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 15 毫秒
1.
Two arsenite-inhibited forms of each of the aldehyde oxidoreductases from Desulfovibrio gigas and Desulfovibrio desulfuricans have been studied by X-ray crystallography and electron paramagnetic resonance (EPR) spectroscopy. The molybdenum site of these enzymes shows a distorted square-pyramidal geometry in which two ligands, a hydroxyl/water molecule (the catalytic labile site) and a sulfido ligand, have been shown to be essential for catalysis. Arsenite addition to active as-prepared enzyme or to a reduced desulfo form yields two different species called A and B, respectively, which show different Mo(V) EPR signals. Both EPR signals show strong hyperfine and quadrupolar couplings with an arsenic nucleus, which suggests that arsenic interacts with molybdenum through an equatorial ligand. X-ray data of single crystals prepared from EPR-active samples show in both inhibited forms that the arsenic atom interacts with the molybdenum ion through an oxygen atom at the catalytic labile site and that the sulfido ligand is no longer present. EPR and X-ray data indicate that the main difference between both species is an equatorial ligand to molybdenum which was determined to be an oxo ligand in species A and a hydroxyl/water ligand in species B. The conclusion that the sulfido ligand is not essential to determine the EPR properties in both Mo–As complexes is achieved through EPR measurements on a substantial number of randomly oriented chemically reduced crystals immediately followed by X-ray studies on one of those crystals. EPR saturation studies show that the electron transfer pathway, which is essential for catalysis, is not modified upon inhibition. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.  相似文献   

2.
 The molybdenum-cofactor (Mo-co) consists of a mononuclear molybdenum or tungsten ion coordinated by one or two molybdopterin ligands. Crystallographic analyses have demonstrated that the molybdopterin ligands are tricyclic and nonplanar, and that they coordinate the metal through their dithiolene sulfurs. Additional ligands to the metal may be provided by amino acid side chains (including serine, cysteine and selenocysteine), as well as one or more nonprotein O or S ligands, such as oxo, hydroxo, and sulfido. The molybdopterin ligand may participate in the various electron transfer reactions associated with the catalytic mechanism of these proteins, as suggested by both oxidation state-dependent changes in the metal coordination environment and the molybdopterin structure, and by the interaction of the molybdopterin with other redox groups within Mo-co-containing enzymes. Received: 14 May 1997 / Accepted: 20 August 1997  相似文献   

3.
Mononuclear Mo-containing enzymes of the xanthine oxidase (XO) family catalyze the oxidative hydroxylation of aldehydes and heterocyclic compounds. The molybdenum active site shows a distorted square-pyramidal geometry in which two ligands, a hydroxyl/water molecule (the catalytic labile site) and a sulfido ligand, have been shown to be essential for catalysis. The XO family member aldehyde oxidoreductase from Desulfovibrio gigas (DgAOR) is an exception as presents in its catalytically competent form an equatorial oxo ligand instead of the sulfido ligand. Despite this structural difference, inactive samples of DgAOR can be activated upon incubation with dithionite plus sulfide, a procedure similar to that used for activation of desulfo-XO. The fact that DgAOR does not need a sulfido ligand for catalysis indicates that the process leading to the activation of inactive DgAOR samples is different to that of desulfo-XO. We now report a combined kinetic and X-ray crystallographic study to unveil the enzyme modification responsible for the inactivation and the chemistry that occurs at the Mo site when DgAOR is activated. In contrast to XO, which is activated by resulfuration of the Mo site, DgAOR activation/inactivation is governed by the oxidation state of the dithiolene moiety of the pyranopterin cofactor, which demonstrates the non-innocent behavior of the pyranopterin in enzyme activity. We also showed that DgAOR incubation with dithionite plus sulfide in the presence of dioxygen produces hydrogen peroxide not associated with the enzyme activation. The peroxide molecule coordinates to molybdenum in a η2 fashion inhibiting the enzyme activity.  相似文献   

