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.
S-Adenosylhomocysteine (AdoHcy) hydrolase catalyzes the reversible hydrolysis of AdoHcy to adenosine (Ado) and homocysteine (Hcy), playing an essential role in modulating the cellular Hcy levels and regulating activities of a host of methyltransferases in eukaryotic cells. This enzyme exists in an open conformation (active site unoccupied) and a closed conformation (active site occupied with substrate or inhibitor) [Turner, M. A., Yang, X., Yin, D., Kuczera, K., Borchardt, R. T., and Howell, P. L. (2000) Cell Biochem. Biophys. 33, 101-125]. To investigate the binding of natural substrates during catalysis, the computational docking program AutoDock (with confirming calculations using CHARMM) was used to predict the binding modes of various substrates or inhibitors with the closed and open forms of AdoHcy hydrolase. The results have revealed that the interaction between a substrate and the open form of the enzyme is nonspecific, whereas the binding of the substrate in the closed form is highly specific with the adenine moiety of a substrate as the main recognition factor. Residues Thr57, Glu59, Glu156, Gln181, Lys186, Asp190, Met351, and His35 are involved in substrate binding, which is consistent with the crystal structure. His55 in the docked model appears to participate in the elimination of water from Ado through the interaction with the 5'-OH group of Ado. In the same reaction, Asp131 removes a proton from the 4' position of the substrate after the oxidation-reduction reaction in the enzyme. To identify the residues that bind the Hcy moiety, AdoHcy was docked to the closed form of AdoHcy hydrolase. The Hcy tail is predicted to interact with His55, Cys79, Asn80, Asp131, Asp134, and Leu344 in a strained conformation, which may lower the reaction barrier and enhance the catalysis rate.  相似文献   

2.
Elrod P  Zhang J  Yang X  Yin D  Hu Y  Borchardt RT  Schowen RL 《Biochemistry》2002,41(25):8134-8142
Residues glutamate 156 (E156), aspartate 190 (D190), asparagine 181 (N181), lysine 186 (K186), and asparagine 191 (N191) in the active site of S-adenosylhomocysteine (AdoHcy) hydrolase have been mutated to alanine (A). AdoHcy hydrolase achieves catalysis of AdoHcy hydrolysis to adenosine (Ado) and homocysteine (Hcy) by means of a redox partial reaction (3'-oxidation of AdoHcy at the beginning and 3'-reduction of Ado at the end of the catalytic cycle) spanning an elimination/addition partial reaction (elimination of Hcy from the oxidized substrate and addition of water to generate the oxidized product), with the enzyme in an open NAD(+) form in the ligand-free state and in a closed NADH form during the elimination/addition partial reaction. Mutation K186A reduces the rate of a model enzymatic reaction for the redox partial reaction by a factor of 280000 and the rate of a model reaction for the elimination/addition partial reaction by a factor of 24000, consistent with a primary catalytic role in both partial reactions as a proton donor/acceptor at the 3'-OH/3'-keto center. Secondary roles for N181 and N191 in localizing the flexible side chain of K186 in a catalytically effective position are supported by rate reduction factors for N181A of 2500 (redox) and 240 (elimination/addition) and for N191A of 730 (redox) and 340 (elimination/addition). A role of D190 in orienting the substrate for effective transition-state stabilization is consistent with rate reduction factors of 1300 (redox) and 30 (elimination/addition) for D190A. Residue E156 may act to maintain K186 in the desired protonation state: rate deduction factors are 1100 (redox) and 70 (elimination/addition). The mutational increases in free energy barriers for k(cat)/K(M) are described by a linear combination of the effects for the partial reactions with the coefficients equal to the fractional degree that each partial reaction determines the rate for k(cat)/K(M). A similar linear equation for k(cat) overestimates the barrier increase by a uniform 5 kJ/mol, probably reflecting reactant-state stabilization by the wild-type enzyme that is abolished by the mutations.  相似文献   

