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.
Hulme EC  Curtis CA  Page KM  Jones PG 《Life sciences》1995,56(11-12):891-898
Site-directed mutagenesis has been used to evaluate the roles of the key aspartate and arginine residues in transmembrane domain three of the muscarinic receptors. The results suggest that the formation of an ionic bond between the Asp carboxylate group and the onium headgroup is essential to anchor acetylcholine in its active, bound conformation in both binary agonist-receptor and ternary agonist-receptor-G-protein complexes, but that secondary, non-productive binding modes, promoted by non-polar forces, may contribute to binary complex formation by other ligands. The positive charge of the arginyl side-chain is central to the recognition, and subsequent activation of G-proteins by the agonist-M1 mAChR complex.  相似文献   

2.
1-Deoxy-d-xylulose-5-phosphate reductoisomerase (DXR) is the second enzyme in the non-mevalonate pathway of isoprenoid biosynthesis. The structure of the apo-form of this enzyme from Zymomonas mobilis has been solved and refined to 1.9-A resolution, and that of a binary complex with the co-substrate NADPH to 2.7-A resolution. The subunit of DXR consists of three domains. Residues 1-150 form the NADPH binding domain, which is a variant of the typical dinucleotide-binding fold. The second domain comprises a four-stranded mixed beta-sheet, with three helices flanking the sheet. Most of the putative active site residues are located on this domain. The C-terminal domain (residues 300-386) folds into a four-helix bundle. In solution and in the crystal, the enzyme forms a homo-dimer. The interface between the two monomers is formed predominantly by extension of the sheet in the second domain. The adenosine phosphate moiety of NADPH binds to the nucleotide-binding fold in the canonical way. The adenine ring interacts with the loop after beta1 and with the loops between alpha2 and beta2 and alpha5 and beta5. The nicotinamide ring is disordered in crystals of this binary complex. Comparisons to Escherichia coli DXR show that the two enzymes are very similar in structure, and that the active site architecture is highly conserved. However, there are differences in the recognition of the adenine ring of NADPH in the two enzymes.  相似文献   

3.
Homoisocitrate dehydrogenase (HICDH) is involved in the α-aminoadipate pathway of lysine biosynthesis in some bacteria and higher fungi, and catalyses the oxidative decarboxylation of (2R,3S)-homoisocitrate into 2-ketoadipate using NAD(+) as a coenzyme. In this study, the crystal structure of Thermus thermophilus HICDH in a binary complex with a designed inhibitor, (2S,3S)-thiahomoisocitrate, has been determined at 2.6 ? resolution. The inhibitor observed as a decarboxylated product interacts through hydrogen bonding to Arg 118, Tyr 125 and Lys 171 in the active site. The induced fitting was also observed around the region consisting of residues 120-141, which shifted up to 2.8 ? towards the active site. In addition, it was found that the complex structure adopts a closed conformation in two domains. While the structure of apo-HICDH shows that a catalytic residue Tyr 125 and Arg 85 that engages in substrate recognition are flipped out of the active site, these residues turn towards the active site in the complex structure. The results revealed that they directly interact with a substrate and are involved in catalysis or substrate recognition. Furthermore, by comparing the binary complex with the quaternary complex of Escherichia coli isocitrate dehydrogenase, the substrate recognition mechanism of HICDH is also discussed.  相似文献   

4.
The crystal structure of the Escherichia coli DNA adenine methyltransferase (EcoDam) in a binary complex with the cofactor product S-adenosyl-L-homocysteine (AdoHcy) unexpectedly showed the bound AdoHcy in two alternative conformations, extended or folded. The extended conformation represents the catalytically competent conformation, identical to that of EcoDam-DNA-AdoHcy ternary complex. The folded conformation prevents catalysis, because the homocysteine moiety occupies the target Ade binding pocket. The largest difference between the binary and ternary structures is in the conformation of the N-terminal hexapeptide ((9)KWAGGK(14)). Cofactor binding leads to a strong change in the fluorescence of Trp(10), whose indole ring approaches the cofactor by 3.3A(.) Stopped-flow kinetics and AdoMet cross-linking studies indicate that the cofactor prefers binding to the enzyme after preincubation with DNA. In the presence of DNA, AdoMet binding is approximately 2-fold stronger than AdoHcy binding. In the binary complex the side chain of Lys(14) is disordered, whereas Lys(14) stabilizes the active site in the ternary complex. Fluorescence stopped-flow experiments indicate that Lys(14) is important for EcoDam binding of the extrahelical target base into the active site pocket. This suggests that the hexapeptide couples specific DNA binding (Lys(9)), AdoMet binding (Trp(10)), and insertion of the flipped target base into the active site pocket (Lys(14)).  相似文献   

