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.
Glycinamide ribonucleotide transformylase (GAR Tfase) has been the target of anti-neoplastic intervention for almost two decades. Here, we use a structure-based approach to design a novel folate analogue, 10-(trifluoroacetyl)-5,10-dideazaacyclic-5,6,7,8-tetrahydrofolic acid (10-CF(3)CO-DDACTHF, 1), which specifically inhibits recombinant human GAR Tfase (K(i) = 15 nM), but is inactive (K(i) > 100 microM) against other folate-dependent enzymes that have been examined. Moreover, compound 1 is a potent inhibitor of tumor cell proliferation (IC(50) = 16 nM, CCRF-CEM), which represents a 10-fold improvement over Lometrexol, a GAR Tfase inhibitor that has been in clinical trials. Thus, this folate analogue 1 is among the most potent and selective inhibitors known toward GAR Tfase. Contributing to its efficacious activity, compound 1 is effectively transported into the cell by the reduced folate carrier and intracellularly sequestered by polyglutamation. The crystal structure of human GAR Tfase with folate analogue 1 at 1.98 A resolution represents the first structure of any GAR Tfase to be determined with a cofactor or cofactor analogue without the presence of substrate. The folate-binding loop of residues 141-146, which is highly flexible in both Escherichia coli and unliganded human GAR Tfase structures, becomes highly ordered upon binding 1 in the folate-binding site. Computational docking of the natural cofactor into this and other apo or complexed structures provides a rational basis for modeling how the natural cofactor 10-formyltetrahydrofolic acid interacts with GAR Tfase, and suggests that this folate analogue-bound conformation represents the best template to date for inhibitor design.  相似文献   

2.
Glutathione reductase (GR) plays a vital role in maintaining the antioxidant levels of the cytoplasm by catalyzing the reduction of glutathione disulfide to reduced glutathione, thereby using NADPH and flavin adenine dinucleotide as cofactors. Chromatiaceae have evolved an unusual homolog that prefers both a modified substrate (glutathione amide disulfide [GASSAG]) and a different cofactor (NADH). Herein, we present the crystal structure of the Chromatium gracile glutathione amide reductase (GAR) both alone and in complex with NAD+. An altered charge distribution in the GASSAG binding pocket explains the difference in substrate specificity. The NADH binding pocket of GAR differs from that of wild-type GR as well as that of a low active GR that was engineered to mimic NADH binding. Based on the GAR structure, we propose two attractive rationales for producing an efficient GR enzyme with NADH specificity.  相似文献   

3.
Glycinamide ribonucleotide transformylase (GAR Tfase) is a key folate-dependent enzyme in the de novo purine biosynthesis pathway and, as such, has been the target for antitumor drug design. Here, we describe the crystal structures of the human GAR Tfase (purN) component of the human trifunctional protein (purD-purM-purN) at various pH values and in complex with its substrate. Human GAR Tfase exhibits pH-dependent enzyme activity with its maximum around pH 7.5-8. Comparison of unliganded human GAR Tfase structures at pH 4.2 and pH 8.5 reveals conformational differences in the substrate binding loop, which at pH 4.2 occupies the binding cleft and prohibits substrate binding, while at pH 8.5 is permissive for substrate binding. The crystal structure of GAR Tfase with its natural substrate, beta-glycinamide ribonucleotide (beta-GAR), at pH 8.5 confirms this conformational isomerism. Surprisingly, several important structural differences are found between human GAR Tfase and previously reported E. coli GAR Tfase structures, which have been used as the primary template for drug design studies. While the E. coli structure gave valuable insights into the active site and formyl transfer mechanism, differences in structure and inhibition between the bacterial and mammalian enzymes suggest that the human GAR Tfase structure is now the appropriate template for the design of anti-cancer agents.  相似文献   

