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.
Bacterial tRNA-specific adenosine deaminase (TadA) catalyzes the essential deamination of adenosine to inosine at the wobble position of tRNAs and is necessary to permit a single tRNA species to recognize multiple codons. The transition state structure of Escherichia coli TadA was characterized by kinetic isotope effects (KIEs) and quantum chemical calculations. A stem loop of E. coli tRNA(Arg2) was used as a minimized TadA substrate, and its adenylate editing site was isotopically labeled as [1'-(3)H], [5'-(3)H2], [1'-(14)C], [6-(13)C], [6-(15)N], [6-(13)C, 6-(15)N] and [1-(15)N]. The intrinsic KIEs of 1.014, 1.022, 0.994, 1.014 and 0.963 were obtained for [6-(13)C]-, [6-(15)N]-, [1-(15)N]-, [1'-(3)H]-, [5'-(3)H2]-labeled substrates, respectively. The suite of KIEs are consistent with a late SNAr transition state with a complete, pro-S-face hydroxyl attack and nearly complete N1 protonation. A significant N6-C6 dissociation at the transition state of TadA is indicated by the large [6-(15)N] KIE of 1.022 and corresponds to an N6-C6 distance of 2.0 A in the transition state structure. Another remarkable feature of the E. coli TadA transition state structure is the Glu70-mediated, partial proton transfer from the hydroxyl nucleophile to the N6 leaving group. KIEs correspond to H-O and H-N distances of 2.02 and 1.60 A, respectively. The large inverse [5'-(3)H] KIE of -3.7% and modest normal [1'-(3)H] KIE of 1.4% indicate that significant ribosyl 5'-reconfiguration and purine rotation occur on the path to the transition state. The late SNAr transition-state established here for E. coli TadA is similar to the late transition state reported for cytidine deaminase. It differs from the early SNAr transition states described recently for the adenosine deaminases from human, bovine, and Plasmodium falciparum sources. The ecTadA transition state structure reveals the detailed architecture for enzymatic catalysis. This approach should be readily transferable for transition state characterization of other RNA editing enzymes.  相似文献   

2.
Kinetic isotope effects (KIEs) and computer modeling are used to approximate the transition state of S. pneumoniae 5'-methylthioadenosine/S-adenosylhomocysteine nucleosidase (MTAN). Experimental KIEs were measured and corrected for a small forward commitment factor. Intrinsic KIEs were obtained for [1'-3H], [1'-14C], [2'-3H], [4'-3H], [5'-3H(2)], [9-15N] and [Me-3H(3)] MTAs. The intrinsic KIEs suggest an SN1 transition state with no covalent participation of the adenine or the water nucleophile. The transition state was modeled as a stable ribooxacarbenium ion intermediate and was constrained to fit the intrinsic KIEs. The isotope effects predicted a 3-endo conformation for the ribosyl oxacarbenium-ion corresponding to H1'-C1'-C2'-H2' dihedral angle of 70 degrees. Ab initio Hartree-Fock and DFT calculations were performed to study the effect of polarization of ribosyl hydroxyls, torsional angles, and the effect of base orientation on isotope effects. Calculations suggest that the 4'-3H KIE arises from hyperconjugation between the lonepair (n(p)) of O4' and the sigma* (C4'-H4') antibonding orbital owing to polarization of the 3'-hydroxyl by Glu174. A [methyl-3H(3)] KIE is due to hyperconjugation between np of sulfur and sigma* of methyl C-H bonds. The van der Waal contacts increase the 1'-3H KIE because of induced dipole-dipole interactions. The 1'-3H KIE is also influenced by the torsion angles of adjacent atoms and by polarization of the 2'-hydroxyl. Changing the virtual solvent (dielectric constant) does not influence the isotope effects. Unlike most N-ribosyltransferases, N7 of the leaving group adenine is not protonated at the transition state of S. pneumoniae MTAN. This feature differentiates the S. pneumoniae and E. coli transition states and explains the 10(3)-fold decrease in the catalytic efficiency of S. pneumoniae MTAN relative to that from E. coli.  相似文献   