5.
A Kamb  J S Finer-Moore  R M Stroud 《Biochemistry》1992,31(51):12876-12884
We have solved crystal structures of two complexes with Escherichia coli thymidylate synthase (TS) bound either to the cofactor analog N10-propargyl-5,8-dideazafolate (CB3717) or to a tighter binding polygutamyl derivative of CB3717. These structures suggest that cofactor binding alone is sufficient to induce the conformational change in TS; dUMP binding is not required. Because polyglutamyl folates are the primary cofactor form in vivo, and because they can bind more tightly than dUMP to TS, these structures may represent a key intermediate along the TS reaction pathway. These structures further suggest that the dUMP binding site is accessible in the TS-cofactor analog binary complexes. Conformational flexibility of the binary complex may permit dUMP to enter the active site of TS while the cofactor is bound. Alternatively, dUMP may enter the active site from the opposite side that the cofactor appears to enter; that is, through a portal flanked by arginines that also coordinate the phosphate group in the active site. Entry of dUMP through this portal may allow dUMP to bind to a TS-cofactor binary complex in which the complex has completed its conformational transition to the catalytically competent structure.  相似文献   

6.
Heteronuclear NMR methods have been used to probe the conformation of four complexes of Escherichia coli dihydrofolate reductase (DHFR) in solution. (1)H(N), (15)N, and (13)C(alpha) resonance assignments have been made for the ternary complex with folate and oxidized NADP(+) cofactor and the ternary complex with folate and a reduced cofactor analog, 5,6-dihydroNADPH. The backbone chemical shifts have been compared with those of the binary complex of DHFR with the substrate analog folate and the binary complex with NADPH (the holoenzyme). Analysis of (1)H(N) and (15)N chemical shifts has led to the identification of marker resonances that report on the active site conformation of the enzyme. Other backbone amide resonances report on the presence of ligands in the pterin binding pocket and in the adenosine and nicotinamide-ribose binding sites of the NADPH cofactor. The chemical shift data indicate that the enzyme populates two dominant structural states in solution, with the active site loops in either the closed or occluded conformations defined by X-ray crystallography; there is no evidence that the open conformation observed in some X-ray structures of E. coli DHFR are populated in solution.  相似文献   

7.
1-Deoxy-d-xylulose-5-phosphate reductoisomerase (DXR) is the second enzyme in the non-mevalonate pathway of isoprenoid biosynthesis. The structure of the apo-form of this enzyme from Zymomonas mobilis has been solved and refined to 1.9-Å resolution, and that of a binary complex with the co-substrate NADPH to 2.7-Å resolution. The subunit of DXR consists of three domains. Residues 1–150 form the NADPH binding domain, which is a variant of the typical dinucleotide-binding fold. The second domain comprises a four-stranded mixed β-sheet, with three helices flanking the sheet. Most of the putative active site residues are located on this domain. The C-terminal domain (residues 300–386) folds into a four-helix bundle. In solution and in the crystal, the enzyme forms a homo-dimer. The interface between the two monomers is formed predominantly by extension of the sheet in the second domain. The adenosine phosphate moiety of NADPH binds to the nucleotide-binding fold in the canonical way. The adenine ring interacts with the loop after β1 and with the loops between α2 and β2 and α5 and β5. The nicotinamide ring is disordered in crystals of this binary complex. Comparisons to Escherichia coli DXR show that the two enzymes are very similar in structure, and that the active site architecture is highly conserved. However, there are differences in the recognition of the adenine ring of NADPH in the two enzymes.  相似文献   

8.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid–base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR–Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

9.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid-base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR-Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