4.
Among the Chromatiaceae, the glutathione derivative gamma-l-glutamyl-l-cysteinylglycine amide, or glutathione amide, was reported to be present in facultative aerobic as well as in strictly anaerobic species. The gene (garB) encoding the central enzyme in glutathione amide cycling, glutathione amide reductase (GAR), has been isolated from Chromatium gracile, and its genomic organization has been examined. The garB gene is immediately preceded by an open reading frame encoding a novel 27.5-kDa chimeric enzyme composed of one N-terminal peroxiredoxin-like domain followed by a glutaredoxin-like C terminus. The 27.5-kDa enzyme was established in vitro to be a glutathione amide-dependent peroxidase, being the first example of a prokaryotic low molecular mass thiol-dependent peroxidase. Amino acid sequence alignment of GAR with the functionally homologous glutathione and trypanothione reductases emphasizes the conservation of the catalytically important redox-active disulfide and of regions involved in binding the FAD prosthetic group and the substrates glutathione amide disulfide and NADH. By establishing Michaelis constants of 97 and 13.2 microm for glutathione amide disulfide and NADH, respectively (in contrast to K(m) values of 6.9 mm for glutathione disulfide and 1.98 mm for NADPH), the exclusive substrate specificities of GAR have been documented. Specificity for the amidated disulfide cofactor partly can be explained by the substitution of Arg-37, shown by x-ray crystallographic data of the human glutathione reductase to hydrogen-bond one of the glutathione glycyl carboxylates, by the negatively charged Glu-21. On the other hand, the preference for the unusual electron donor, to some extent, has to rely on the substitution of the basic residues Arg-218, His-219, and Arg-224, which have been shown to interact in the human enzyme with the NADPH 2'-phosphate group, by Leu-197, Glu-198, and Phe-203. We suggest GAR to be the newest member of the class I flavoprotein disulfide reductase family of oxidoreductases.  相似文献   

5.
Multisubstrate adduct inhibitors (MAI) of glycinamide ribonucleotide transformylase (GAR Tfase), which incorporate key features of the folate cofactor and the beta-GAR substrate, typically exhibit K(i)'s in the picomolar range. However, these compounds have reduced bioavailability due to the incorporation of a negatively charged phosphate moiety that prevents effective cellular uptake. Thus, a folate analogue that is capable of adduct formation with the substrate on the enzyme active site could lead to a potent GAR Tfase inhibitor that takes advantage of the cellular folate transport systems. We synthesized a dibromide folate analogue, 10-bromo-10-bromomethyl-5,8,10-trideazafolic acid, that was an intermediate designed to assemble with the substrate beta-GAR on the enzyme active site. We have now determined the crystal structure of the Escherichia coli GAR Tfase/MAI complex at 1.6 A resolution to ascertain the nature and mechanism of its time-dependent inhibition. The high-resolution crystal structure clearly revealed the existence of a covalent adduct between the substrate beta-GAR and the folate analogue (K(i) = 20 microM). However, the electron density map surprisingly indicated a C10 hydroxyl in the adduct rather than a bromide and suggested that the multisubstrate adduct is not formed directly from the dibromide but proceeds via an epoxide. Subsequently, we demonstrated the in situ conversion of the dibromide to the epoxide. Moreover, synthesis of the authentic epoxide confirmed that its inhibitory, time-dependent, and cytotoxic properties are comparable to those of the dibromide. Further, inhibition was strongest when the dibromide or epoxide is preincubated with both enzyme and substrate, indicating that inhibition occurs via the enzyme-dependent formation of the multisubstrate adduct. Thus, the crystal structure revealed the successful formation of an enzyme-assembled multisubstrate adduct and highlighted a potential application for epoxides, and perhaps aziridines, in the design of efficacious GAR Tfase inhibitors.  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.
N6-(N-[(4-Azido-3,5,6-trifluoro)pyridin-2-yl]-2-aminoethyl)- adenosine 5'-monophosphate has been synthesized and evidence presented for its structural assignment by ultraviolet and 19F-NMR spectroscopies. Its photolysis was shown to occur within 5 min. This AMP derivative behaves as a competitive inhibitor of NAD+ in horse-liver-alcohol-dehydrogenase-promoted oxidation of ethanol, with a Ki (0.95 mM) comparable to the Ki of AMP (1.9 mM). Moreover it is an activator of the enzyme when nicotinamide ribose is used as the oxidation cofactor. This activation is as good as that promoted by AMP or by the well known 8-azido-AMP. Upon photolysis of this new derivative in the presence of horse liver alcohol dehydrogenase, a covalent enzyme--analogue complex was isolated and assayed as a catalyst in the oxidation of ethanol using nicotinamide ribose as the cofactor. The reaction took place without complementation of AMP, indicating clearly that the AMP analogue is mainly covalently bound in the AMP-binding site, and that the linkage formed between the enzyme and the azido derivative has not dramatically altered the active site of the enzyme. A similar experiment with 8-azido-AMP produced a completely inactive complex.  相似文献   