4.
 The nature of the catalytic centres of the oxomolybdoenzymes is considered with particular reference to the results of recent protein crystallographic studies. The different nature of these centres, with one or two molecules of a special pyranopterin (molybdopterin) ligating the metal through a dithiolene group, the presence or absence of a nucleotide appended to the phosphate of the molybdopterin AND the variation in the coordination chemistry at the metal render the term "THE molybdenum cofactor" meaningless and confusing. Rather, there is a series of such cofactors, related by the common denominators of a single molybdenum atom bound to the dithiolene group of the molybdopterin and, at some stage in the catalytic cycle, at least one terminal oxo group. This Mo(O)(molybdopterin) moiety is considered to be the metal-centred functional unit (McFU) of the oxomolybdoenzymes. Variations in the coordination chemistry and, therefore, the properties of the metal centre occur with the binding of other ligands, which can include: a terminal oxo or sulfido group, OH and/or H2O group(s), a second pterin, and/or a serine, a cysteine or selenocysteine group from the polypeptide backbone of the protein. The role of molybdopterin is considered with particular reference to its potential involvement in the various redox processes necessary for the operation of the catalytic cycles of these enzymes; special attention is given to the possible cooperativity between formally metal-based and pterin-based redox processes. Received: 17 June 1997 / Accepted: 20 August 1997  相似文献   

5.
Re-evaluation of the crystallographic data of the molybdenum-containing E. coli formate dehydrogenase H (Boyington et al. Science 275:1305–1308, 1997), reported in two redox states, reveals important structural differences for the formate-reduced form, with large implications for the reaction mechanism proposed in that work. We have re-refined the reduced structure with revised protocols and found substantial rearrangement in some parts of it. The original model is essentially correct but an important loop close to the molybdenum active site was mistraced, and, therefore, catalytic relevant residues were located in wrong positions. In particular selenocysteine-140, a ligand of molybdenum in the original work, and essential for catalysis, is no longer bound to the metal after reduction of the enzyme with formate. These results are incompatible with the originally proposed reaction mechanism. On the basis of our new interpretation, we have revised and proposed a new reaction mechanism, which reconciles the new X-ray model with previous biochemical and extended X-ray absorption fine structure data.  相似文献   

6.
Despite significant influence of secondary bile acids on human health and disease, limited structural and biochemical information is available for the key gut microbial enzymes catalyzing its synthesis. Herein, we report apo‐ and cofactor bound crystal structures of BaiA2, a short chain dehydrogenase/reductase from Clostridium scindens VPI 12708 that represent the first protein structure of this pathway. The structures elucidated the basis of cofactor specificity and mechanism of proton relay. A conformational restriction involving Glu42 located in the cofactor binding site seems crucial in determining cofactor specificity. Limited flexibility of Glu42 results in imminent steric and electrostatic hindrance with 2′‐phosphate group of NADP(H). Consistent with crystal structures, steady state kinetic characterization performed with both BaiA2 and BaiA1, a close homolog with 92% sequence identity, revealed specificity constant (kcat/KM) of NADP+ at least an order of magnitude lower than NAD+. Substitution of Glu42 with Ala improved specificity toward NADP+ by 10‐fold compared to wild type. The cofactor bound structure uncovered a novel nicotinamide‐hydroxyl ion (NAD+‐OH?) adduct contraposing previously reported adducts. The OH? of the adduct in BaiA2 is distal to C4 atom of nicotinamide and proximal to 2′‐hydroxyl group of the ribose moiety. Moreover, it is located at intermediary distances between terminal functional groups of active site residues Tyr157 (2.7 Å) and Lys161 (4.5 Å). Based on these observations, we propose an involvement of NAD+‐OH? adduct in proton relay instead of hydride transfer as noted for previous adducts. Proteins 2014; 82:216–229. © 2013 Wiley Periodicals, Inc.  相似文献   