10.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

11.
The resonance Raman spectrum has been recorded for two different binary complexes formed between 2-carboxy-2'-hydroxy-5'-sulfoformazylbenzene (zincon) and liver alcohol dehydrogenase. The shifts in the zincon spectrum upon complexation with enzyme in one complex are similar to those in model compounds containing azo or formazyl linkages upon complexation of these with zinc. The results are interpreted in terms of complexation of zincon to a zinc atom at the enzyme active site. Since zincon is a coenzyme competitive inhibitor, it is probably bound at or near the coenzyme binding site; the results of this study, therefore, are useful in understanding the chemistry of zinc at the enzyme active site.  相似文献   

12.
CD38 is a novel multifunctional protein that serves not only as an antigen but also as an enzyme. It catalyzes the metabolism of cyclic ADP-ribose and nicotinic acid adenine dinucleotide phosphate, two structurally and functionally distinct Ca(2+) messengers targeting, respectively, the endoplasmic reticulum and lysosomal Ca(2+) stores. The protein has recently been crystallized and its three-dimensional structure solved to a resolution of 1.9 A. The crystal structure of a binary complex reveals critical interactions between residues at the active site and a bound substrate, providing mechanistic insights to its novel multi-functional catalysis. This article reviews the current advances in the understanding of the structural determinants that control the multiple enzymatic reactions catalyzed by CD38.  相似文献   

13.
L C Kurz  G R Drysdale 《Biochemistry》1987,26(9):2623-2627
The infrared spectrum of oxaloacetate bound in the active site of citrate synthase has been measured in the binary complex and in the ternary complex with the acetyl coenzyme A (CoA) enolate analogue carboxymethyl-CoA. The carbonyl stretching frequency of oxaloacetate in binary and ternary complexes is found at 1697 cm-1, a shift of 21 cm-1 to lower frequency relative to that of the free ligand. The line widths of the carbonyl absorption in enzyme complexes differ from that of the free ligand, decreasing from a value of 20 cm-1 for the free ligand to 10 cm-1 in the binary complex and 7 cm-1 in the ternary complex with carboxymethyl-CoA. The integrated absorbance of the carbonyl absorption in these enzyme complexes is significantly increased over that of the free ligand at the same concentration, increasing approximately 2-fold in the binary complex and approximately 3-fold in the ternary complex. These results indicate strong polarization of the carbonyl bond in the enzyme-substrate complexes and suggest that ground-state destabilization is a major catalytic strategy of citrate synthase.  相似文献   

14.
Although a single binary functional complex between cytochrome P450 (P450 or CYP for a specific isoform) and cytochrome P450 reductase (CPR) has been generally accepted in the literature, this simple model failed to explain the experimentally observed catalytic activity of recombinant CYP2E1 in dependence on the total concentration of the added CPR-K56Q mutant. Our rejection of the simplest 1:1 binding model was based on two independent lines of experimental evidence. First, under the assumption of the 1:1 binding model, separate analyses of titration curves obtained while varying either P450 or CPR concentrations individually produced contradictory results. Second, an asymmetric Job plot suggested the existence of higher order molecular complexes. To identify the most probable complexation mechanism, we generated a comprehensive data set where the concentrations of both P450 and P450 were varied simultaneously, rather than one at a time. The resulting two-dimensional data were globally fit to 32 candidate mechanistic models, involving the formation of binary, ternary, and quaternary P450.CPR complexes, in the absence or presence or P450 and CPR homodimers. Of the 32 candidate models (mechanisms), two models were approximately equally successful in explaining our experimental data. The first plausible model involves the binary complex P450.CPR, the quaternary complex (P450)2.(CPR)2, and the homodimer (P450)2. The second plausible model additionally involves a weakly bound ternary complex (P450)2.CPR. Importantly, only the binary complex P450.CPR seems catalytically active in either of the two most probable mechanisms.  相似文献   