8.
B Sherry  R H Abeles 《Biochemistry》1985,24(11):2594-2605
Methanol oxidase isolated from Hansenula polymorpha contains two distinct flavin cofactors in approximately equal amounts. One has been identified as authentic FAD and the other as a modified form of FAD differing only in the ribityl portion of the ribityldiphosphoadenosine side chain. The significance of this finding is as yet unknown. Previous studies have shown that cyclopropanol irreversibly inactivates methanol oxidase [Mincey, T., Tayrien, G., Mildvan, A. S., & Abeles, R. H. (1980) Proc. Natl. Acad. Sci. U.S.A. 77, 7099-7101]. We have now established that inactivation is accompanied by covalent modification of the flavin cofactor. The stoichiometry of this reaction is 1 mol of cyclopropanol/mol of active flavin. The structure of the covalent adduct was determined by NMR, IR, and UV spectral studies to be an N5,C4a-cyclic 4a,5-dihydroflavin. Reduction of the covalent adduct with NaBH4 at pH 9.0 before removal from the enzyme converted it to the 1-(ribityldiphosphoadenosine)-substituted 4-(3-hydroxypropyl)-2,3-dioxoquinoxaline. Cyclopropyl ring cleavage accompanies inactivation, and covalent bond formation occurs between a methylene carbon of cyclopropanol and N5 of flavin. Methanol oxidase was also reconstituted with 5-deazaflavin adenine dinucleotide (dFAD). Reconstituted enzyme did not catalyze the oxidation of alcohols to the corresponding aldehydes, nor did reduced reconstituted enzyme catalyze the reverse reaction. Incubation of reconstituted enzyme with cyclopropanol resulted in an absorbance decrease at 399 nm, but no irreversible covalent modification of the deazaflavin cofactor. A reversible addition complex between cyclopropanol and dFAD is formed. The structure of that complex was not definitively established, but it is likely that it is formed through the addition of cyclopropoxide to C5 of dFAD. The failure of dFAD-reconstituted methanol oxidase to catalyze the oxidation of substrate, as well as the lack of reaction with cyclopropanol, supports a radical mechanism for alcohol oxidation and cyclopropanol inactivation. Methanol oxidase catalyzes the oxidation of cyclopropylcarbinol to the corresponding aldehyde. No ring-opened products were detected. The failure to form ring-opened products has been used as an argument against radical processes [MacInnes, I., Nonhebel, D. C., Orsculik, S. T., & Suckling, C. J. (1982) J. Chem. Soc., Chem. Commun., 121-122]. We present arguments against this interpretation.  相似文献   

9.
We demonstrate here that Escherichia coli synthesizes two different glycinamide ribonucleotide (GAR) transformylases, both catalyzing the third step in the purine biosynthetic pathway. One is coded for by the previously described purN gene (GAR transformylase N), and a second, hitherto unknown, enzyme is encoded by the purT gene (GAR transformylase T). Mutants defective in the synthesis of the purN- and the purT-encoded enzymes were isolated. Only strains defective in both genes require an exogenous purine source for growth. Our results suggest that both enzymes may function to ensure normal purine biosynthesis. Determination of GAR transformylase T activity in vitro required formate as the C1 donor. Growth of purN mutants was inhibited by glycine. Under these conditions GAR accumulated. Addition of purine compounds or formate prevented growth inhibition. The regulation of the level of GAR transformylase T is controlled by the PurR protein and hypoxanthine.  相似文献   

10.
The purD gene of Escherichia coli encoding the enzyme glycinamide ribonucleotide (GAR) synthetase, which catalyzes the conversion of phosphoribosylamine (PRA), glycine, and MgATP to glycinamide ribonucleotide, MgADP, and Pi, has been cloned and sequenced. The protein, as deduced by the structural gene sequence, contains 430 amino acids and has a calculated Mr of 45,945. Construction of an overproducing strain behind a lambda pL promoter allowed a 4-fold purification of the protein to homogeneity. N-Terminal sequence analysis and comparison of the sequence with those of other GAR synthetases confirm the amino acid sequence deduced from the gene sequence. Initial velocity studies and product and dead-end inhibition studies are most consistent with a sequential ordered mechanism of substrate binding and product release in which PRA binds first followed by MgATP and then glycine; Pi leaves first, followed by loss of MgADP and finally GAR. Incubation of [18O]glycine, ATP, and PRA results in quantitative transfer of the 18O to Pi. GAR synthetase is very specific for its substrate glycine.  相似文献   

