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 125 毫秒
1.
Purple acid phosphatase from pig uterine fluid (uteroferrin), a representative of the diverse family of binuclear metallohydrolases, requires a heterovalent Fe(III)Fe(II) center for catalytic activity. The active-site structure and reaction mechanism of this enzyme were probed with a combination of methods including metal ion replacement and biomimetic studies. Specifically, the asymmetric ligand 2-bis{[(2-pyridylmethyl)-aminomethyl]-6-[(2-hydroxybenzyl)(2-pyridylmethyl)]aminomethyl}-4-methylphenol and two symmetric analogues that contain the softer and harder sites of the asymmetric unit were employed to assess the site selectivity of the trivalent and divalent metal ions using (71)Ga NMR, mass spectrometry and X-ray crystallography. An exclusive preference of the harder site of the asymmetric ligand for the trivalent metal ion was observed. Comparison of the reactivities of the biomimetics with Ga(III)Zn(II) and Fe(III)Zn(II) centers indicates a higher turnover for the former, suggesting that the M(III)-bound hydroxide acts as the reaction-initiating nucleophile. Catalytically active Ga(III)Zn(II) and Fe(III)Zn(II) derivatives were also generated in the active site of uteroferrin. As in the case of the biomimetics, the Ga(III) derivative has increased reactivity, and a comparison of the pH dependence of the catalytic parameters of native uteroferrin and its metal ion derivatives supports a flexible mechanistic strategy whereby both the mu-(hydr)oxide and the terminal M(III)-bound hydroxide can act as nucleophiles, depending on the metal ion composition, the geometry of the second coordination sphere and the substrate.  相似文献   

2.
3.
The function of Mn(III) in plant acid phosphatase has been investigated by a metal-substitution study, and some properties of the Fe(III)-substituted enzyme were compared with those of the native Mn(III) enzyme and mammalian Fe(III)-containing acid phosphatases. 19F nuclear magnetic resonance (NMR) and proton relaxation rate measurements showed that inhibitors such as F and nitrilotriacetic acid interact with paramagnetic Mn(III) active site. The 31P-NMR signal of the enzyme-phosphate complex was also broadened by the paramagnetic effect of Mn(III). In the metal-substitution experiments of the Mn(III)-acid phosphatase with Fe(III), Zn(II) and Cu(II), only the iron gave satisfactory substitution. The Fe(III)-substituted plant acid phosphatase exhibited an absorption maximum at 525 nm (ε = 3000), typical high spin ferric ESR signal at g = 4.39, and lower pH optimum (pH 4.8) than the native Mn(III)-enzyme (pH 5.8). The phosphatase activity of the Fe(III)-substituted enzyme was reduced to about 53% of that of the native enzyme. The substrate specificities of both metallophosphatases were remarkably similar, but different from that of the Fe(III)-containing uteroferrin. The present results indicate that Mn(III) and Fe(IIII) in the acid phosphatase play an important role on effective binding of phosphate and acceleration of hydrolysis of phosphomonoesters at pH 4–6.  相似文献   

4.
The choreography of restriction endonuclease catalysis is a long-standing paradigm in molecular biology. Bivalent metal ions are required almost for all PD..D/ExK type enzymes, but the number of cofactors essential for the DNA backbone scission remained ambiguous. On the basis of crystal structures and biochemical data for various restriction enzymes, three models have been developed that assign critical roles for one, two, or three metal ions during the phosphodiester hydrolysis. To resolve this apparent controversy, we investigated the mechanism of BamHI catalysis using quantum mechanical/molecular mechanical simulation techniques and determined the activation barriers of three possible pathways that involve a Glu-113 or a neighboring water molecule as a general base or an external nucleophile that penetrated from bulk solution. The extrinsic mechanism was found to be the most favorable with an activation free energy of 23.4 kcal/mol, in reasonable agreement with the experimental data. On the basis of the effect of the individual metal ions on the activation barrier, metal ion A was concluded to be pivotal for the reaction, while the enzyme lacking metal ion B still has moderate efficiency. Thus, we propose that the catalytic scheme of BamHI does not involve a general base for nucleophile generation and requires one obligatory metal ion for catalysis that stabilizes the attacking nucleophile and coordinates it throughout the nucleophilic attack. Such a model may also explain the variation in the number of metal ions in the crystal structures and thus could serve as a framework for a unified catalytic scheme of type II restriction endonucleases.  相似文献   