3.
Kinetic isotope effects (KIEs) and computer modeling using density functional theory were used to approximate the transition state of human 5'-methylthioadenosine phosphorylase (MTAP). KIEs were measured on the arsenolysis of 5'-methylthioadenosine (MTA) catalyzed by MTAP and were corrected for the forward commitment to catalysis. Intrinsic KIEs were obtained for [1'-(3)H], [1'-(14)C], [2'-(3)H], [4'-(3)H], [5'-(3)H(2)], [9-(15)N], and [Me-(3)H(3)] MTAs. The primary intrinsic KIEs (1'-(14)C and 9-(15)N) suggest that MTAP has a dissociative S(N)1 transition state with its cationic center at the anomeric carbon and insignificant bond order to the leaving group. The 9-(15)N intrinsic KIE of 1.039 also establishes an anionic character for the adenine leaving group, whereas the alpha-primary 1'-(14)C KIE of 1.031 indicates significant nucleophilic participation at the transition state. Computational matching of the calculated EIEs to the intrinsic isotope effects places the oxygen nucleophile 2.0 Angstrom from the anomeric carbon. The 4'-(3)H KIE is sensitive to the polarization of the 3'-OH group. Calculations suggest that a 4'-(3)H KIE of 1.047 is consistent with ionization of the 3'-OH group, indicating formation of a zwitterion at the transition state. The transition state has cationic character at the anomeric carbon and is anionic at the 3'-OH oxygen, with an anionic leaving group. The isotope effects predicted a 3'-endo conformation for the ribosyl zwitterion, corresponding to a H1'-C1'-C2'-H2' torsional angle of 33 degrees. The [Me-(3)H(3)] and [5'-(3)H(2)] KIEs arise predominantly from the negative hyperconjugation of the lone pairs of sulfur with the sigma (C-H) antibonding orbitals. Human MTAP is characterized by a late S(N)1 transition state with significant participation of the phosphate nucleophile.  相似文献   

4.
The remote 5'-3H V/K kinetic isotope effect (KIE) observed in human thymidine phosphorylase (6.1%) is significantly larger than can be explained by the reaction chemistry. One hypothesis connects the 5'-3H KIE in purine nucleoside phosphorylase to that enzyme's SN1 transition state. The transition state of thymidine phosphorylase, however, is an SN2 nucleophilic displacement. Here we report equilibrium binding isotope effects sufficiently large to explain the presence of this substantial KIE in thymidine phosphorylase.  相似文献   

5.
Recombinant human thymidine phosphorylase catalyzes the reaction of arsenate with thymidine to form thymine and 2-deoxyribose 1-arsenate, which rapidly decomposes to 2-deoxyribose and inorganic arsenate. The transition-state structure of this reaction was determined using kinetic isotope effect analysis followed by computer modeling. Experimental kinetic isotope effects were determined at physiological pH and 37 degrees C. The extent of forward commitment to catalysis was determined by pulse-chase experiments to be 0.70%. The intrinsic kinetic isotope effects for [1'-(3)H]-, [2'R-(3)H]-, [2'S-(3)H]-, [4'-(3)H]-, [5'-(3)H]-, [1'-(14)C]-, and [1-(15)N]-thymidines were determined to be 0.989 +/- 0.002, 0.974 +/- 0.002, 1.036 +/- 0.002, 1.020 +/- 0.003, 1.061 +/- 0.003, 1.139 +/- 0.005, and 1.022 +/- 0.005, respectively. A computer-generated model, based on density functional electronic structure calculations, was fit to the experimental isotope effect. The structure of the transition state confirms that human thymidine phosphorylase proceeds through an S(N)2-like transition state with bond orders of 0.50 to the thymine leaving group and 0.33 to the attacking oxygen nucleophile. The reaction differs from the dissociative transition states previously reported for N-ribosyl transferases and is the first demonstration of a nucleophilic transition state for an N-ribosyl transferase. The large primary (14)C isotope effect of 1.139 can occur only in nucleophilic displacements and is the largest (14)C primary isotope effect reported for an enzymatic reaction. A transition state structure with substantial bond order to the attacking nucleophile and leaving group is confirmed by the slightly inverse 1'-(3)H isotope effect, demonstrating that the transition state is compressed by the impinging steric bulk of the nucleophile and leaving group.  相似文献   