11.
The enzyme glycinamide ribonucleotide transformylase (GART) catalyzes the transfer of a formyl group from formyl tetrahydrofolate (fTHF) to glycinamide ribonucleotide (GAR), a process that is pH-dependent with pK(a) of approximately 8. Experimental studies of pH-rate profiles of wild-type and site-directed mutants of GART have led to the proposal that His108, Asp144, and GAR are involved in catalysis, with His108 being an acid catalyst, while forming a salt bridge with Asp144, and GAR being a nucleophile to attack the formyl group of fTHF. This model implied a protonated histidine with pK(a) of 9.7 and a neutral GAR with pK(a) of 6.8. These proposed unusual pK(a)s have led us to investigate the electrostatic environment of the active site of GART. We have used Poisson-Boltzmann-based electrostatic methods to calculate the pK(a)s of all ionizable groups, using the crystallographic structure of a ternary complex of GART involving the pseudosubstrate 5-deaza-5,6,7,8-THF (5dTHF) and substrate GAR. Theoretical mutation and deletion analogs have been constructed to elucidate pairwise electrostatic interactions between key ionizable sites within the catalytic site. Also, a construct of a more realistic catalytic site including a reconstructed pseudocofactor with an attached formyl group, in an environment with optimal local van der Waals interactions (locally minimized) that imitates closely the catalytic reactants, has been used for pK(a) calculations. Strong electrostatic coupling among catalytic residues His108, Asp144, and substrate GAR was observed, which is extremely sensitive to the initial protonation and imidazole ring flip state of His108 and small structural changes. We show that a proton can be exchanged between GAR and His108, depending on their relative geometry and their distance to Asp144, and when the proton is attached on His108, catalysis could be possible. Using the formylated locally minimized construct of GART, a high pK(a) for His108 was calculated, indicating a protonated histidine, and a low pK(a) for GAR(NH(2)) was calculated, indicating that GAR is in neutral form. Our results are in qualitative agreement with the current mechanistic picture of the catalytic process of GART deduced from the experimental data, but they do not reproduce the absolute magnitude of the pK(a)s extracted from fits of k(cat)-pH profiles, possibly because the static time-averaged crystallographic structure does not describe adequately the dynamic nature of the catalytic site during binding and catalysis. In addition, a strong effect on the pK(a) of GAR(NH(2)) is produced by the theoretical mutations of His108Ala and Asp144Ala, which is not in agreement with the observed insensitivity of the pK(a) of GAR(NH(2)) modeled from the experimental data using similar mutations. Finally, we show that important three-way electrostatic interactions between highly conserved His137, with His108 and Asp144, are responsible for stabilizing the electrostatic microenvironment of the catalytic site. In conclusion, our data suggest that further detailed computational and experimental work is necessary.  相似文献   

12.
The atomic structure of glycinamide ribonucleotide transformylase, an essential enzyme in purine biosynthesis, has been determined at 3.0 A resolution. The last three C-terminal residues and a sequence stretch of 18 residues (residues 113 to 130) are not visible in the electron density map. The enzyme forms a dimer in the crystal structure. Each monomer is divided into two domains, which are connected by a central mainly parallel seven-stranded beta-sheet. The N-terminal domain contains a Rossmann type mononucleotide fold with a phosphate ion bound to the C-terminal end of the first beta-strand. A long narrow cleft stretches from the phosphate to a conserved aspartic acid, Asp144, which has been suggested as an active-site residue. The cleft is lined by a cluster of residues, which are conserved between bacterial, yeast, avian and human enzymes, and likely represents the binding pocket and active site of the enzyme. GAR Tfase binds a reduced folate cofactor and glycinamide ribonucleotide for the catalysis of one of the initial steps in purine biosynthesis. Folate analogs and multi-substrate inhibitors of the enzyme have antineoplastic effects and the structure determination of the unliganded enzyme and enzyme-inhibitor complexes will aid the development of anti-cancer drugs.  相似文献   