3.
S-adenosylhomocysteine hydrolase (AdoHcyase) catalyzes the hydrolysis of S-adenosylhomocysteine (AdoHcy) to form adenosine and homocysteine. The crystal structure of the K185N mutated enzyme, which has weak catalytic activity (0.1%), has been determined at 2.8 A resolution and supports the previously predicted mechanism [Takata, Y., Yamada, T., Huang, Y., Komoto, J., Gomi, T., Ogawa, H., Fujioka, M., & Takusagawa, F. (2002). Catalytic mechanism of S-adenosylhomocysteine hydrolase. Site-directed mutagenesis of Asp-130, Lys-185, Asp-189, and Asn-190. J. Biol. Chem. 277, 22670-22676]. The mutated enzyme has an intermediate structure between the open and closed conformation, observed in the substrate-free enzyme and in the inhibitor complexes, respectively. H54, H300, and H352 were mutated to asparagine, respectively, to identify the roles of the histidine residues in catalysis. The kinetic data of H54N, H300N, and H354N mutated enzymes suggest that H54 is the amino acid residue that acts as a general acid/base to cleave the C5'-S(D) bond of AdoHcy. The E155Q mutated enzyme retained a large portion of the catalytic activity (31%), while the E155D mutated enzyme lost most of it (0.3%). The NADH accumulation measurements of the mutated enzymes indicated that the C3'-oxidation and the C4'-proton abstraction are a concerted event and the C5'-S(D) bond cleavage is an independent event. The C4'-proton exchange measurements indicate that the enzyme has an open conformation when AdoHcy is converted to 3'-keto-4', 5'-dehydro-Ado in the active site. With the results of this study and those of the previous studies, a detailed catalytic mechanism of AdoHcyase is described. K185 facilitates the C3'-oxidation, D130 abstracts the C4'-proton, D189, and E155 act as a communicator between the concerted C3'-oxidation and C4'-proton abstraction, and H54 plays as a general acid to cleave the C5'-S(D) bond of AdoHcy.  相似文献   

4.
Abstract

Crucial to the rational design of inhibitors of S-adenosyl-L-homocysteine (AdoHcy) hydrolase was the elucidation of its mechanism of catalysis by Palmer and Abeles (J. Biol. Chem. 254, 1217–1226, 1979). This mechanism involves an NAD+-dependent oxidation (oxidative activity) of the 3′-hydroxyl group of AdoHcy followed by elimination of homocysteine (Hcy) to form 4′,5′-didehydro-3′-keto-Ado. Addition of water at the 5′-position (hydrolytic activity) of this tightly bound intermediate followed by an NADH-dependent reduction results in the formation of adenosine (Ado). Many inhibitors of this enzyme have been shown to serve as substrates [e.g., 9-(trans-2-trans-3-dihydroxycyclopent-4-en-1-yl)adenine, DHCeA)] for the oxidative activity of AdoHcy hydrolase, affording the 3′-keto-derivative (e.g., 3′-keto-DHCeA), which is tightly bound to the enzyme, and converting the enzyme from its active form (NAD+) to its inactive form (NADH) (Type I mechanism-based inhibitors; Wolfe and Borchardt, J. Med. Chem. 34, 1521–1530, 1991). More recently, substrates [e.g., (E)-5.,6′-didehydro-6′-deoxy-6′-fluorohomoadenosine, EDDFHA] for the hydrolytic activity of AdoHcy hydrolase have been identified by our laboratories. Identification of hydrolytic substrates affords a new strategy for the design of more potent and more specific inhibitors of AdoHcy hydrolase.  相似文献   