6.
Adenosine deaminases (ADAs) from human, bovine, and Plasmodium falciparum sources were analyzed by kinetic isotope effects (KIEs) and shown to have distinct but related transition states. Human adenosine deaminase (HsADA) is present in most mammalian cells and is involved in B- and T-cell development. The ADA from Plasmodium falciparum (PfADA) is essential in this purine auxotroph, and its inhibition is expected to have therapeutic effects for malaria. Therefore, ADA is of continuing interest for inhibitor design. Stable structural mimics of ADA transition states are powerful inhibitors. Here we report the transition-state structures of PfADA, HsADA, and bovine ADA (BtADA) solved using competitive kinetic isotope effects (KIE) and density functional calculations. Adenines labeled at [6-13C], [6-15N], [6-13C, 6-15N], and [1-15N] were synthesized and enzymatically coupled with [1'-14C] ribose to give isotopically labeled adenosines as ADA substrates for KIE analysis. [6-13C], [6-15N], and [1-15N]adenosines reported intrinsic KIE values of (1.010, 1.011, 1.009), (1.005, 1.005, 1.002), and (1.004, 1.001, 0.995) for PfADA, HsADA, and BtADA, respectively. The differences in intrinsic KIEs reflect structural alterations in the transition states. The [1-15N] KIEs and computational modeling results indicate that PfADA, HsADA, and BtADA adopt early SNAr transition states, where N1 protonation is partial and the bond order to the attacking hydroxyl nucleophile is nearly complete. The key structural variation among PfADA, HsADA, and BtADA transition states lies in the degree of N1 protonation with the decreased bond lengths of 1.92, 1.55, and 1.28 A, respectively. Thus, PfADA has the earliest and BtADA has the most developed transition state. This conclusion is consistent with the 20-36-fold increase of kcat in comparing PfADA with HsADA and BtADA.  相似文献   

7.
Nucleophile (11)C/ (14)C [ k (11)/ k (14)] and secondary alpha-deuterium [( k H/ k D) alpha] kinetic isotope effects (KIEs) were measured for the S N2 reactions between tetrabutylammonium cyanide and ethyl iodide, bromide, chloride, and tosylate in anhydrous DMSO at 20 degrees C to determine whether these isotope effects can be used to determine the structure of S N2 transition states. Interpreting the experimental KIEs in the usual fashion (i.e., that a smaller nucleophile KIE indicates the Nu-C alpha transition state bond is shorter and a smaller ( k H/ k D) alpha is found when the Nu-LG distance in the transition state is shorter) suggests that the transition state is tighter with a slightly shorter NC-C alpha bond and a much shorter C alpha-LG bond when the substrate has a poorer halogen leaving group. Theoretical calculations at the B3LYP/aug-cc-pVDZ level of theory support this conclusion. The results show that the experimental nucleophile (11)C/ (14)C KIEs can be used to determine transition-state structure in different reactions and that the usual method of interpreting these KIEs is correct. The magnitude of the experimental secondary alpha-deuterium KIE is related to the nucleophile-leaving group distance in the S N2 transition state ( R TS) for reactions with a halogen leaving group. Unfortunately, the calculated and experimental ( k H/ k D) alpha's change oppositely with leaving group ability. However, the calculated ( k H/ k D) alpha's duplicate both the trend in the KIE with leaving group ability and the magnitude of the ( k H/ k D) alpha's for the ethyl halide reactions when different scale factors are used for the high and the low energy vibrations. This suggests it is critical that different scaling factors for the low and high energy vibrations be used if one wishes to duplicate experimental ( k H/ k D) alpha's. Finally, neither the experimental nor the theoretical secondary alpha-deuterium KIEs for the ethyl tosylate reaction fit the trend found for the reactions with a halogen leaving group. This presumably is found because of the bulky (sterically hindered) leaving group in the tosylate reaction. From every prospective, the tosylate reaction is too different from the halogen reactions to be compared.  相似文献   

8.
The mechanistic details of aerobic alcohol oxidation with catalytic Pd(IiPr)(OAc)(2)(H(2)O) (IiPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) are disclosed. Under optimal conditions, beta-hydride elimination is rate-limiting supported by kinetic studies including a high primary kinetic isotope effect (KIE) value of 5.5 +/- 0.1 and a Hammett rho value of -0.48 +/- 0.04. On the basis of these studies, a late transition state is proposed for beta-hydride elimination, which is further corroborated by theoretical calculations using density functional theory. Additive acetic acid modulates the rates of both the alcohol oxidation sequence and regeneration of the Pd catalyst. With no additive [HOAc], turnover-limiting reprotonation of intermediate palladium peroxo is kinetically competitive with beta-hydride elimination, allowing for reversible oxygenation and decomposition of Pd(0). With additive [HOAc] (>2 mol %), reprotonation of the palladium peroxo is fast and beta-hydride elimination is the single rate-controlling step. This proposal is supported by an apparent decomposition pathway modulated by [HOAc], a change in alcohol concentration dependence, a lack of [O(2)] dependence at high [HOAc], and significant changes in the KIE values at different HOAc concentrations.  相似文献   

