleavage of 2a-d having various leaving group aryl substitutions, and particularly the 4-nitrophenyl derivative (2b), catalyzed by Pd-complex 3 is presented. Plots of k(obs) versus palladacycle [3] demonstrate strong saturation binding to form 2b:3. Numerical fits of the kinetic data to a universal binding equation provide binding constants, K(b), and first order catalytic rate constants for the methanolysis reaction of the 2b:3 complex (k(cat)) which, when corrected for buffer effects, give corrected (k(cat)(corr)) rate constants. A sigmoidal shaped plot of log(k(cat)(corr)) versus (s)(s)pH (in methanol) for the cleavage of 2b displays a broad (s)(s)pH independent region from 5.6 ≤ (s)(s)pH ≤ 10 with a k(minimum) = (1.45 ± 0.24) × 10(-2) s(-1) and a [lyoxide] dependent wing plateauing above a kinetically determined (s)(s)pK(a) of 12.71 ± 0.17 to give a k(maximum) = 7.1 ± 1.7 s(-1). Br?nsted plots were constructed for reaction of 2a-d at (s)(s)pH 8.7 and 14.1, corresponding to reaction in the midpoints of the low and high (s)(s)pH plateaus. The Br?nsted coefficients (β(LG)) are computed as -0.01 ± 0.03 and -0.86 ± 0.004 at low and high (s)(s)pH, respectively. In the low (s)(s)pH plateau, and under conditions of saturating 3, a solvent deuterium kinetic isotope effect of k(H)/k(D) = 1.17 ± 0.08 is observed; activation parameters (ΔH(Pd)(++) = 14.0 ± 0.6 kcal/mol and ΔS(Pd)(++)= -20 ± 2 cal/mol·K) were obtained for the 3-catalyzed cleavage reaction of 2b. Possible mechanisms are discussed for the reactions catalyzed by 3 at low and high sspH. This catalytic system is shown to promote the methanolytic cleavage of O,O-dimethyl phosphorothioate in CD3OD, producing (CD3O)2P═O(S(-)) with a half time for reaction of 34 min.  相似文献   

5.
We test the hypothesized pathway by which protons are passed from the substrate, ascorbate, to the ferryl oxygen in the heme enzyme ascorbate peroxidase (APX). The role of amino acid side chains and bound solvent is demonstrated. We investigated solvent kinetic isotope effects (SKIE) for the wild-type enzyme and several site-directed replacements of the key residues which form the proposed proton path. Kinetic constants for H(2)O(2)-dependent enzyme oxidation to Compound I, k(1), and subsequent reduction of Compound II, k(3), were determined in steady-state assays by variation of both H(2)O(2) and ascorbate concentrations. A high value of the SKIE for wild type APX ((D)k(3) = 4.9) as well as a clear nonlinear dependence on the deuterium composition of the solvent in proton inventory experiments suggest the simultaneous participation of several protons in the transition state for proton transfer. The full SKIE and the proton inventory data were modeled by applying Gross-Butler-Swain-Kresge theory to a proton path inferred from the known structure of APX. The model has been tested by constructing and determining the X-ray structures of the R38K and R38A variants and accounts for their observed SKIEs. This work confirms APX uses two arginine residues in the proton path. Thus, Arg38 and Arg172 have dual roles, both in the formation of the ferryl species and binding of ascorbate respectively and to facilitate proton transfer between the two.  相似文献   