7.
Possible reaction intermediates of the histidine ammonia-lyase (HAL) reaction were investigated within the tightly closed active site of HAL from Pseudomonas putida (PpHAL). The closed structure of PpHAL was derived from the crystal structure of PpHAL inhibited with l-cysteine, in which the 39–80 loop including the catalytically essential Tyr53 was replaced. This modified loop with closed conformation was modeled using the structure of phenylalanine ammonia-lyase from Anabaena variabilis (AvPAL) with a tightly closed active site as a template. Three hypothetical structures of the covalently bound intermediate in the PpHAL active site were investigated by conformational analysis. The distances between the acidic pro-S β-hydrogen of the ligand and the appropriate oxygen atoms of Tyr53, Ty280 and Glu414 − which may act as enzymic bases − in the conformations of the three hypothetical intermediate structures were analyzed together with the substrate and product arrangements. The calculations indicated that the most plausible HAL reaction pathway involves the N-MIO intermediate structure in which the L-histidine substrate is covalently bound to the N-3,5-dihydro-5-methylidene-4H-imidazol-4-one (MIO) prosthetic group of the apoenzyme via the amino group. Density functional theory (DFT) calculations − on a truncated model of the N-MIO intermediate containing a Zn2+ ion coordinated to the imidazole ring of the ligand and to His83, Met382 and a water molecule − indicated that Zn-complex formation plays a role in the reactivity and substrate specificity of HAL.  相似文献   

8.
The active site of ß‐galactosidase (E. coli) contains a Mg2+ ion ligated by Glu‐416, His‐418 and Glu‐461 plus three water molecules. A Na+ ion binds nearby. To better understand the role of the active site Mg2+ and its ligands, His‐418 was substituted with Asn, Glu and Phe. The Asn‐418 and Glu‐418 variants could be crystallized and the structures were shown to be very similar to native enzyme. The Glu‐418 variant showed increased mobility of some residues in the active site, which explains why the substitutions at the Mg2+ site also reduce Na+ binding affinity. The Phe variant had reduced stability, bound Mg2+ weakly and could not be crystallized. All three variants have low catalytic activity due to large decreases in the degalactosylation rate. Large decreases in substrate binding affinity were also observed but transition state analogs bound as well or better than to native. The results indicate that His‐418, together with the Mg2+, modulate the central role of Glu‐461 in binding and as a general acid/base catalyst in the overall catalytic mechanism. Glucose binding as an acceptor was also dramatically decreased, indicating that His‐418 is very important for the formation of allolactose (the natural inducer of the lac operon).  相似文献   

9.
Saccharomyces cerevisiae phosphoenolpyruvate (PEP) carboxykinase catalyzes the reversible formation of oxaloacetate and adenosine triphosphate from PEP, adenosine diphosphate and carbon dioxide, and uses Mn2+ as the activating metal ion. Comparison with the crystalline structure of homologous Escherichia coli PEP carboxykinase [Tari et al. (1997) Nature Struct. Biol. 4, 990–994] shows that Lys213 is one of the ligands to Mn2+ at the enzyme active site. Coordination of Mn2+ to a lysyl residue is not common and suggests a low pK a value for the ε-NH2 group of Lys213. In this work, we evaluate the role of neighboring Phe216 in contributing to provide a low polarity microenvironment suitable to keep the ε-NH2 of Lys213 in the unprotonated form. Mutation Phe216Tyr shows that the introduction of a hydroxyl group in the lateral chain of the residue produces a substantial loss in the enzyme affinity for Mn2+, suggesting an increase of the pK a of Lys213. In agreement with this interpretation, theoretical calculations indicate an alkaline shift of 2.8 pH units in the pK a of the ε-amino group of Lys213 upon Phe216Tyr mutation.  相似文献   