9.
A capillary gas chromatographic-mass spectrometric method for the simultaneous determination of stable isotopically labelled L-histidine (L-[3,3-2H2,1',3'-15N2]histidine, L-His-[M + 4]) and urocanic acid ([3-2H,1',3'-15N2]urocanic acid, UA-[M + 3]) in human plasma was developed using DL-[2,3,3,5'-2H4,2'-13C,1',3'-15N2]histidine (DL-His-[M + 7]) and [2,3,5'-2H3,2'-13C,1',3'-15N2]urocanic acid (UA-[M + 6]) as internal standards. L-Histidine and urocanic acid were derivatized to alpha N-(trifluoroacetyl)-imN-(ethoxycarbonyl)-L-histidine n-butyl ester and imN-(ethoxycarbonyl)urocanic acid n-butyl ester. Quantification was carried out by selected ion monitoring of the molecular ions of the respective derivatives of L-His-[M + 4], DL-His-[M + 7], UA-[M + 3] and UA-[M + 6]. The sensitivity, specificity, precision and accuracy of the method were demonstrated to be satisfactory for measuring plasma concentrations of L-His-[M + 4] and UA-[M + 3] following administration of trace amounts of L-His-[M + 4] to humans.  相似文献   

10.
A one-dimensional supramolecular head-to-tail N+ -H...N-type hydrogen-bonded chain of the complex [FeII(L)2H](ClO4)3.MeOH [L = 4'-(4'-pyridyl)-1,2':6'1'-bis(pyrazolyl)pyridine] exhibits a reversible, thermally driven spin transition at 286 K with a hysteresis loop of ca. 2 K.  相似文献   

11.
12.
Variational transition state theory has been used to calculate the kinetic isotope effects affecting product ratios in the reaction between (1)O(2) and d(6)-tetramethylethylene. The minimum energy path on the potential energy surface for this process reaches a valley-ridge inflection point and then bifurcates leading to the two final products. Using canonical variational transition state theory, two distinct dynamical bottlenecks were located corresponding to the H- and the D-abstraction, respectively. The calculated KIE at 263 K turns out to be 1.126. Analogously, a H/T KIE of 1.17 at the same temperature has been found for the reaction of (1)O(2) with the tritiated derivative of tetramethylethylene.  相似文献   

13.
Palladium-catalyzed intramolecular cyclization of N-(N'-tert-butylformimidoyl)-6-[2-(2-iodophenyl)ethyl]-1,2,3,4-tetrahydropyridine (1a) and N-(N'-tert-butylformimidoyl)-6-[3-(2-iodophenyl)propyl]-1,2,3,4-tetrahydropyridine (1b) respectively results in formation of spiro compounds 1'-(N-tert-butylformimidoyl)-3',4'-dihydrospiro[indan-1,2'(1'H)-pyridine] (4a), 1'-(N-tert-butylformimidoyl)-1',6'-dihydrospiro[indan-1,2'(3'H)-pyridine] (5a), and 1'-(N-tert-butylformimidoyl)-5',6'-dihydrospiro[indan-1,2'(1'H)-pyridine] (6a) and 1'-(N-tert-butylformimidoyl)-3,3',4,4'-tetrahydrospiro[naphthalene-1(2H),2'(1'H)-pyridine] (4b), 1'-(N-tert-butylformimidoyl)-1',3,4,6'-tetrahydrospiro[naphthalene-1(2H),2'(3'H)-pyridine] (5b), and 1'-(N-tert-butylformimidoyl)-3,4,5',6'-tetrahydrospiro[naphthalene-1(2H),2'(1'H)-pyridine] (6b). The double-bond migration process can be controlled, and any of the three double-bond isomers can be prepared by employing proper ligands. A combination of BINAP and the amidine function was required to obtain the isomers 5a and 5b with the double bond in the homoallylic position relative to the aryl group. An electrospray ionization mass spectrometric study was conducted to support suggested reaction intermediates.  相似文献   