5.
The dinuclear aminopeptidase from Streptomyces griseus (SgAP) and its metal derivatives catalyze the hydrolysis of the phosphoester bis(p-nitrophenyl) phosphate (BNPP) and the phosphonate ester p-nitrophenyl phenylphosphonate with extraordinary rate enhancements at pH 7.0 and 25 °C [A. Ercan, H. I. Park, L.-J. Ming, Biochemistry 45, (2006) 13779-13793.], reaching 6.7 billion-fold in terms of the first-order rate constant of the di-Co(II) derivative with respect to the autohydrolytic rates. Since phosphoesters are transition state-like inhibitors in peptide hydrolysis, their hydrolysis by SgAP is quite novel. Herein, we report the investigation of this proficient alternative catalysis of SgAP and the role of each metal ion in the dinuclear site toward peptide and BNPP hydrolysis. Mn(II) selectively binds to one of the dinuclear metal sites (M1), affording MnE-SgAP with an empty (E) second site for the binding of another metal (M2), including Mn(II), Co(II), Ni(II), Zn(II), and Cd(II). Peptide hydrolysis is controlled by M2, wherein the kcat values for the derivatives MnM2-SgAP are different yet similar between MnCo- and CoCo-SgAP and pairs of other metal derivatives. On the other hand, BNPP hydrolysis is affected by metals in both sites. Thus, the two hydrolytic catalyses must follow different mechanisms. Based on crystal structures, docking, and the results presented herein, the M1 site is close to the hydrophobic specific site and the M2 site is next to Tyr246 that is H-bonded to a coordinated nucleophilic water molecule in peptide hydrolysis; whereas a coordinated water molecule on M1 becomes available as the nucleophile in phosphodiester hydrolysis.  相似文献   

6.
Uteroferrin, a purple-colored, iron-containing glycoprotein, purified from the uterine fluid of progesterone-stimulated pigs, owes its natural purple color to a broad absorption band centered at 545 nm. Laser excitation within the visible absorption band of uteroferrin results in an intense resonance Raman spectrum which bears a striking resemblance to that reported for Fe(III)-transferrin, the iron transport protein of serum. Excitation profiles for the four resonance-enhanced bands of uteroferrin were obtained from 4579 A to 6741 A, using lines from Ar+ and Kr+ lasers. Each of the profiles have maxima near 545 nm. The spectral similarities of uteroferrin and Fe(III)-transferrin lead to the belief that the Fe(III) binding sites of the two proteins must be, at least in some respects, quite similar. In particular, it is concluded that, as in Fe(III)-transferrin, the metal binding site of uteroferrin contains a tyrosine ligand and that the visible absorption spectrum of uteroferrin results from a phenolate to Fe(III) charge transfer.  相似文献   

7.
The Cu(II) or Ni(II) ion-catalyzed hydrolysis of methyl 2-carboxy-6-(2-imidazoleazo)benzoate (1) and the corresponding dimethyl ester (2) was studied kinetically at various pH values. For 2, the ester group located at the o position to the azo substiuent was hydrolyzed. From the rate data obtained at various metal concentrations, the values of kcat and Kf were estimated at each pH value. For the Ni(II)-catalyzed hydrolysis of 1 at pH < 4, kcat increases as pH is lowered, indicating bifunctional catalysis by the carboxyl group and the metal ion. For most of the reactions investigated under other conditions, the ester hydrolysis was subjected to sole catalysis by the metal ions. Detailed analysis of kinetic data obtained for these reactions indicated that the metal-ion catalysis involves the rate-determining breakdown of the tetrahedral intermediates formed by the addition of a water molecule or hydroxide ion. The bifunctional catalysis by the carboxyl group and Ni(II) ion can be considered as a model for carboxypeptidase A. The kinetic data indicate that the bifunctional catalysis proceeds through the nucleophilic attack of the carboxylate ion at the Ni(II)-coordinated carbonyl group.  相似文献   