10.
[Mo(SSCH3)(S2C2(CH3)2)2] x complexes with charges x between −3 and +3 were investigated by density functional theory computations as minimal nitrate reductase active-site models. The strongly reduced species (x = −2, −3) exist preferentially as pentacoordinate sulfo complexes separated from a thiolate anion. The oxidized extremes (x > 0) clearly prefer hexacoordinate complexes with an η2-MeSS ligand. Among the neutral and especially for the singly negatively charged species structures with η2-MeSS and η1-MeSS ligands are energetically close to the sulfo methyl sulfide complex without SS bonding. For x = −1 the three isomers lie in a 1.5 kcal mol−1 energy range. Putative mechanistic pathways for nitrate reduction from the literature were investigated computationally: (1) reduction at a pentacoordinate sulfo complex, (2) reduction at the ligand, and (3) reduction at the molybdenum center with an R–S–S ligand. All three pathways could be traced at least for some overall charges but no definite conclusion can be drawn about the mechanism. Complexes with larger dithiolato ligands were also computed in order to model the tricyclic metallopterin framework more accurately: the first heterocyclus (5,6-dihydro-2H-pyran) stabilizes the nitrate complex and the molybdenum oxo product complex by approximately 10 kcal mol−1 and also reduces the activation barrier (by approximately 5 kcal mol−1). The effect of the second (1,2,3,4-tetrahydropyrazin) and third heterocyclus (2-amino-3H-pyrimidin-4-one) on the relative energies is relatively small. For bigger models derived from an experimental protein structure, nitrate reduction at a persulfo molybdenum(IV) complex fragment (mechanism 3) is clearly favored over the oxidation of a molybdenum-bound sulfur atom (mechanism 2). Mechanism 1 could not be investigated for the big models but seems the least favorable on the basis of the results from smaller models.  相似文献   

11.
 A two-step concerted mechanism for the conversion of methane to methanol catalyzed by soluble methane monooxygenase (sMMO) is discussed. We propose that the enzymatic reaction mechanism is essentially the same as that of the gas-phase methane-methanol conversion by the bare FeO+ complex. In the initial stage of our mechanism, the ferryl (Fe—O) "iron" active site of intermediate Q and substrate methane come into contact to form the initial Q (CH4) complex with an OFe—CH4 bond. The C—H bonds of methane are significantly weakened by the formation of a five-coordinate carbon species, through orbital interactions between a C 3v - or D 2d -distorted methane and the Fe—O active site. The important transition state for an H atom abstraction exhibits a four-centered structure. The generated intermediate involves an HO—Fe—CH3 moiety, and it is then converted into the final product complex including methanol as a ligand through a methyl migration that occurs via a three-centered transition state. The two-step concerted mechanism is consistent with recent experiments on regioselectivity of enzyme-catalyzed alkane hydroxylations. Received: 15 September 1997 / Accepted: 20 December 1997  相似文献   

12.
L Lebioda  B Stec 《Biochemistry》1991,30(11):2817-2822
Enolase in the presence of Mg2+ catalyzes the elimination of H2O from 2-phosphoglyceric acid (PGA) to form phosphoenolpyruvate (PEP) and the reverse reaction, the hydration of PEP to PGA. The structure of the ternary complex yeast enolase-Mg2(+)-PGA/PEP has been determined by X-ray diffraction and refined by crystallographic restrained least-squares to an R = 16.9% for those data with I/sigma (I) greater than or equal to 2 to 2.2-A resolution with a good geometry of the model. The structure indicates the substrate molecule in the active site has its hydroxyl group coordinated to the Mg2+ ion. The carboxylic group interacts with the side chains of His373 and Lys396. The phosphate group is H-bonded to the guanidinium group of Arg374. A water molecule H-bonded to the carboxylic groups of Glu168 and Glu211 is located at a 2.6-A distance from carbon-2 of the substrate in the direction of its proton. We propose that this cluster functions as the base abstracting the proton in the catalytic process. The proton is probably transferred, first to the water molecule, then to Glu168, and further to the substrate hydroxyl to form a water molecule. Some analogy is apparent between the initial stages of the enolase reverse reaction, the hydration of PEP, and the proteolytic mechanism of the metallohydrolases carboxypeptidase A and thermolysin. The substrate/product binding is accompanied by large movements of loops Ser36-His43 and Ser158-Gly162. The role of these conformational changes is not clear at this time.  相似文献   