5.
Neplanocin A, a cyclopentenyl analog of adenosine, has been reported by S. Yaginuma, N. Muto, M. Tsujino, Y. Sudate, M. Hayashi, and M. Otari (1981) J. Antibiot. 34, 359-366 to exhibit antibacterial activity against Alcaligenes faecalis. Since neplanocin A (NpcA) is a known inhibitor of eukaryotic S-adenosylhomocysteine (AdoHcy) hydrolase (EC 3.3.1.1) (R. T. Borchardt, B. T. Keller, and U. Patel-Thombre (1984) J. Biol. Chem. 259, 4353-4358), the present study was undertaken to determine the effects of this carbocyclic nucleoside on AdoHcy hydrolase isolated from a prokaryotic source (A. faecalis). AdoHcy hydrolase was purified to homogeneity by affinity chromatography on an AdoHcy-agarose matrix from A. faecalis. Neplanocin A inactivated the purified AdoHcy hydrolase in a time- and concentration-dependent manner and the enzyme activity could not be recovered by dialysis. The inactivation of this bacterial enzyme by neplanocin A is accompanied by a reduction of three of the six enzyme-bound NAD+s to NADHs. These results suggest that the prokaryotic enzyme, like the eukaryotic AdoHcy hydrolase, is susceptible to inhibition by neplanocin A. The mechanism of inactivation in both cases appears to be a Kcat mechanism involving the reduction of the enzyme-bound NAD+ to NADH. The fact that total inhibition of the prokaryotic AdoHcy hydrolase by NpcA results in a reduction of only three of the six enzyme-bound NAD+s to NADHs suggests that the enzyme shows half-site reactivity (i.e., only three of the six subunits are catalytically active).  相似文献   

6.
On the mechanism of the metallo-beta-lactamase from Bacteroides fragilis.   总被引:1,自引:0,他引:1  
Z Wang  W Fast  S J Benkovic 《Biochemistry》1999,38(31):10013-10023
The catalytic mechanism of metallo-beta-lactamase from Bacteroides fragilis, a dinuclear Zn(II)-containing enzyme responsible for multiple antibiotic resistance, has been investigated by using nitrocefin as a substrate. Rapid-scanning and single-wavelength stopped-flow studies revealed the accumulation during turnover of an enzyme-bound intermediate with intense absorbance at 665 nm (epsilon = 30 000 M(-1) cm(-1)). The proposed minimum kinetic mechanism for the B. fragilis metallo-beta-lactamase-catalyzed nitrocefin hydrolysis [Wang, Z., and Benkovic, S. J. (1998) J. Biol. Chem. 273, 22402-22408] was confirmed, and more accurate kinetic parameters were obtained from computer simulations and fitting. The intermediate was shown to be a novel anionic species bound to the enzyme through a Zn-acyl linkage and contains a negatively charged nitrogen leaving group. This is the first time such an intermediate was observed in the catalytic cycle of a Zn(II)-containing hydrolase and is evidence for a unique beta-lactam hydrolysis mechanism in which the amine can leave as an anion; prior protonation is not required. The electrostatic interaction between the negatively charged intermediate and the positively charged dinuclear Zn(II) center of the enzyme is important for stabilization of the intermediate. The catalytic reaction was accelerated in the presence of exogenous nucleophiles or anions, and neither the product nor the enzyme was modified during turnover, indicating that a Zn-bound hydroxide (rather than Asp-103) is the active site nucleophile. On the basis of all the information on hand, a catalytic mechanism of the B. fragilis metallo-beta-lactamase is proposed.  相似文献   