8.
Oxidative stress markers characterize the neuropathology both of Alzheimer's disease and of amyloid-bearing transgenic mice. The neurotoxicity of amyloid A beta peptides has been linked to peroxide generation in cell cultures by an unknown mechanism. We now show that human A beta directly produces hydrogen peroxide (H2O2) by a mechanism that involves the reduction of metal ions, Fe(III) or Cu(II), setting up conditions for Fenton-type chemistry. Spectrophotometric experiments establish that the A beta peptide reduces Fe(III) and Cu(II) to Fe(II) and Cu(I), respectively. Spectrochemical techniques are used to show that molecular oxygen is then trapped by A beta and reduced to H2O2 in a reaction that is driven by substoichiometric amounts of Fe(II) or Cu(I). In the presence of Cu(II) or Fe(III), A beta produces a positive thiobarbituric-reactive substance (TBARS) assay, compatible with the generation of the hydroxyl radical (OH.). The amounts of both reduced metal and TBARS reactivity are greatest when generated by A beta 1-42 > A beta 1-40 > rat A beta 1-40, a chemical relationship that correlates with the participation of the native peptides in amyloid pathology. These findings indicate that the accumulation of A beta could be a direct source of oxidative stress in Alzheimer's disease.  相似文献   

9.
Myeloperoxidase (MPO), an abundant protein in neutrophils, monocytes, and subpopulations of tissue macrophages, is believed to play a critical role in host defenses and inflammatory tissue injury. To perform these functions, an array of diffusible radicals and reactive oxidant species may be formed through oxidation reactions catalyzed at the heme center of the enzyme. Myeloperoxidase and inducible nitric-oxide synthase are both stored in and secreted from the primary granules of activated leukocytes, and nitric oxide (nitrogen monoxide; NO) reacts with the iron center of hemeproteins at near diffusion-controlled rates. We now demonstrate that NO modulates the catalytic activity of MPO through distinct mechanisms. NO binds to both ferric (Fe(III), the catalytically active species) and ferrous (Fe(II)) forms of MPO, generating stable low-spin six-coordinate complexes, MPO-Fe(III).NO and MPO-Fe(II).NO, respectively. These nitrosyl complexes were spectrally distinguishable by their Soret absorbance peak and visible spectra. Stopped-flow kinetic analyses indicated that NO binds reversibly to both Fe(III) and Fe(II) forms of MPO through simple one-step mechanisms. The association rate constant for NO binding to MPO-Fe(III) was comparable to that observed with other hemoproteins whose activities are thought to be modulated by NO in vivo. In stark contrast, the association rate constant for NO binding to the reduced form of MPO, MPO-Fe(II), was over an order of magnitude slower. Similarly, a 2-fold decrease was observed in the NO dissociation rate constant of the reduced versus native form of MPO. The lower NO association and dissociation rates observed suggest a remarkable conformational change that alters the affinity and accessibility of NO to the distal heme pocket of the enzyme following heme reduction. Incubation of NO with the active species of MPO (Fe(III) form) influenced peroxidase catalytic activity by dual mechanisms. Low levels of NO enhanced peroxidase activity through an effect on the rate-limiting step in catalysis, reduction of Compound II to the ground-state Fe(III) form. In contrast, higher levels of NO inhibited MPO catalysis through formation of the nitrosyl complex MPO-Fe(III)-NO. NO interaction with MPO may thus serve as a novel mechanism for modulating peroxidase catalytic activity, influencing the regulation of local inflammatory and infectious events in vivo.  相似文献   

10.
In an effort to better understand the structure, metal content, the nature of the metal centers, and enzyme activity of Arabidopsis thaliana Glx2-2, the enzyme was overexpressed, purified, and characterized using metal analyses, kinetics, and UV–vis, EPR, and 1H NMR spectroscopies. Glx2-2-containing fractions that were purple, yellow, or colorless were separated during purification, and the differently colored fractions were found to contain different amounts of Fe and Zn(II). Spectroscopic analyses of the discrete fractions provided evidence for Fe(II), Fe(III), Fe(III)–Zn(II), and antiferromagnetically coupled Fe(II)–Fe(III) centers distributed among the discrete Glx2-2-containing fractions. The individual steady-state kinetic constants varied among the fractionated species, depending on the number and type of metal ion present. Intriguingly, however, the catalytic efficiency constant, k cat/K m, was invariant among the fractions. The value of k cat/K m governs the catalytic rate at low, physiological substrate concentrations. We suggest that the independence of k cat/K m on the precise makeup of the active-site metal center is evolutionarily related to the lack of selectivity for either Fe versus Zn(II) or Fe(II) versus Fe(III), in one or more metal binding sites.  相似文献   