13.
This mini-review briefly summarizes our current knowledge on the reaction pattern of light-driven water splitting and the structure of Photosystem II that acts as a water:plastoquinone oxidoreductase. The overall process comprises three types of reaction sequences: (a) light-induced charge separation leading to formation of the radical ion pair P680+•QA−•; (b) reduction of plastoquinone to plastoquinol at the QB site via a two-step reaction sequence with QA−• as reductant and (c) oxidative water splitting into O2 and four protons at a manganese-containing catalytic site via a four-step sequence driven by P680+• as oxidant and a redox active tyrosine YZ acting as mediator. Based on recent progress in X-ray diffraction crystallographic structure analysis the array of the cofactors within the protein matrix is discussed in relation to the functional pattern. Special emphasis is paid on the structure of the catalytic sites of PQH2 formation (QB-site) and oxidative water splitting (Mn4O x Ca cluster). The energetics and kinetics of the reactions taking place at these sites are presented only in a very concise manner with reference to recent up-to-date reviews. It is illustrated that several questions on the mechanism of oxidative water splitting and the structure of the catalytic sites are far from being satisfactorily answered.  相似文献   

14.
Conversion of iron(II) verdoheme to iron(II) biliverdin in the presence of hydroxyl ion as a nucleophile and imidazole, pyridine, water, hydroxyl, cyanide, phenolate, chloride, thiolate and imidazolate as axial ligands was investigated using the B3LYP method and the 6-31G basis set. In the five-coordinated pathway the reactants and products are in the ground triplet state. In this path, hydroxyl ion directly attacks the macrocycle. The exothermic energy for addition of hydroxyl ion to iron(II) verdoheme with various ligands is 169.55, 166.34 and 164 kcal mol−1 for water, pyridine and imidazole, energies which are around 30–60 kcal mol−1 more exothermic than those for the other axial ligands used in this study. Therefore, imidazole, water and pyridine axial ligands can facilitate hydrolytic cleavage of iron(II) verdoheme to form open-chained helical iron(II) biliverdin complexes. The activation barrier for the conversion of iron(II) verdoheme hydroxyl species to the iron(II) biliverdin complex is estimated to be 5.2, 4.2, 4.35, 13.76 and 14.05 kcal mol−1 for imidazole, water, cyanide, thiolate and imidazolate, respectively. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

15.
The non-functional form of xanthine oxidase known as the desulpho enzyme was compared with the functional enzyme in various ways, to obtain information on the structure of the molybdenum centre and the mechanism of the catalytic reaction. The desulpho enzyme, like the functional one, possesses a site for the binding of anions, presumably as ligands of molybdenum. Evidence is presented that in the Mo(V) e.p.r. signal from the desulpho-enzyme, as in that from the functional enzyme, a weakly coupled proton, in addition to a strongly coupled proton, interacts with the metal. Measurements were carried out by e.p.r. on the rate at which the proton strongly coupled to molybdenum exchanged, on diluting enzyme samples with 2H2O. For the desulpho enzyme the exchange rate constant was 0.40s-1, at pH 8.2 and 12 degrees C, and for the functional enzyme it was 85 s-1. It is shown that the great majority of reported differences between the enzyme forms are consistent with functional enzyme containing an (Enzyme)-Mo=S grouping, replaced in the desulpho form by (Enzyme)-Mo=O. Protonation of these groups, with pK values of about 8 and 10 respectively, would give (Enzyme)-Mo-SH and (Enzyme)-Mo-OH, these being the forms observed by e.p.r. The accepting group in the functional enzyme, for the proton transferred from the substrate while molybdenum is reduced in the catalytic reaction [Gutteridge, Tanner & Bray (1978) Biochem J. 175 869-878], is thus taken to be Mo=S.  相似文献   