7.
Comparison of crystal structures of S-adenosylhomocysteine (AdoHcy) hydrolase in the substrate-free, NAD(+) form [Hu, Y., Komoto, J., Huang, Y., Gomi, T., Ogawa, H., Takata, Y., Fujioka, M., and Takusagawa, F. (1999) Biochemistry 38, 8323-8333] and a substrate-bound, NADH form [Turner, M. A., Yuan, C.-S., Borchardt, R. T., Hershfield, M. S., Smith, G. D., and Howell, P. L. (1998) Nat. Struct. Biol. 5, 369-376] indicates large differences in the spatial arrangement of the catalytic and NAD(+) binding domains. The substrate-free, NAD(+) form exists in an "open" form with respect to catalytic and NAD(+) binding domains, whereas the substrate-bound, NADH form exists in a closed form with respect to those domains. To address whether domain closure is induced by substrate binding or its subsequent oxidation, we have measured the rotational dynamics of spectroscopic probes covalently bound to Cys(113) and Cys(421) within the catalytic and carboxyl-terminal domains. An independent domain motion is associated with the catalytic domain prior to substrate binding, suggesting the presence of a flexible hinge element between the catalytic and NAD(+) binding domains. Following binding of substrates (i.e., adenosine or neplanocin A) or a nonsubstrate (i.e., 3'-deoxyadenosine), the independent domain motion associated with the catalytic domain is essentially abolished. Likewise, there is a substantial decrease in the average hydrodynamic volume of the protein that is consistent with a reduction in the overall dimensions of the homotetrameric enzyme following substrate binding and oxidation observed in earlier crystallographic studies. Thus, the catalytic and NAD(+) binding domains are stabilized to form a closed active site through interactions with the substrate prior to substrate oxidation.  相似文献   

8.
S-Adenosylhomocysteine (AdoHcy) hydrolase (EC 3.3.1.1) catalyzes the reversible hydrolysis of AdoHcy to adenosine and homocysteine. Neplanocin A, a cyclopentyl analog of adenosine, has previously been shown to act as a tight-binding inhibitor of the purified bovine liver enzyme, binding with a stoichiometry of one molecule per tetramer of enzyme (R.T. Borchardt, B.T. Keller, and U.G. Patel-Thombre, 1984, J. Biol. Chem. 259, 4353-4358). In the current study neplanocin A was also shown to act as a stoichiometric inhibitor of the L929 cell enzyme having Ki = 0.2 nM. Using this inhibitor to titrate the AdoHcy hydrolase, the concentration of the enzyme in intact L929 cells was calculated to be 0.8 microM, assuming a 1:1 inhibitor:protein stoichiometry. It was observed that the specific activity of AdoHcy hydrolase as measured in the hydrolytic direction increased 270% over a 12-h period after L929 cells were given fresh serum-free medium or when the cell extract was dialyzed first against phosphate buffer. Using the neplanocin A titration technique, it was found that the enzyme concentration in L929 cells remained constant over a 48-h period after refeeding the cultures. These results suggest the presence of an endogenous inhibitor or a readily reversible-type enzyme modification which is responsible for regulating AdoHcy hydrolase in vivo.  相似文献   

9.
6-Oxo camphor hydrolase (OCH) is an enzyme of the crotonase superfamily that catalyzes carbon-carbon bond cleavage in bicyclic beta-diketones via a retro-Claisen reaction (Grogan, G., Roberts, G. A., Bougioukou, D., Turner, N. J., and Flitsch, S. L. (2001) J. Biol. Chem. 276, 12565-12572). The native structure of OCH has been solved at 2.0-A resolution with selenomethionine multiple wave anomalous dispersion and refined to a final R(free) of 19.0. The structure of OCH consists of a dimer of trimers that resembles the "parent" enzyme of the superfamily, enoyl-CoA hydratase. In contrast to enoyl-CoA hydratase, however, two octahedrally coordinated sodium atoms are found at the 3-fold axis of the hexamer of OCH, and the C-terminal helix of OCH does not form a discrete domain. Models of the substrate, 6-oxo camphor, and a proposed enolate intermediate in the putative active site suggest possible mechanistic roles for Glu-244, Asp-154, His-122, His-45, and His-145.  相似文献   