11.
Poplar (Populus jacquemontiana var. glauca cv. Kopeczkii) was grown in hydroponics containing 10 μM Cd(II), Ni(II) or Pb(II), and Fe as Fe(III) EDTA or Fe(III) citrate in identical concentrations. The present study was designed to compare the accumulation and distribution of Fe, Cd, Ni and Pb within the different plant compartments. Generally, Fe and heavy-metal accumulation were higher by factor 2-7 and 1.6-3.3, respectively, when Fe(III) citrate was used. Iron transport towards the shoot depended on the Fe(III) chelate and, generally, on the heavy metal used. Lead was accumulated only in the root. The amounts of Fe and heavy metals accumulated by poplar were very similar to those of cucumber grown in an identical way, indicating strong Fe uptake regulation of these two Strategy I plants: a cultivar and a woody plant. The Strategy I Fe uptake mechanism (i.e. reducing Fe(III) followed by Fe(II) uptake), together with the Fe(III) chelate form in the nutrient solution had significant effects on Fe and heavy metal uptake. Poplar appears to show phytoremediation potential for Cd and Ni, as their transport towards the shoot was characterized by 51-54% and 26-48% depending on the Fe(III) supply in the nutrient solution.  相似文献   

12.
13.
Resonance Raman (RR) spectroscopy is used to examine porphyrin substrate, product, and inhibitor interactions with the active site of murine ferrochelatase (EC 4.99.1.1), the terminal enzyme in the biosynthesis of heme. The enzyme catalyzes in vivo Fe(2+) chelation into protoporphyrin IX to give heme. The RR spectra of native ferrochelatase show that the protein, as isolated, contains varying amounts of endogenously bound high- or low-spin ferric heme, always at much less than 1 equiv. RR data on the binding of free-base protoporphyrin IX and its metalated complexes (Fe(III), Fe(II), and Ni(II)) to active wild-type protein were obtained at varying ratios of porphyrin to protein. The binding of ferric heme, a known inhibitor of the enzyme, leads to the formation of a low-spin six-coordinate adduct. Ferrous heme, the enzyme's natural product, binds in the ferrous high-spin five-coordinate state. Ni(II) protoporphyrin, a metalloporphyrin that has a low tendency toward axial ligation, becomes distorted when bound to ferrochelatase. Similarly for free-base protoporphyrin, the natural substrate of ferrochelatase, the RR spectra of porphyrin-protein complexes reveal a saddling distortion of the porphyrin. These results corroborate and extend our previous findings that porphyrin distortion, a crucial step of the catalytic mechanism, occurs even in the absence of bound metal substrate. Moreover, RR data reveal the presence of an amino acid residue in the active site of ferrochelatase which is capable of specific axial ligation to metals.  相似文献   

14.
The glycerophosphodiesterase from Enterobacter aerogenes (GpdQ) is a highly promiscuous dinuclear metallohydrolase with respect to both substrate specificity and metal ion composition. While this promiscuity may adversely affect the enzyme's catalytic efficiency its ability to hydrolyse some organophosphates (OPs) and by-products of OP degradation have turned GpdQ into a promising candidate for bioremedial applications. Here, we investigated both metal ion binding and the effect of the metal ion composition on catalysis. The prevalent in vivo metal ion composition for GpdQ is proposed to be of the type Fe(II)Zn(II), a reflection of natural abundance rather than catalytic optimisation. The Fe(II) appears to have lower binding affinity than other divalent metal ions, and the catalytic efficiency of this mixed metal center is considerably smaller than that of Mn(II), Co(II) or Cd(II)-containing derivatives of GpdQ. Interestingly, metal ion replacements do not only affect catalytic efficiency but also the optimal pH range for the reaction, suggesting that different metal ion combinations may employ different mechanistic strategies. These metal ion-triggered modulations are likely to be mediated via an extensive hydrogen bond network that links the two metal ion binding sites via residues in the substrate binding pocket. The observed functional diversity may be the cause for the modest catalytic efficiency of wild-type GpdQ but may also be essential to enable the enzyme to evolve rapidly to alter substrate specificity and enhance kcat values, as has recently been demonstrated in a directed evolution experiment. This article is part of a Special Issue entitled: Chemistry and mechanism of phosphatases, diesterases and triesterases.  相似文献   

15.
Magnesium is essential for the catalysis reaction of Escherichia coli primase, the enzyme synthesizing primer RNA chains for initiation of DNA replication. To map the Mg(2+) binding site in the catalytic center of primase, we have employed the iron cleavage method in which the native bound Mg(2+) ions were replaced with Fe(2+) ions and the protein was then cleaved in the vicinity of the metal binding site by adding DTT which generated free hydroxyl radicals from the bound iron. Three Fe(2+) cleavages were generated at sites designated I, II, and III. Adding Mg(2+) or Mn(2+) ions to the reaction strongly inhibited Fe(2+) cleavage; however, adding Ca(2+) or Ba(2+) ions had much less effect. Mapping by chemical cleavage and subsequent site-directed mutagensis demonstrated that three acidic residues, Asp345 and Asp347 of a conserved DPD sequence and Asp269 of a conserved EGYMD sequence, were the amino acid residues that chelated Mg(2+) ions in the catalytic center of primase. Cleavage data suggested that binding to D345 is significantly stronger than to D347 and somewhat stronger than to D269.  相似文献   