16.
Contact-dependent growth inhibition is a mechanism of interbacterial competition mediated by delivery of the C-terminal toxin domain of CdiA protein (CdiA–CT) into neighboring bacteria. The CdiA–CT of enterohemorrhagic Escherichia coli EC869 (CdiA–CTEC869) cleaves the 3′-acceptor regions of specific tRNAs in a reaction that requires the translation factors Tu/Ts and GTP. Here, we show that CdiA–CTEC869 has an intrinsic ability to recognize a specific sequence in substrate tRNAs, and Tu:Ts complex promotes tRNA cleavage by CdiA–CTEC869. Uncharged and aminoacylated tRNAs (aa-tRNAs) were cleaved by CdiA–CTEC869 to the same extent in the presence of Tu/Ts, and the CdiA–CTEC869:Tu:Ts:tRNA(aa-tRNA) complex formed in the presence of GTP. CdiA–CTEC869 interacts with domain II of Tu, thereby preventing the 3′-moiety of tRNA to bind to Tu as in canonical Tu:GTP:aa-tRNA complexes. Superimposition of the Tu:GTP:aa-tRNA structure onto the CdiA–CTEC869:Tu structure suggests that the 3′-portion of tRNA relocates into the CdiA–CTEC869 active site, located on the opposite side to the CdiA–CTEC869 :Tu interface, for tRNA cleavage. Thus, CdiA–CTEC869 is recruited to Tu:GTP:Ts, and CdiA–CT:Tu:GTP:Ts recognizes substrate tRNAs and cleaves them. Tu:GTP:Ts serves as a reaction scaffold that increases the affinity of CdiA–CTEC869 for substrate tRNAs and induces a structural change of tRNAs for efficient cleavage by CdiA–CTEC869.  相似文献   

17.
The enzyme 2′-aminobiphenyl-2,3-diol-1,2-dioxygenase (CarB), encoded by two genes (carBa and carBb), is an α2β2 heterotetramer that presents meta-cleavage activity toward the hydroxylated aromatic ring in the carbazole degradation pathway from petroleum-degrader bacteria Pseudomonas spp. The 1082-base, pair polymerase chain reaction product corresponding to, carBaBb genes from Pseudomonas stutzeri ATCC 31258 was cloned by site-specific recombination and expressed in high levels in Escherichia coli BL21-SI with a histidine-tag and in native form. The CarB activity toward 2,3-dihydroxybiphenyl was similar for these two constructions. The α2β2 3D model of CarB dioxygenase was proposed by homology modeling using the protocatechuate 4,5-dioxygenase (LigAB) structure as template. Accordingly, His12, His53, and Glu230 coordinate the Fe(II) in the catalytic site at the subunit CarBb. The model also indicates that His182 is the catalytic base responsible for deprotonating one of the hydroxyl group of the substrate by a hydrogen bond. The hydrophobic residues Trp257 and Phe258 in the CarB structure substituted the LigAB amino acid residues Ser269 and Asn270. These data could explain why the CarB was active for 2,3-dihydroxybiphenyl and not for protocatechuate.  相似文献   