6.
The E. coli isopentenyl diphosphate isomerase (IDI) catalyzed reaction of isopentenyl diphosphate (IPP) in D(2)O gives a 66% yield of dimethylallyl diphosphate labeled with deuterium at the (E)-methyl group (d-DMAPP) and a 34% yield of IPP labeled with 1 mol of deuterium at C-2 (d-IPP). This shows that the release to D(2)O of the initial product of the IDI-catalyzed reaction (d-DMAPP) is slower than its conversion to d-IPP. Product dissociation is therefore rate determining for isomerization of IPP with a rate constant k(dis) ≈ k(cat) = 0.08 s(-1). The data provide an estimated rate constant of k(as) = 6 × 10(3) M(-1) s(-1) for binding of DMAPP to E. coli IDI that is similar to rate constants determined for the binding of N-protonated 2-amino ethyl diphosphate intermediate analogs to IDI from yeast [Reardon, J. E.; Abeles, R. H. Biochemistry1986, 25, 5609-5616]. We propose that ligand binding to IDI is relatively slow because there is a significant kinetic barrier to reorganization of the initial encounter complex between enzyme, substrate, and an essential Mg(2+) to form the Michaelis complex where the metal cation bridges the protein and the substrate diphosphate group.  相似文献   

7.
The cis,cis-[(bpy)(2)Ru(III)(OH(2))](2)O(4+) micro-oxo dimeric coordination complex is an efficient catalyst for water oxidation by strong oxidants that proceeds via intermediary formation of cis,cis-[(bpy)(2)Ru(V)(O)](2)O(4+) (hereafter, [5,5]). Repetitive mass spectrometric measurement of the isotopic distribution of O(2) formed in reactions catalyzed by (18)O-labeled catalyst established the existence of two reaction pathways characterized by products containing either one atom each from a ruthenyl O and solvent H(2)O or both O atoms from solvent molecules. The apparent activation parameters for micro-oxo ion-catalyzed water oxidation by Ce(4+) and for [5,5] decay were nearly identical, with DeltaH(++) = 7.6 (+/-1.2) kcal/mol, DeltaS() = -43 (+/-4) cal/deg mol (23 degrees C) and DeltaH(++) = 7.9 (+/-1.1) kcal/mol, DeltaS(++) = -44 (+/-4) cal/deg mol, respectively, in 0.5 M CF(3)SO(3)H. An apparent solvent deuterium kinetic isotope effect (KIE) of 1.7 was measured for O(2) evolution at 23 degrees C; the corresponding KIE for [5,5] decay was 1.6. The (32)O(2)/(34)O(2) isotope distribution was also insensitive to solvent deuteration. On the basis of these results and previously established chemical properties of this class of compounds, mechanisms are proposed that feature as critical reaction steps H(2)O addition to the complex to form covalent hydrates. For the first pathway, the elements of H(2)O are added as OH and H to the adjacent terminal ruthenyl O atoms, and for the second pathway, OH is added to a bipyridine ring and H is added to one of the ruthenyl O atoms.  相似文献   

8.
The generation of the active species for the enzyme cytochrome P450 by using the highly versatile oxygen surrogate iodosylbenzene (PhIO) often produces different results compared with the native route, in which the active species is generated through O(2) uptake and reduction by NADPH. One of these differences that is addressed here is the deuterium kinetic isotope effect (KIE) jump observed during N-dealkylation of N,N-dimethylaniline (DMA) by P450, when the reaction conditions change from the native to the PhIO route. The paper presents a theoretical analysis targeted to elucidate the mechanism of the reaction of PhIO with heme, to form the high-valent iron-oxo species Compound I (Cpd I), and define the origins of the KIE jump in the reaction of Cpd I with DMA. It is concluded that the likely origin of the KIE jump is the spin-selective chemistry of the enzyme cytochrome P450 under different preparation procedures. In the native route, the reaction proceeds via the doublet spin state of Cpd I and leads to a low KIE value. PhIO, however, diverts the reaction to the quartet spin state of Cpd I, which leads to the observed high KIE values. The KIE jump is reproduced here experimentally for the dealkylation of N,N-dimethyl-4-(methylthio)aniline, by using intra-molecular KIE measurements that avoid kinetic complexities. The effect of PhIO is compared with N,N-dimethylaniline-N-oxide (DMAO), which acts both as the oxygen donor and the substrate and leads to the same KIE values as the native route.  相似文献   