15.
The structure of amylosucrase from Neisseria polysaccharea in complex with beta-D-glucose has been determined by X-ray crystallography at a resolution of 1.66 A. Additionally, the structure of the inactive active site mutant Glu328Gln in complex with sucrose has been determined to a resolution of 2.0 A. The D-glucose complex shows two well-defined D-glucose molecules, one that binds very strongly in the bottom of a pocket that contains the proposed catalytic residues (at the subsite -1), in a nonstrained (4)C(1) conformation, and one that binds in the packing interface to a symmetry-related molecule. A third weaker D-glucose-binding site is located at the surface near the active site pocket entrance. The orientation of the D-glucose in the active site emphasizes the Glu328 role as the general acid/base. The binary sucrose complex shows one molecule bound in the active site, where the glucosyl moiety is located at the alpha-amylase -1 position and the fructosyl ring occupies subsite +1. Sucrose effectively blocks the only visible access channel to the active site. From analysis of the complex it appears that sucrose binding is primarily obtained through enzyme interactions with the glucosyl ring and that an important part of the enzyme function is a precise alignment of a lone pair of the linking O1 oxygen for hydrogen bond interaction with Glu328. The sucrose specificity appears to be determined primarily by residues Asp144, Asp394, Arg446, and Arg509. Both Asp394 and Arg446 are located in an insert connecting beta-strand 7 and alpha-helix 7 that is much longer in amylosucrase compared to other enzymes from the alpha-amylase family (family 13 of the glycoside hydrolases).  相似文献   

16.
The dissociation constant for the complex of rhodanese and Cibacron Blue, determined by analytical affinity chromatography using rhodanese immobilized on controlled-pore glass (CPG) beads (200 nm pore diameter) and aminohexyl-Cibacron Blue, was 44 microM which agreed well with the kinetic inhibition constant, suggesting that the dye binds at or near the active site of this enzyme. Formation of a binary complex of the dye and lactate dehydrogenase (LDH) was also characterized by direct chromatography of LDH on CPG/immobilized Cibacron Blue (KD = 0.29 microM). The binary complex formed between LDH and NADH was characterized by analytical affinity chromatography using both CPG/immobilized LDH and immobilized Cibacron Blue. Since the dye competes with NADH in binding to the active site of LDH, competitive elution chromatography using the immobilized dye allows determination of the dissociation constant of the soluble LDH.NADH complex. Agreement between the dissociation constants determined by direct chromatography of NADH on immobilized LDH (KD = 1.4 microM) and that determined for the soluble complex (KD = 2.4 microM) indicates that immobilization of LDH did not affect the interaction. Formation of various binary, ternary and quaternary complexes of bovine liver glutamate dehydrogenase (GDH) with glutamate, NADPH, NADH, and ADP was also investigated using immobilized GDH. This approach allows characterization of the enzyme/ligand interactions without the complicating effect of enzyme self-association. The affinity for NADPH is considerably greater in the ternary complex (including glutamate) as compared to the binary complex (0.38 microM vs 22 microM); however, occupancy of the regulatory site by ADP greatly reduces the affinity in both complexes (6.4 microM and 43 microM, respectively).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The extracellular portions of the chains that comprise the human type I interferon receptor, IFNAR1 and IFNAR2, have been expressed and purified as recombinant soluble His-tagged proteins, and their interactions with each other and with human interferon-beta-1a (IFN-beta-1a) were studied by gel filtration and by cross-linking. By gel filtration, no stable binary complexes between IFN-beta-1a and IFNAR1, or between IFNAR1 and IFNAR2 were detected. However, a stable binary complex formed between IFN-beta-1a and IFNAR2. Analysis of binary complex formation using various molar excesses of IFN-beta-1a and IFNAR2 indicated that the complex had a 1:1 stoichiometry, and reducing SDS-PAGE of the binary complex treated with the cross-linking reagent dissucinimidyl glutarate (DSG) indicated that the major cross-linked species had an apparent Mr consistent with the sum of its two individual components. Gel filtration of a mixture of IFNAR1 and the IFN-beta-1a/IFNAR2 complex indicated that the three proteins formed a stable ternary complex. Analysis of ternary complex formation using various molar excesses of IFNAR1 and the IFN-beta-1a/IFNAR2 complex indicated that the ternary complex had a 1:1:1 stoichiometry, and reducing SDS-PAGE of the ternary complex treated with DSG indicated that the major cross-linked species had an apparent Mr consistent with the sum of its three individual components. We conclude that the ternary complex forms by the sequential association of IFN-beta-1a with IFNAR2, followed by the association of IFNAR1 with the preformed binary complex. The ability to produce the IFN-beta-1a/IFNAR2 and IFN-beta-1a/IFNAR1/IFNAR2 complexes make them attractive candidates for X-ray crystallography studies aimed at determining the molecular interactions between IFN-beta-1a and its receptor.  相似文献   