10.
Domain motions of S-adenosyl-l-homocysteine (AdoHcy) hydrolase have been detected by time-resolved fluorescence anisotropy measurements. Time constants for reorientational motions in the native enzyme were compared with those for enzymes where key residues were altered by site-directed mutation. Mutations M351P, H353A, and P354A were selected in a hinge region for motion between the open and closed forms of the enzyme, as identified in a previous normal-mode study [Wang et al. (2005) Domain motions and the open-to-closed conformational transition of an enzyme: A normal-mode analysis of S-adenosyl-l-homocysteine hydrolase, Biochemistry 44, 7228-7239]. In wild-type, substrate-free AdoHcy hydrolase (NAD(+) cofactor in each subunit), reorientational motions were detected on time scales of 10-20 and 80-90 ns. The faster motion is attributed to the domain motion, and the slower motion is attributed to the tumbling of the enzyme. The domain motion was also detected for the enzyme complexes E(NADH/3'-keto-adenosine) and E(NAD(+)/3'-deoxyadenosine) but was absent for the complex E(NADH/3'-keto-neplanocin A). The results indicate that AdoHcy hydrolase exists in equilibrium of open and closed structures, with the equilibrium shifted toward the more mobile open form for the substrate-free enzyme, E(NAD(+)), and for intermediates formed early in the catalytic cycle after substrate binding or formed late prior to product release, E(NAD(+)/ligand). However, the strong inhibitor neplanocin A upon binding undergoes oxidation, forming the complex E(NADH/3'-keto-neplanocin). For this complex, which is analogous to the enzyme complex with the central catalytic intermediate, the equilibrium was shifted toward the more rigid closed form. A similar pattern was observed for M351P and P354A mutants. In contrast, the domain motion could not be detected, either in the absence or presence of ligands or with the cofactor in either the oxidized or reduced state, for the H353A protein, suggesting that this mutation changes the hinge-bending dynamics of the enzyme.  相似文献   

11.
Guanosine 5'-O-(3-thio)triphosphate (GTP gamma S) was found to be a substrate of pig heart succinyl-CoA synthetase with Km and kcat values of 3 microM and 0.23 s-1, respectively. The corresponding values with GTP as substrate were 48 microM and 65 s-1. 35S-thiophosphorylated enzyme was prepared by incubation of pig heart succinyl-CoA synthetase with [35S]GTP gamma S. A comparison was made of thiophosphoryl group release by substrates from this alpha beta (one active site) enzyme with that of the alpha 2 beta 2 (two active sites) Escherichia coli enzyme (Wolodko, W. T., Brownie, E. R., O'Connor, M. D., and Bridger, W. A. (1983) J. Biol. Chem. 258, 14116-14119; Nishimura, J. S., and Mitchell, T. (1984) J. Biol. Chem. 259, 9642-9645). It was found, as in the case of the E. coli enzyme, that thiophosphoryl group release by GDP and by succinate plus CoA was stimulated by succinyl-CoA and GTP, respectively. The same result was observed at 1, 0.1, and 0.01 mg/ml, lending assurance that these phenomena were not exhibited by an aggregated form of the pig heart enzyme. While an alternating-sites catalytic cooperativity model is not ruled out for the E. coli enzyme, it is proposed that the NTP- and succinyl-CoA-stimulated release of thiophosphoryl groups from either enzyme involves a "same-site" mechanism, to be distinguished from an "other-site" mechanism.  相似文献   

12.
13.
Phosphonoacetaldehyde hydrolase (phosphonatase) from Bacillus cereus catalyzes hydrolytic P-C bond cleavage of phosphonoacetaldehyde (Pald) via a Schiff base intermediate formed with Lys53. A single turnover requires binding of Pald to the active site of the core domain, closure of the cap domain containing the Lys53 over the core domain, and dissociation of the products following catalysis. The ligand binding and dissociation steps occur from the "open conformer" (domains are separated and the active site is solvent-exposed), while catalysis occurs from the "closed conformer" (domains are bound together and the active site is sequestered from solvent). To test the hypothesis that bound substrate stabilizes the closed conformer, thus facilitating catalysis, the rates of chemical modification of Lys53 in the presence and absence of inert substrate and/or product analogues were compared. Acetylation of Lys53 with 2,4-dinitrophenylacetate (DNPA) resulted in the loss of enzyme activity. The pseudo-first-order rate constant for inactivation varied with pH. The pH profile of inactivation is consistent with a pK(a) of 9.3 for Lys53. The inhibitors tungstate and vinyl sulfonate, which are known to bind to active site residues comprising the core domain, protected Lys53 from acetylation. These results are consistent with a dynamic equilibrium between the open and closed conformations of phosphonatase and the hypothesis that ligand binding stabilizes the closed conformation required for catalytic turnover.  相似文献   