9.
Using a refined pulsed laser photolysis/pulsed laser-induced fluorescence (PLP/PLIF) technique, the kinetics of the reaction of a surrogate three-ring polynuclear aromatic hydrocarbons (PAH), anthracene (and its deuterated form), with hydroxyl (OH) radicals was investigated over the temperature range of 373 to 1200 K. This study represents the first examination of the OH kinetics for this class of reactions at elevated temperatures (>470 K). The results indicate a complex temperature dependence similar to that observed for simpler aromatic compounds, e.g., benzene. At low temperatures (373-498 K), the rate measurements exhibited Arrhenius behavior (k = 1.82 x 10(-11) exp(542.35/T) in units of cm3 molecule(-1) s(-1)), and the kinetic isotope effect (KIE) measurements were consistent with an OH-addition mechanism. The low-temperature results are extrapolated to atmospheric temperatures and compared with previous measurements. Rate measurements between 673 and 923 K exhibited a sharp decrease in the magnitude of the rate coefficients (a factor of 9). KIE measurements under these conditions were still consistent with an OH-addition mechanism. The following modified Arrhenius equation is the best fit to our anthracene measurements between 373 and 923 K (in units of cm3 molecule(-1) s(-1)): k(1) (373-923 K) = 8.17 x 10(14) T(-8.3) exp(-3171.71/T). For a limited temperature range between 1000 and 1200 K, the rate measurements exhibited an apparent positive temperature dependence with the following Arrhenius equation, the best fit to the data (in units of cm3 molecule(-1) s(-1)): k1 (999-1200 K) = 2.18 x 10(-11) exp(-1734.11/T). KIE measurements above 999 K were slightly larger than unity but inclusive regarding the mechanism of the reaction. Theoretical calculations of the KIE indicate the mechanism of reaction at these elevated temperatures is dominated by OH addition with H abstraction being a minor contributor.  相似文献   

10.
A facile and convenient synthesis of the chiral phthalide framework catalyzed by cationic iridium was developed. The method utilized cationic iridium/bisphosphine‐catalyzed asymmetric intramolecular carbonyl hydroacylation of 2‐keto benzaldehydes to furnish the corresponding optically active phthalide products in good to excellent enantioselectivities (up to 98% ee). The mechanistic studies using a deuterium‐labelled substrate suggested that the reaction involved an intramolecular carbonyl insertion mechanism to iridium hydride intermediate. In addition, we investigated the kinetic isotope effect (KIE) of intramolecular hydroacylation with deuterated substrate and determined that the C?H activation step is not included in the turnover‐limiting step.  相似文献   

11.
Hydroxylation of mesitylene by a nonheme manganese(IV)–oxo complex, [(N4Py)MnIV(O)]2+ ( 1 ), proceeds via one‐step hydrogen‐atom transfer (HAT) with a large deuterium kinetic isotope effect (KIE) of 3.2(3) at 293 K. In contrast, the same reaction with a triflic acid‐bound manganese(IV)‐oxo complex, [(N4Py)MnIV(O)]2+‐(HOTf)2 ( 2 ), proceeds via electron transfer (ET) with no KIE at 293 K. Interestingly, when the reaction temperature is lowered to less than 263 K in the reaction of 2 , however, the mechanism changes again from ET to HAT with a large KIE of 2.9(3). Such a switchover of the reaction mechanism from ET to HAT is shown to occur by changing only temperature in the boundary region between ET and HAT pathways when the driving force of ET from toluene derivatives to 2 is around ?0.5 eV. The present results provide a valuable and general guide to predict a switchover of the reaction mechanism from ET to the others, including HAT.  相似文献   