18.
The interaction between cytochrome P-450 and NADPH-cytochrome c reductase during catalysis has been investigated with a reconstituted monooxygenase system composed of the two purified enzyme components and synthetic phospholipid. Steady state kinetic data are consistent with a scheme in which the formation of a binary complex between the two proteins precedes catalysis. The formation of this binary complex is described by a simple mass action equation. In agreement with this equation, the observed Vmax for benzphetamine N-demethylation was found to be directly proportional to the calculated concentration of the cytochrome P-450 . reductase complex. Furthermore, with appropriate reductase/cytochrome P-450 mole ratios, the Vmax could be shown to be linearly dependent on either the reductase or the cytochrome P-450 concentration alone. In contrast, the Km parameter is independent of the complex concentration, indicating that no change in the rate-limiting step has occurred. Thus a distinction should be made between a rate-limiting enzyme component and the rate-limiting step in this multienzyme system.  相似文献   

19.
The carbon-13 NMR spectrum of oxaloacetate bound in the active site of citrate synthase has been obtained at 90.56 MHz. In the binary complex with enzyme, the positions of the resonances of oxaloacetate are shifted relative to those of the free ligand as follows: C-1 (carboxylate), -2.5 ppm; C-2 (carbonyl), +4.3 ppm; C-3 (methylene), -0.6 ppm; C-4 (carboxylate), +1.3 ppm. The change observed in the carbonyl chemical shift is successively increased in ternary complexes with the product [coenzyme A (CoA)], a substrate analogue (S-acetonyl-CoA), and an acetyl-CoA enolate analogue (carboxymethyl-CoA), reaching a value of +6.8 ppm from the free carbonyl resonance. Binary complexes are in intermediate to fast exchange on the NMR time scale with free oxaloacetate; ternary complexes are in slow exchange. Line widths of the methylene resonance in the ternary complexes suggest complete immobilization of oxaloacetate in the active site. Analysis of line widths in the binary complex suggests the existence of a dynamic equilibrium between two or more forms of bound oxaloacetate, primarily involving C-4. The changes in chemical shifts of the carbonyl carbon indicate strong polarization of the carbonyl bond or protonation of the carbonyl oxygen. Some of this carbonyl polarization occurs even in the binary complex. Development of positive charge on the carbonyl carbon enhances reactivity toward condensation with the carbanion/enolate of acetyl-CoA in the mechanism which has been postulated for this enzyme. The very large change in the chemical shift of the reacting carbonyl in the presence of an analogue of the enolate of acetyl-CoA supports this interpretation.  相似文献   

20.
Sanghani PC  Robinson H  Bosron WF  Hurley TD 《Biochemistry》2002,41(35):10778-10786
The human glutathione-dependent formaldehyde dehydrogenase is unique among the structurally studied members of the alcohol dehydrogenase family in that it follows a random bi bi kinetic mechanism. The structures of an apo form of the enzyme, a binary complex with substrate 12-hydroxydodecanoic acid, and a ternary complex with NAD+ and the inhibitor dodecanoic acid were determined at 2.0, 2.3, and 2.3 A resolution by X-ray crystallography using the anomalous diffraction signal of zinc. The structures of the enzyme and its binary complex with the primary alcohol substrate, 12-hydroxydodecanoic acid, and the previously reported binary complex with the coenzyme show that the binding of the first substrate (alcohol or coenzyme) causes only minor changes to the overall structure of the enzyme. This is consistent with the random mechanism of the enzyme where either of the substrates binds to the free enzyme. The catalytic-domain position in these structures is intermediate to the "closed" and "open" conformations observed in class I alcohol dehydrogenases. More importantly, two different tetrahedral coordination environments of the active site zinc are observed in these structures. In the apoenzyme, the active site zinc is coordinated to Cys44, His66 and Cys173, and a water molecule. In the inhibitor complex, the coordination environment involves Glu67 instead of the solvent water molecule. The coordination environment involving Glu67 as the fourth ligand likely represents an intermediate step during ligand exchange at the active site zinc. These observations provide new insight into metal-assisted catalysis and substrate binding in glutathione-dependent formaldehyde dehydrogenase.  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