14.
The gene encoding S-adenosylhomocysteine (AdoHcy) hydrolase in Leishmania donovani was subcloned into an expression vector (pPROK-1) and expressed in Escherichia coli. Recombinant L. donovani AdoHcy hydrolase was then purified from cell-free extracts of E. coli using three chromatographic steps (DEAE-cellulose chromatofocusing, Sephacryl S-300 gel filtration, and Q-Sepharose ion exchange). The purified recombinant L. donovani enzyme exists as a tetramer with a molecular weight of approximately 48 kDa for each subunit. Unlike recombinant human AdoHcy hydrolase, the catalytic activity of the recombinant L. donovani enzyme was shown to be dependent on the concentration of NAD+ in the incubation medium. The dissociation constant (Kd) for NAD+ with the L. donovani enzyme was estimated to be 2.1 +/- 0.2 microM. The Km values for the natural substrates of the enzyme, AdoHcy, Ado, and Hcy, were determined to be 21 +/- 3, 8 +/- 2, and 82 +/- 5 microM, respectively. Several nucleosides and carbocyclic nucleosides were tested for their inhibitory effects on this parasitic enzyme, and the results suggested that L. donovani AdoHcy hydrolase has structural requirements for binding inhibitors different than those of the human enzyme. Thus, it may be possible to eventually exploit these differences to design specific inhibitors of this parasitic enzyme as potential antiparasitic agents.  相似文献   

15.
Fluoroacetate dehalogenase from Moraxella sp. B (FAc-DEX) catalyzes the hydrolytic dehalogenation of fluoroacetate and other haloacetates. Asp(105) of the enzyme acts as a nucleophile to attack the alpha-carbon of haloacetate to form an ester intermediate, which is subsequently hydrolyzed by a water molecule activated by His(272) [Liu, J.Q., Kurihara, T., Ichiyama, S., Miyagi, M., Tsunasawa, S., Kawasaki, H., Soda, K., and Esaki, N. (1998) J. Biol. Chem. 273, 30897-30902]. In this study, we found that FAc-DEX is inactivated concomitantly with defluorination of fluoroacetate by incubation with ammonia. Mass spectrometric analyses revealed that the inactivation of FAc-DEX is caused by nucleophilic attack of ammonia on the ester intermediate to convert the catalytic residue, Asp(105), into an asparagine residue. The results indicate that ammonia reaches the active site of FAc-DEX without losing its nucleophilicity. Analysis of the three-dimensional structure of the enzyme by homology modeling showed that the active site of the enzyme is mainly composed of hydrophobic and basic residues, which are considered to be essential for an ammonia molecule to retain its nucleophilicity. In a normal enzyme reaction, the hydrophobic environment is supposed to prevent hydration of the highly electronegative fluorine atom of the substrate and contribute to fluorine recognition by the enzyme. Basic residues probably play a role in counterbalancing the electronegativity of the substrate. These results demonstrate that catalysis-linked inactivation is useful for characterizing the active-site environment as well as for identifying the catalytic residue.  相似文献   