12.
Detailed mechanistic studies on the ligand hydroxylation reaction mediated by a copper bis(imine) complex are presented. Starting from a structural analysis of the CuI complex and the CuII product with a hydroxylated ligand, the optical absorption and vibrational spectra of starting material and product are analyzed. Kinetic analysis of the ligand hydroxylation reaction shows that O2 binding is the rate-limiting step. The reaction proceeds much faster in methanol than in acetonitrile. Moreover, an inverse kinetic isotope effect (KIE) is evidenced for the reaction in acetonitrile, which is attributed to a sterically congested transition state leading to the peroxo adduct. In methanol, however, no KIE is observed. A DFT analysis of the oxygenation reaction mediated by the micro-eta2:eta2 peroxo core demonstrates that the major barrier after O2 binding corresponds to electrophilic attack on the arene ring. The relevant orbital interaction occurs between the sigma* orbital of the Cu2O2 unit and the HOMO of the ligand. On the basis of the activation energy for the rate-limiting step (18.3 kcal mol(-1)) this reaction is thermally allowed, in agreement with the experimental observation. The calculations also predict the presence of a stable dienone intermediate which, however, escaped experimental detection so far. Reasons for these findings are considered. The implications of the results for the mechanism of tyrosinase are discussed.  相似文献   

13.
The reaction of molecular oxygen with a Pd(II)-hydride species to form a Pd(II)-hydroperoxide represents one of the proposed catalyst reoxidation pathways in Pd-catalyzed aerobic oxidation reactions, but well-defined examples of this reaction were discovered only recently. Here, we present a mechanistic study of the reaction of O2 with trans-(IMes) 2Pd(H)(OBz), 1 (IMes = 1,3-dimesitylimidazol-2-ylidene), which yields trans-(IMes) 2Pd(OOH)(OBz), 2. The reaction was monitored by (1)H NMR spectroscopy in benzene-d6, and kinetic studies reveal a two-term rate law, rate = k1[1] + k2[1][BzOH], and a small deuterium kinetic isotope effect, k(Pd-H)/k(Pd-D) = 1.3(1). The rate is independent of the oxygen pressure. The data support a stepwise mechanism for the conversion of 1 into 2 consisting of rate-limiting reductive elimination of BzOH from 1 followed by rapid reaction of molecular oxygen with (IMes) 2Pd(0) and protonolysis of a Pd-O bond of the eta(2)-peroxo complex (IMes) 2Pd(O2). Benzoic acid and other protic additives (H2O, ArOH) catalyze the oxygenation reaction, probably by stabilizing the transition state for reductive elimination of BzOH from 1. This study provides the first experimental validation of the mechanism traditionally proposed for aerobic oxidation of Pd-hydride species.  相似文献   

14.
Deuterium kinetic isotope effects (KIEs) are reported for the first time for the dissociation of a protein-ligand complex in the gas phase. Temperature-dependent rate constants were measured for the loss of neutral ligand from the deprotonated ions of the 1:1 complex of bovine β-lactoglobulin (Lg) and palmitic acid (PA), (Lg + PA)(n-) → Lg(n-) + PA, at the 6- and 7- charge states. At 25 °C, partial or complete deuteration of the acyl chain of PA results in a measurable inverse KIE for both charge states. The magnitude of the KIEs is temperature dependent, and Arrhenius analysis of the rate constants reveals that deuteration of PA results in a decrease in activation energy. In contrast, there is no measurable deuterium KIE for the dissociation of the (Lg + PA) complex in aqueous solution at pH 8. Deuterium KIEs were calculated using conventional transition-state theory with an assumption of a late dissociative transition state (TS), in which the ligand is free of the binding pocket. The vibrational frequencies of deuterated and non-deuterated PA in the gas phase and in various solvents (n-hexane, 1-chlorohexane, acetone, and water) were established computationally. The KIEs calculated from the corresponding differences in zero-point energies account qualitatively for the observation of an inverse KIE but do not account for the magnitude of the KIEs nor their temperature dependence. It is proposed that the dissociation of the (Lg + PA) complex in aqueous solution also proceeds through a late TS in which the acyl chain is extensively hydrated such that there is no significant differential change in the vibrational frequencies along the reaction coordinate and, consequently, no significant KIE.  相似文献   