14.
The overall reductive elimination of RH from the ansa-molybdenocene and -tungstenocene complexes [Me(2)Si(C(5)Me(4))(2)]Mo(Ph)H and [Me(2)Si(C(5)Me(4))(2)]W(R)H (R = Me, Ph) is characterized by an inverse primary kinetic isotope effect (KIE) for the tungsten system but a normal KIE for the molybdenum system. Oxidative addition of PhH to [[Me(2)Si(C(5)Me(4))(2)]M] also differs for the two systems, with the molybdenum system exhibiting a substantial intermolecular KIE, while no effect is observed for the tungsten system. These differences in KIEs indicate a significant difference in the reactivity of the hydrocarbon adducts [Me(2)Si(C(5)Me(4))(2)]M(RH) for the molybdenum and tungsten systems. Specifically, oxidative cleavage of [Me(2)Si(C(5)Me(4))(2)]M(RH) is favored over RH dissociation for the tungsten system, whereas RH dissociation is favored for the molybdenum system. A kinetics analysis of the interconversion of [Me(2)Si(C(5)Me(4))(2)]W(CH(3))D and [Me(2)Si(C(5)Me(4))(2)]W(CH(2)D)H, accompanied by elimination of methane, provides evidence that the reductive coupling step in this system is characterized by a normal KIE. This observation demonstrates that the inverse KIE for overall reductive elimination is a result of an inverse equilibrium isotope effect (EIE) and is not a result of an inverse KIE for a single step. A previous report of an inverse kinetic isotope effect of 0.76 for C-H reductive coupling in the [Tp]Pt(CH(3))H(2) system is shown to be erroneous. Finally, a computational study provides evidence that the reductive coupling of [Me(2)Si(C(5)Me(4))(2)]W(Ph)H proceeds via the initial formation of a benzene sigma-complex, rather than an eta(2)-pi-benzene complex.  相似文献   

15.
Density functional calculations were performed in response to the controversies regarding the identity of the oxidant species in cytochrome P450. The calculations were used to gauge the relative C-H hydroxylation reactivity of three potential oxidant species of the enzyme, the high-valent oxo-iron species Compound I (Cpd I), the ferric hydroperoxide Compound 0 (Cpd 0), and the ferric-hydrogen peroxide complex Fe(H(2)O(2)). The results for the hydroxylation of a radical probe substrate, 1, show the following trends: (a) Cpd I is the most reactive species; in its presence the other two reagents will be silent. (b) In the absence of Cpd I, substrate oxidation by Cpd 0 and Fe(H(2)O(2)) will take place via a stepwise mechanism that involves initial O-O homolysis followed by H-abstraction from 1. (c) Cpd 0 will undergo mostly porphyrin hydroxylation and only approximately 15% of substrate oxidation producing mostly the rearranged alcohol, 3 (Scheme 2). (d) Fe(H(2)O(2)) will generate mostly free hydrogen peroxide (uncoupling). A small fraction will perform substrate oxidation and lead mostly to 3. Reactivity probes for these reagents are kinetic isotope effect (KIE) and the product ratio of unrearranged to rearranged alcohols, [2/3]. Thus, for substrate oxidation by Cpd 0 or Fe(H(2)O(2)) KIE will be small, approximately 2, while Cpd I will have large KIE values. Typically both Cpd 0 and Fe(H(2)O(2)) will lead to a [2/3] ratio < 1, while Cpd I will lead to ratios > 1. In addition, the product isotope effect (KIE(2)/KIE(3) not equal 1) is expected from the reactivity of Cpd I.  相似文献   

16.
The mechanism of cyclopropenations of alkynes with ethyl diazoacetate catalyzed by Rh2(OAc)4 and Rh2(OAc)(DPTI)3 (1) is studied by a combination of kinetic isotope effects and theoretical calculations. With each catalyst, a significant normal 13C KIE was observed for the terminal acetylenic carbon, while a very small 13C KIE was observed at the internal acetylenic carbon. These isotope effects are predicted well from canonical variational transition structures for cyclopropenations with intact tetrabridged rhodium carbenoids. A viable mechanism based on the recently proposed importance of a [2 + 2] cycloaddition on a tribridged rhodium carbenoid could not be identified. An explanation for the enantioselectivity with DPTI ligands is described.  相似文献   