16.
2-Hydroxy-6-keto-nona-2,4-diene 1,9-dioic acid 5,6-hydrolase (MhpC) from Escherichia coli catalyses the hydrolytic cleavage of the extradiol ring fission product on the phenylpropionate catabolic pathway and is a member of the alpha/beta hydrolase family. The catalytic mechanism of this enzyme has previously been shown to proceed via initial ketonization of the dienol substrate (Henderson, I. M. J., and Bugg, T. D. H. (1997) Biochemistry 36, 12252-12258), followed by stereospecific fragmentation. Despite the implication of an active site serine residue in the alpha/beta hydrolase family, attempts to verify a putative acyl enzyme intermediate by radiochemical trapping methods using a (14)C-labeled substrate yielded a stoichiometry of <1% covalent intermediate, which could be accounted for by nonenzymatic processes. In contrast, incorporation of 5-6% of two atoms of (18)O from H(2)(18)O into succinic acid was observed using the natural substrate, consistent with the reversible formation of a gem-diol intermediate. Furthermore, time-dependent incorporation of (18)O from H(2)(18)O into the carbonyl group of a nonhydrolysable analogue 4-keto-nona-1,9-dioic acid was observed in the presence of MhpC, consistent with enzyme-catalyzed attack of water at the ketone carbonyl. These results favor a catalytic mechanism involving base-catalyzed attack of water, rather than nucleophilic attack of an active site serine. The implication of this work is that the putative active site serine in this enzyme may have an alternative function, for example, as a base.  相似文献   

17.
Neoplanocin A, a cyclopentenyl analog of adenosine, has been shown recently to be a tight binding inhibitor of S-adenosylhomocysteine (AdoHcy) hydrolase (EC 3.3.1.1), exhibiting a stoichiometry of one molecule of inhibitor per molecule of the enzyme tetramer (Borchardt, R. T., Keller, B. T., and Patel-Thombre, U. (1984) J. Biol. Chem. 259, 4353-4358). In the present study a detailed analysis was performed of the possible role of the enzyme-bound NAD+ in the inactivation of AdoHcy hydrolase by neplanocin A. The NAD+/NADH content was quantitated using a fluorescence technique. The native enzyme showed intrinsic fluorescence with an emission maximum at 460 nm when excited at 340 nm, partially due to NADH bound to the enzyme. It was found that the content of NAD+ and NADH in freshly prepared, native enzyme is equal, having a stoichiometry of two nucleotides per enzyme molecule (tetramer). In addition, it was observed that the enzymatic activity of the native enzyme can be increased by about 30% following preincubation with NAD+. Furthermore, it was demonstrated that the mechanism of inhibition of AdoHcy hydrolase by neplanocin A involves the reduction of enzymatically bound NAD+ to NADH. Catalytic activity of the inactivated enzyme could be fully recovered in a time-dependent manner by further incubation with NAD+ (but not NADH). It was also found that inhibition by neplanocin A does not involve dissociation of the bound NAD+ or NADH from the enzyme, but simply reduction of the NAD+ to NADH.  相似文献   

18.
We have recently reported that a functional alpha-L-fucosidase could be expressed by a single insertional mutation in the region of overlap between the ORFs SSO11867 and SSO3060 of the hyperthermophilic Archaeon Sulfolobus solfataricus [Cobucci-Ponzano et al. J. Biol. Chem. (2003) 278, 14622-14631]. This enzyme, belonging to glycoside hydrolase family 29 (GH29), showed micromolar specificity for p-nitrophenyl-alpha-L-fucoside (pNp-Fuc) and promoted transfucosylation reactions by following a reaction mechanism in which the products retained the anomeric configuration of the substrate. The active site residues in GH29 enzymes are still unknown. We describe here the identification of the catalytic nucleophile of the reaction in the alpha-L-fucosidase from S. solfataricus by reactivation with sodium azide of the mutant Asp242Gly that shows a 10(3)-fold activity reduction on pNp-Fuc. The detailed stereochemical analysis of the fucosyl-azide produced by the mutant reactivated on pNp-Fuc revealed its inverted (beta-fucosyl azide) configuration compared with the substrate. This allows for the first time the unambiguous assignment of Asp242, and its homologous residues, as the nucleophilic catalytic residues of GH29 alpha-L-fucosidases. This is the first time that this approach is used for alpha-L-glycosidases, widening the applicability of this method.  相似文献   