15.
The intermolecular primary deuterium isotope effects on the individual C-H bond cleavage steps involved in dihydroceramide Delta(4) desaturation have been determined for the first time by incubating rat liver microsomes with 1:1 mixtures of nonlabeled substrate and the appropriate regiospecifically dideuterated analogue. Analysis of the enzymatic products via gas chromatography coupled to mass spectrometry showed that the introduction of the (E) double bond between C-4 and C-5 occurs in two discrete steps: cleavage of the C4-H bond was found to be very sensitive to isotopic substitution (k(H)/k(D) = 8.0 +/- 0.8), while a negligible isotope effect (k(H)/k(D) = 1.02 +/- 0.07) was observed for the C5-H bond-breaking step. According to a mechanistic model that we have previously proposed, these results suggest that initial oxidation for this desaturation reaction occurs at C-4. This finding correlates nicely with the observation that 4-hydroxylated products are produced from a similar substrate by a closely related oxidative enzyme in yeast.  相似文献   

16.
Reaction rate constants have been acquired for the gaseous unimolecular decomposition reaction of the Co(+)(OC(CH(3))(2)) cluster ion and its deuterium labeled analog. Each rate constant is measured at a well resolved cluster internal energy within the range 12,300-16,100 cm(-1). The weighted, averaged kinetic isotope effect (KIE), k(H)/k(D) = 1.54 ± 0.05, is about three times smaller than the KIE measured for the rate-determining rate constants in the similar Ni(+)(OC(CH(3))(2)) decomposition reaction. These reactions likely follow the same oxidative addition-reductive elimination mechanism. Thus, this unexpected change in the KIE magnitudes is not due to differences in the dissociative reaction coordinates. Rather, we propose that the unique dissociation dynamics of these two similar systems is due to differences in the low-lying electronic structure of each transition metal ion.  相似文献   

17.
The solvolysis of α-d-glucopyranosyl fluoride in hexafluoro-2-propanol gives two products, 1,1,1,3,3,3-hexafluoropropan-2-yl α-d-glucopyranoside and 1,6-anhydro-β-D-glucopyranose. The ratio of these two products is essentially unchanged for reactions that are performed between 56 and 100 °C. The activation parameters for the solvolysis reaction are as follows: ΔH(++) = 81.4 ± 1.7 kJ mol(-1), and ΔS(++) = -90.3 ± 4.6 J mol(-1) K(-1). To characterize, by use of multiple kinetic isotope effect (KIE) measurements, the TS for the solvolysis reaction in hexafluoro-2-propanol, we synthesized a series of isotopically labeled α-d-glucopyranosyl fluorides. The measured KIEs for the C1 deuterium, C2 deuterium, C5 deuterium, anomeric carbon, ring oxygen, O6, and solvent deuterium are 1.185 ± 0.006, 1.080 ± 0.010, 0.987 ± 0.007, 1.008 ± 0.007, 0.997 ± 0.006, 1.003 ± 0.007, and 1.68 ± 0.07, respectively. The transition state for the solvolysis reaction was modeled computationally using the experimental KIE values as constraints. Taken together, the reported data are consistent with the retained solvolysis product being formed in an S(N)i (D(N)(++)*A(Nss)) reaction with a late transition state in which cleavage of the glycosidic bond is coupled to the transfer of a proton from a solvating hexafluoro-2-propanol molecule. In comparison, the inverted product, 1,6-anhydro-β-D-glucopyranose, is formed by intramolecular capture of a solvent-equilibrated glucopyranosylium ion, which results from dissociation of the solvent-separated ion pair formed in the rate-limiting ionization reaction (D(N)(++) + A(N)). The implications that this model reaction have for the mode of action of retaining glycosyltransferases are discussed.  相似文献   