16.
Ruzzini AC  Horsman GP  Eltis LD 《Biochemistry》2012,51(29):5831-5840
meta-Cleavage product (MCP) hydrolases catalyze C-C bond fission in the aerobic catabolism of aromatic compounds by bacteria. These enzymes utilize a Ser-His-Asp triad to catalyze hydrolysis via an acyl-enzyme intermediate. BphD, which catalyzes the hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) in biphenyl degradation, catalyzed the hydrolysis of an ester analogue, p-nitrophenyl benzoate (pNPB), with a k(cat) value (6.3 ± 0.5 s(-1)) similar to that of HOPDA (6.5 ± 0.5 s(-1)). Consistent with the breakdown of a shared intermediate, product analyses revealed that BphD catalyzed the methanolysis of both HOPDA and pNPB, partitioning the products to benzoic acid and methyl benzoate in similar ratios. Turnover of HOPDA was accelerated up to 4-fold in the presence of short, primary alcohols (methanol > ethanol > n-propanol), suggesting that deacylation is rate-limiting during catalysis. In the steady-state hydrolysis of HOPDA, k(cat)/K(m) values were independent of methanol concentration, while both k(cat) and K(m) values increased with methanol concentration. This result was consistent with a simple model of nucleophilic catalysis. Although the enzyme could not be saturated with pNPB at methanol concentrations of >250 mM, k(obs) values from the steady-state turnover of pNPB at low methanol concentrations were also consistent with a nucleophilic mechanism of catalysis. Finally, transient-state kinetic analysis of pNPB hydrolysis by BphD variants established that substitution of the catalytic His reduced the rate of acylation by more than 3 orders of magnitude. This suggests that for pNPB hydrolysis, the serine nucleophile is activated by the His-Asp dyad. In contrast, rapid acylation of the H265Q variant during C-C bond cleavage suggests that the serinate forms via a substrate-assisted mechanism. Overall, the data indicate that ester hydrolysis proceeds via the same acyl-enzyme intermediate as that of the physiological substrate but that the serine nucleophile is activated via a different mechanism.  相似文献   

17.
Human NUDT5 (hNUDT5) is an ADP-ribose (ADPR) pyrophosphatase (ADPRase) that plays important roles in controlling the intracellular levels of ADPR and preventing non-enzymatic ADP-ribosylation of proteins by hydrolyzing ADPR to AMP and ribose 5′-phosphate. We report the crystal structure of hNUDT5 in complex with a non-hydrolyzable ADPR analogue, α,β-methyleneadenosine diphosphoribose, and three Mg2 + ions representing the transition state of the enzyme during catalysis. Analysis of this structure and comparison with previously reported hNUDT5 structures identify key residues involved in substrate binding and catalysis. In the transition-state structure, three metal ions are bound at the active site and are coordinated by surrounding residues and water molecules. A conserved water molecule is at an ideal position for nucleophilic attack on the α-phosphate of ADPR. The side chain of Glu166 on loop L9 changes its conformation to interact with the conserved water molecule compared with that in the substrate-bound structure and appears to function as a catalytic base. Mutagenesis and kinetic studies show that Trp28 and Trp46 are important for the substrate binding; Arg51 is involved in both the substrate binding and the catalysis; and Glu112 and Glu116 of the Nudix motif, Glu166 on loop L9, and Arg111 are critical for the catalysis. The structural and biochemical data together reveal the molecular basis of the catalytic mechanism of ADPR hydrolysis by hNUDT5. Specifically, Glu166 functions as a catalytic base to deprotonate a conserved water molecule that acts as a nucleophile to attack the α-phosphate of ADPR, and three Mg2 + ions are involved in the activation of the nucleophile and the binding of the substrate. Structural comparison of different ADPRases also suggests that most dimeric ADPRases may share a similar catalytic mechanism of ADPR hydrolysis.  相似文献   