19.
Insect acetylcholinesterase (AChE), an enzyme whose catalytic site is located at the bottom of a gorge-like structure, hydrolyzes its substrate over a wide range of concentrations (from 2 microm to 300 mm). AChE is activated at low substrate concentrations and inhibited at high substrate concentrations. Several rival kinetic models have been developed to try to describe and explain this behavior. One of these models assumes that activation at low substrate concentrations partly results from an acceleration of deacetylation of the acetylated enzyme. To test this hypothesis, we used a monomethylcarbamoylated enzyme, which is considered equivalent to the acylated form of the enzyme and a non-hydrolyzable substrate analog, 4-oxo-N,N,N-trimethylpentanaminium iodide. It appears that this substrate analog increases the decarbamoylation rate by a factor of 2.2, suggesting that the substrate molecule bound at the activation site (K(d) = 130 +/- 47 microm) accelerates deacetylation. These two kinetic parameters are consistent with our analysis of the hydrolysis of the substrate. The location of the active site was investigated by in vitro mutagenesis. We found that this site is located at the rim of the active site gorge. Thus, substrate positioning at the rim of the gorge slows down the entrance of another substrate molecule into the active site gorge (Marcel, V., Estrada-Mondaca, S., Magné, F., Stojan, J., Klaébé, A., and Fournier, D. (2000) J. Biol. Chem. 275, 11603-11609) and also increases the deacylation step. This results in an acceleration of enzyme turnover.  相似文献   

20.
Glycine N-methyltransferase (S-adenosyl-l-methionine: glycine methyltransferase, EC 2.1.1.20; GNMT) catalyzes the AdoMet-dependent methylation of glycine to form sarcosine (N-methylglycine). Unlike most methyltransferases, GNMT is a tetrameric protein showing a positive cooperativity in AdoMet binding and weak inhibition by S-adenosylhomocysteine (AdoHcy). The first crystal structure of GNMT complexed with AdoMet showed a unique "closed" molecular basket structure, in which the N-terminal section penetrates and corks the entrance of the adjacent subunit. Thus, the apparent entrance or exit of the active site is not recognizable in the subunit structure, suggesting that the enzyme must possess a second, enzymatically active, "open" structural conformation. A new crystalline form of the R175K enzyme has been grown in the presence of an excess of AdoHcy, and its crystal structure has been determined at 3.0 A resolution. In this structure, the N-terminal domain (40 amino acid residues) of each subunit has moved out of the active site of the adjacent subunit, and the entrances of the active sites are now opened widely. An AdoHcy molecule has entered the site occupied in the "closed" structure by Glu15 and Gly16 of the N-terminal domain of the adjacent subunit. An AdoHcy binds to the consensus AdoMet binding site observed in the other methyltransferase. This AdoHcy binding site supports the glycine binding site (Arg175) deduced from a chemical modification study and site-directed mutagenesis (R175K). The crystal structures of WT and R175K enzymes were also determined at 2.5 A resolution. These enzyme structures have a closed molecular basket structure and are isomorphous to the previously determined AdoMet-GNMT structure. By comparing the open structure to the closed structure, mechanisms for auto-inhibition and for the forced release of the product AdoHcy have been revealed in the GNMT structure. The N-terminal section of the adjacent subunit occupies the AdoMet binding site and thus inhibits the methyltransfer reaction, whereas the same N-terminal section forces the departure of the potentially potent inhibitor AdoHcy from the active site and thus facilitates the methyltransfer reaction. Consequently GNMT is less active at a low level of AdoMet concentration, and is only weakly inhibited by AdoHcy. These properties of GNMT are particularly suited for regulation of the cellular AdoMet/AdoHcy ratio.  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