18.
The oxidation of hydroxylamine by [IrCl6]2- has been studied spectrophotometrically in deoxygenated aqueous solutions in the range of pH 4-9 at 25 degrees C. The reaction is catalyzed by Cu2+, Fe2+, and impurities of aquochloroiridium complexes. Oxalate is a very effective inhibitor of catalysis by copper and iron ions. With excess hydroxylamine, the reaction follows pseudo-first-order kinetics, and the stoichiometric ratio (DeltanIr(IV)/Deltanhydroxylamine) is 1.05 at pH 5.9. Over the pH range 4.2-8.8, the empirical rate law is -d[IrCl(6)2-]/dt=k[IrCl6(2-)][NH2OH]tot, with k=k1Ka1/([H+]+Ka1)+k'Ka1/([H+]([H+]+Ka1)), where Ka1 is the dissociation constant of NH3OH+. Least-squares fitting yields k1=(17.05+/-0.47) M-1 s(-1) and k'=(2.59+/-0.09)x10(-6) s(-1) at ionic strength of 0.1 M (adjusted by NaClO4) and 25 degrees C. The kinetic isotope effects (KIE) (kH/kD) for k1 and k' are 4.4 and 9.8, correspondingly. A mechanism is inferred in which k1 corresponds to concerted proton-coupled electron transfer (PCET) and k' corresponds to electron transfer from NH2O-. In this mechanism, the large KIE for k' is due almost entirely to the equilibrium isotope effect for the pKa of NH2OH.  相似文献   

19.
The mutarotation of N-acetylneuraminic acid (Neu5Ac) proceeds by four kinetically distinct pathways: (i) the acid-catalyzed reaction of neutral Neu5Ac; (ii) the spontaneous reaction of the carboxylic acid (the kinetically equivalent acid-catalyzed reaction on the anion being ruled out by the solvent deuterium kinetic isotope effect of 3.74 ± 0.68); (iii) a spontaneous, water-catalyzed, reaction of the anion; and (iv) a specific-base catalyzed reaction of the anion. The magnitude of the solvent kinetic isotope effect, k(H2O)/k(D2O) = 4.48 ± 0.74 is consistent with a ring-opening transition state in which a water molecule is deprotonating the anomeric hydroxyl group in concert with strengthening solvation of the ring oxygen atom. The mechanistic implications for Neu5Ac mutarotases are discussed.  相似文献   

20.
Song N  Stanbury DM 《Inorganic chemistry》2011,50(24):12762-12773
It has been previously established that the aqueous oxidation of phenol by a deficiency of [IrCl(6)](2-) proceeds through the production of [IrCl(6)](3-) and phenoxyl radicals. Coupling of the phenoxyl radicals leads primarily to 4,4'-biphenol, 2,2'-biphenol, 2,4'-biphenol, and 4-phenoxyphenol. Overoxidation occurs through the further oxidation of these coupling products, leading to a rather complex mixture of final products. The rate laws for oxidation of the four coupling products by [IrCl(6)](2-) have the same form as those for the oxidation of phenol itself: -d[Ir(IV)]/dt = {(k(ArOH) + k(ArO(-))K(a)/[H(+)])/(1 + K(a)/[H(+)])}[ArOH](tot)[Ir(IV)]. Values for k(ArOH) and k(ArO(-)) have been determined for the four substrates at 25 °C and are assigned to H(2)O-PCET and electron-transfer mechanisms, respectively. Kinetic simulations of a combined mechanism that includes the rate of oxidation of phenol as well as the rates of these overoxidation steps show that the degree of overoxidation is rather limited at high pH but quite extensive at low pH. This pH-dependent overoxidation leads to a pH-dependent stoichiometric factor in the rate law for oxidation of phenol and causes some minor deviations in the rate law for oxidation of phenol. Empirically, these minor deviations can be accommodated by the introduction of a third term in the rate law that includes a "pH-dependent rate constant", but this approach masks the mechanistic origins of the effect.  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