13.
R67 dihydrofolate reductase (DHFR) shares no sequence or structural homology with chromosomal DHFRs. This enzyme arose recently in response to the clinical use of the antibacterial drug trimethoprim. R67 DHFR is a homotetramer possessing a single active site pore. A high-resolution crystal structure shows the homotetramer possesses exact 222 symmetry [Narayana, N., et al. (1995) Nat. Struct. Biol. 2, 1018-1025]. This symmetry dictates four symmetry-related binding sites must exist for each substrate as well as each cofactor. Isothermal titration calorimetry studies, however, indicate only two molecules bind: either two dihydrofolate molecules, two NADPH molecules, or one substrate and one cofactor [Bradrick, T. D., et al. (1996) Biochemistry 35, 11414-11424]. The latter is the productive ternary complex. To evaluate the role of S65, Q67, I68, and Y69 residues, located near the center of the active site pore, site-directed mutagenesis was performed. One mutation in the gene creates four mutations per active site pore which typically result in large cumulative effects. Steady state kinetic data indicate the mutants have altered K(m) values for both cofactor and substrate. For example, the Y69F R67 DHFR displays an 8-fold increase in the K(m) for dihydrofolate and a 20-fold increase in the K(m) for NADPH. Residues involved in ligand binding in R67 DHFR display very little, if any, specificity, consistent with their possessing dual roles in binding. These results support a model where R67 DHFR utilizes an unusual "hot spot" binding surface capable of binding both ligands and indicate this enzyme has adopted a novel yet simple approach to catalysis.  相似文献   

14.
The pathway for de novo biosynthesis of purine nucleotides contains two one-carbon transfer reactions catalyzed by glycinamide ribotide (GAR) and 5-aminoimidazole-4-carboxamide ribotide (AICAR) transformylases in which N10-formyltetrahydrofolate is the one-carbon donor. We have found that the antifolates methotrexate (MTX) and piritrexim (PTX) completely block the de novo purine pathway in mouse L1210 leukemia cells growing in culture but with only minor accumulations of GAR and AICAR to less than 5% of the polyphosphate derivatives of N-formylglycinamide ribotide (FGAR) which accumulate when the pathway is blocked completely by azaserine. This azaserine-induced accumulation of FGAR polyphosphates is completely abolished by MTX, indicating that inhibition of the pathway is at or before GAR transformylase (reaction 3; Lyons, S. D., and Christopherson, R. I. (1991) Biochem. Int. 24, 187-197). Three h after the addition of MTX (0.1 microM), cellular 5-phosphoribosyl-1-pyrophosphate has accumulated 3.4-fold while 6-methyl-mercaptopurine riboside (25 microM) induces a 6.3-fold accumulation. These data suggest that amido phosphoribosyltransferase catalyzing reaction 1 of the pathway is the primary site of inhibition. In support of this conclusion, we have found that dihydrofolate-Glu5, which accumulates in MTX-treated cells, is a noncompetitive inhibitor of amido phosphoribosyltransferase with a dissociation constant of 3.41 +/- 0.08 microM for interaction with the enzyme-glutamine complex in vitro. Folate-Glu5, MTX-Glu5, PTX, dihydrotriazine benzenesulfonyl fluoride, and AICAR also inhibit amido phosphoribosyltransferase.  相似文献   

15.
Heparin inhibits the intrinsic tenase complex (factor IXa-factor VIIIa) via interaction with a factor IXa exosite. To define the role of this exosite, human factor IXa with alanine substituted for conserved surface residues (R126, N129, K132, R165, N178) was characterized. Chromogenic substrate hydrolysis by the mutant proteases was reduced 20-30% relative to factor IXa wild type. Coagulant activity was moderately (N129A, K132A, K126A) or dramatically (R165A) reduced relative to factor IXa wild type. Kinetic analysis demonstrated a marked reduction in apparent cofactor affinity (23-fold) for factor IXa R165, and an inability to stabilize cofactor activity. Factor IXa K126A, N129A, and K132A demonstrated modest reductions ( approximately 2-fold) in apparent cofactor affinity, and accelerated decay of intrinsic tenase activity. In the absence of factor VIIIa, factor IXa N178A and R165A demonstrated a defective Vmax(app) for factor X activation. In the presence of factor VIIIa, Vmax(app) varied in proportion to the predicted factor IXa-factor VIIIa concentration. However, factor IXa R165A had a 65% reduction in the kcat for factor X, suggesting an additional effect on catalysis. The ability of factor IXa to compete for physical assembly into the intrinsic tenase complex was enhanced by EGR-chloromethylketone bound to the factor IXa active site or addition of factor X, and reduced by selected mutations in the heparin-binding exosite (N178A, K126A, R165A). These results suggest that the factor IXa heparin-binding exosite participates in both cofactor binding and protease activation, and cofactor affinity is linked to active site conformation and factor X interaction during enzyme assembly.  相似文献   