17.
The isomeric ferrocene phosphine-carboxamides, 1-(diphenylphosphino)-1'-{[N-(2-pyridyl)-methyl]carbamoyl}ferrocene (1) and 1-(diphenylphosphino)-1'-{[N-(4-pyridyl)methyl]carbamoyl}ferrocene (2) have been studied as ligands in group-12 metal bromide complexes. The reactions of 1 with CdBr2 x 4H2O and HgBr(2) at 1:1 mole ratio gave the discrete tetracadmium complex [Cd2(micro-Br)2(-1kappa2O,N2)2[micro-1kappa2O,N2:2kappaP-(C5H4N)CH2NHC(O)fcPPh2-CdBr3]2] (7; fc = ferrocene-1,1'-diyl) and the halogeno-bridged dimer [[Hg(micro-Br)Br(-kappaP)]2] (8), respectively. In the presence of acetic acid, the CdBr2-1 system furnished a zwitterionic complex featuring protonated 1 as the P-monodentate donor, [CdBr3[Ph2PfcC(O)NHCH2(C5H4NH)-kappaP]] x H2O (6 x H2O). Under neutral conditions, compound , whose terminal donor groups are better arranged for the formation of extended assemblies, gave rise to one-dimensional coordination polymers [MBr2[micro(P,N)-]](n) (M = Cd, 4; M = Hg, ). The crystal structures of 2 x H2O, its corresponding phosphine oxide (3 x H2O), and complexes 4, 5, 6 x H2O, and have been determined, revealing extensive hydrogen bonding interactions in the solid state.  相似文献   

18.
Solid complexes of lanthanide picrates with N,N-dibenzyl-2-[2'-[(dibenzylcarbamoyl)-methoxy]-[1,1']binaphthalenyl-2-yloxy]-acetamide (L), [Ln(pic)(3)L] (Ln = La, Nd, Eu, Gd, Tb, Dy, Y), have been prepared and characterized by elemental analysis, IR, 1H NMR, UV-Vis spectra and conductivity measurements. The fluorescence property of the europium complex in solid state and in CHCl(3), acetone, AcOEt and DMF was studied.  相似文献   

19.
Rates, kinetic isotope effects (KIE), and Swain-Schaad exponents (SSE) have been calculated for a variety of isotopologues for the [1,5] shift in (Z)-1,3-pentadiene using mPW1K/6-31+G(d,p). Quantum mechanical effects along the reaction coordinate were incorporated with the zero-curvature tunneling (ZCT) model and with the multidimensional small curvature tunneling (SCT) model, which allows for coupling of modes perpendicular to the reaction coordinate. The latter model gives the best agreement with experimental rates and primary KIEs. The small quasiclassical primary KIE (2.6) is rationalized in terms of a nonlinear transition state. For sp3 to sp2 rehybridization, the quasiclassical alpha-secondary KIE shows an unusual inverse effect due to compression of the nonbonding hydrogens in the suprafacial transition state. SCT transmission coefficients (kappa) increase the rates by as much as one order of magnitude. Tunneling allows the reactant to evade 1-2.5 kcal/mol of the barrier depending on the isotope. Inclusion of tunneling in the secondary KIE increases it beyond the equilibrium isotope effect and converts the inverse effect (0.95) into a normal KIE (1.12). Tunneling was found to deflate the primary y SSE but by an amount too small to distinguish it from the quasiclassical SSE. On the other hand, when a specific labeling pattern is used, the difference between the quasiclassical secondary SSE (4.1) and the tunneling secondary SSE (2.3) may be sufficiently large to detect tunneling. The mixed secondary SSE shows even larger differences.  相似文献   

20.
Dry carbonic acid has recently been shown to be kinetically stable even at room temperature. Addition of water molecules reduces this stability significantly, and the decomposition (H2CO3 + nH2O --> (n+1)H2O + CO2) is extremely accelerated for n = 1, 2, 3. By including two water molecules, a reaction rate that is a factor of 3000 below the experimental one (10 s(-1)) at room temperature was found. In order to further remove the gap between experiment and theory, we increased the number of water molecules involved to 3 and took into consideration different mechanisms for thorough elucidation of the reaction. A mechanism whereby the reaction proceedes via a six-membered transition state turns out to be the most efficient one over the whole examined temperature range. The determined reaction rates approach experimental values in aqueous solution reasonably well; most especially, a significant increase in the rates in comparison to the decomposition reaction with fewer water molecules is found. Further agreement with experiment is found in the kinetic isotope effects (KIE) for the deuterated species. For water-free carbonic acid, the KIE (i.e., kH2CO3/kD2CO3) for the decomposition reaction is predicted to be 220 at 300 K, whereas it amounts to 2.2-3.0 for the investigated mechanisms including three water molecules. This result is therefore reasonably close to the experimental value of 2 (at 300 K). These KIEs are in much better accordance with the experiment than the KIE for decomposition with fewer water entities.  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