18.
Hall RS  Xiang DF  Xu C  Raushel FM 《Biochemistry》2007,46(27):7942-7952
NagA is a member of the amidohydrolase superfamily and catalyzes the deacetylation of N-acetyl-d-glucosamine-6-phosphate. The catalytic mechanism of this enzyme was addressed by the characterization of the catalytic properties of metal-substituted derivatives of NagA from Escherichia coli with a variety of substrate analogues. The reaction mechanism is of interest since NagA from bacterial sources is found with either one or two divalent metal ions in the active site. This observation indicates that there has been a divergence in the evolution of NagA and suggests that there are fundamental differences in the mechanistic details for substrate activation and hydrolysis. NagA from E. coli was inactivated by the removal of the zinc bound to the active site and the apoenzyme reactivated upon incubation with 1 equiv of Zn2+, Cd2+, Co2+, Mn2+, Ni2+, or Fe2+. In the proposed catalytic mechanism the reaction is initiated by the polarization of the carbonyl group of the substrate via a direct interaction with the divalent metal ion and His-143. The invariant aspartate (Asp-273) found at the end of beta-strand 8 in all members of the amidohydrolase superfamily abstracts a proton from the metal-bound water molecule (or hydroxide) to promote the hydrolytic attack on the carbonyl group of the substrate. A tetrahedral intermediate is formed and then collapses with cleavage of the C-N bond after proton transfer to the leaving group amine by Asp-273. The lack of a solvent isotope effect by D2O and the absence of any changes to the kinetic constants with increases in solvent viscosity indicate that net product formation is not limited to any significant extent by proton-transfer steps or the release of products. N-Trifluoroacetyl-d-glucosamine-6-phosphate is hydrolyzed by NagA 26-fold faster than the corresponding N-acetyl derivative. This result is consistent with the formation or collapse of the tetrahedral intermediate as the rate limiting step in the catalytic mechanism of NagA.  相似文献   

19.
Divalent metal ions play a crucial role in catalysis by many RNA and protein enzymes that carry out phosphoryl transfer reactions, and defining their interactions with substrates is critical for understanding the mechanism of biological phosphoryl transfer. Although a vast amount of structural work has identified metal ions bound at the active site of many phosphoryl transfer enzymes, the number of functional metal ions and the full complement of their catalytic interactions remain to be defined for any RNA or protein enzyme. Previously, thiophilic metal ion rescue and quantitative functional analyses identified the interactions of three active site metal ions with the 3'- and 2'-substrate atoms of the Tetrahymena group I ribozyme. We have now extended these approaches to probe the metal ion interactions with the nonbridging pro-S(P) oxygen of the reactive phosphoryl group. The results of this study combined with previous mechanistic work provide evidence for a novel assembly of catalytic interactions involving three active site metal ions. One metal ion coordinates the 3'-departing oxygen of the oligonucleotide substrate and the pro-S(P) oxygen of the reactive phosphoryl group; another metal ion coordinates the attacking 3'-oxygen of the guanosine nucleophile; a third metal ion bridges the 2'-hydroxyl of guanosine and the pro-S(P) oxygen of the reactive phosphoryl group. These results for the first time define a complete set of catalytic metal ion/substrate interactions for an RNA or protein enzyme catalyzing phosphoryl transfer.  相似文献   

20.
《Bioorganic chemistry》1987,15(2):167-182
The kinetics of the Ni(II)-catalyzed ester hydrolysis of O-acetyl-2-pyridine-carboxaldoxime, O-acetyl-2-acetylpyridineketoxime, and O-acetyl-6-carboxy-2-pyridine-carboxaldoxime are measured and the values of various kcat parameters are calculated for reaction paths involving one metal ion (kcatW and kcatOH) and two metal ions kcatA and kcatB). Examination of the kinetic data reveals that the kcatW and kcatOH paths for the Ni(II)-catalyzed reactions involve the same mechanism as those for the previously reported Cu(II)-catalyzed reactions. For the kcatA and kcatB paths, the mechanism involving binuclear Ni(II) ions is preferred by analogy with the previously reported Zn(II)-catalyzed reactions. Comparison of kcatOH values for the Cu(II)- and Ni(II)-catalyzed hydrolysis of 1–3 indicates that markedly different steric effects are exerted by the substituents of 2 and 3 on the catalytic behavior of the two metal ions. This is explained in terms of differences in the fit of the metal ions in the metal complexes of 1–3. Present results demonstrate that slight changes in the geometry around the central metal atom can affect the catalytic outcome significantly. The implications of the present results on metal substitution in metalloenzymes are also discussed.  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