18.
 The role of the cysteine thiolate ligand for the unusual copper coordination geometry in the blue copper proteins has been studied by comparing the electronic structure, geometry, and energetics of a number of small Cu(II) complexes. The geometries have been optimised with the density functional B3LYP method, and energies have been calculated by multiconfigurational second-order perturbation theory (the CASPT2 method). Most small inorganic Cu(II) complexes assume a tetragonal geometry, where four ligands make σ bonds to a Cu 3d orbital. If a ligand lone-pair orbital instead forms a π bond to the copper ion, it formally occupies two ligand positions in a square coordination, and the structure becomes trigonal. Large, soft, and polarisable ligands, such as SH and SeH, give rise to covalent copper-ligand bonds and structures close to a tetrahedron, which might be trigonal or tetragonal with approximately the same stability. On the other hand, small and hard ligands, such as NH3, OH2, and OH, give ionic bonds and flattened tetragonal structures. It is shown that axial type 1 (blue) copper proteins have a trigonal structure with a π bond to the cysteine sulphur atom, whereas rhombic type 1 and type 2 proteins have a tetragonal structure with σ bonds to all strong ligands. The soft cysteine ligand is essential for the stabilisation of a structure that is close to a tetrahedron (either trigonal or tetragonal), which ensures a low reorganisation energy during electron transfer. Received: 9 July 1997 / 26 November 1997  相似文献   

19.
Conversion of pyrogallol to phloroglucinol was studied with the molybdenum enzyme transhydroxylase of the strictly anaerobic fermenting bacterium Pelobacter acidigallici. Transhydroxylation experiments in H218O revealed that none of the hydroxyl groups of phloroglucinol was derived from water, confirming the concept that this enzyme transfers a hydroxyl group from the cosubstrate 1,2,3, 5-tetrahydroxybenzene (tetrahydroxybenzene) to the acceptor pyrogallol, and simultaneously regenerates the cosubstrate. This concept requires a reaction which synthesizes the cofactor de novo to maintain a sufficiently high intracellular pool during growth. Some sulfoxides and aromatic N-oxides were found to act as hydroxyl donors to convert pyrogallol to tetrahydroxybenzene. Again, water was not the source of the added hydroxyl groups; the oxides reacted as cosubstrates in a transhydroxylation reaction rather than as true oxidants in a net hydroxylation reaction. No oxidizing agent was found that supported a formation of tetrahydroxybenzene via a net hydroxylation of pyrogallol. However, conversion of pyrogallol to phloroglucinol in the absence of tetrahydroxybenzene was achieved if little pyrogallol and a high amount of enzyme preparation was used which had been pre-exposed to air. Obviously, the enzyme was oxidized by air to form sufficient amounts of tetrahydroxybenzene from pyrogallol to start the reaction. A reaction mechanism is proposed which combines an oxidative hydroxylation with a reductive dehydroxylation via the molybdenum cofactor, and allows the transfer of a hydroxyl group between tetrahydroxybenzene and pyrogallol without involvement of water. With this, the transhydroxylase differs basically from all other hydroxylating molybdenum enzymes which all use water as hydroxyl source.  相似文献   

20.
Modeling of the catalytic mechanism of penicillin acylase, a member of the N-terminal nucleophile hydrolase superfamily, is for the first time conducted at ab initio quantum chemistry level. The uniqueness of this family of enzymes is that their active site lacks His and Asp (Glu) residues, comprising together with a serine residue the classical catalytic triad. The current investigation confirms that the amino group of the N-terminal serine residue in N-terminal hydrolases is capable of activating its own hydroxyl group. Using the MP2/RHF method with the 6−31+G** basis set, stationary points on the potential energy surface of the considered molecular system were located, corresponding to local minima (complexes of reagents, products, intermediate) and to saddle points (transition states). It turned out that the stage of acyl-serine formation proceeds via two transition states; the first one, which separates reagents from the so-called tetrahedral intermediate, has the highest relative energy (30 kcal/mol). In contrast to recently proposed empiric suggestions, we have found that participation of a bridging water molecule in proton shuttling is not necessary for the catalysis. The quantum chemical calculations showed a crucial role of a specific solvation in decreasing the activation barrier of the reaction by approximately 10 kcal/mol. Published in Russion in Biokhimiya, 2007, Vol. 72, No. 5, pp. 615–621.  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