16.
The human mitochondrial NAD(P)+-dependent malic enzyme (m-NAD-ME) is a malic enzyme isoform with dual cofactor specificity and substrate binding cooperativity. Previous kinetic studies have suggested that Lys362 in the pigeon cytosolic NADP+-dependent malic enzyme has remarkable effects on the binding of NADP+ to the enzyme and on the catalytic power of the enzyme (Kuo, C. C., Tsai, L. C., Chin, T. Y., Chang, G.-G., and Chou, W. Y. (2000) Biochem. Biophys. Res. Commun. 270, 821-825). In this study, we investigate the important role of Gln362 in the transformation of cofactor specificity from NAD+ to NADP+ in human m-NAD-ME. Our kinetic data clearly indicate that the Q362K mutant shifted its cofactor preference from NAD+ to NADP+. The Km(NADP) and kcat(NADP) values for this mutant were reduced by 4-6-fold and increased by 5-10-fold, respectively, compared with those for the wild-type enzyme. Furthermore, up to a 2-fold reduction in Km(NADP)/Km(NAD) and elevation of kcat(NADP)/kcat(NAD) were observed for the Q362K enzyme. Mutation of Gln362 to Ala or Asn did not shift its cofactor preference. The Km(NADP)/Km(NAD) and kcat(NADP)/kcat(NAD) values for Q362A and Q362N were comparable with those for the wild-type enzyme. The DeltaG values for Q362A and Q362N with either NAD+ or NADP+ were positive, indicating that substitution of Gln with Ala or Asn at position 362 brings about unfavorable cofactor binding at the active site and thus significantly reduces the catalytic efficiency. Our data also indicate that the cooperative binding of malate became insignificant in human m-NAD-ME upon mutation of Gln362 to Lys because the sigmoidal phenomenon appearing in the wild-type enzyme was much less obvious that that in Q362K. Therefore, mutation of Gln362 to Lys in human m-NAD-ME alters its kinetic properties of cofactor preference, malate binding cooperativity, and allosteric regulation by fumarate. However, the other Gln362 mutants, Q362A and Q362N, have conserved malate binding cooperativity and NAD+ specificity. In this study, we provide clear evidence that the single mutation of Gln362 to Lys in human m-NAD-ME changes it to an NADP+-dependent enzyme, which is characteristic because it is non-allosteric, non-cooperative, and NADP+-specific.  相似文献   

17.
Glucooligosaccharide oxidase from Acremonium strictum has been screened for potential applications in oligosaccharide acid production and alternative carbohydrate detection, because it catalyzes the oxidation of glucose, maltose, lactose, cellobiose and cello- and maltooligosaccharides. We report the crystal structures of the enzyme and of its complex with an inhibitor, 5-amino-5-deoxy- cellobiono-1,5-lactam at 1.55- and 1.98-A resolution, respectively. Unexpectedly, the protein structure demonstrates the first known double attachment flavinylation, 6-S-cysteinyl, 8alpha-N1-histidyl FAD. The FAD cofactor is cross-linked to the enzyme via the C(6) atom and the 8alpha-methyl group of the isoalloxazine ring with Cys(130) and His(70), respectively. This sugar oxidase possesses an open carbohydrate-binding groove, allowing the accommodation of higher oligosaccharides. The complex structure suggests that this enzyme may prefer a beta-d-glucosyl residue at the reducing end with the conserved Tyr(429) acting as a general base to abstract the OH(1) proton in concert with the H(1) hydride transfer to the flavin N(5). Finally, a detailed comparison illustrates the structural conservation as well as the divergence between this protein and its related flavoenzymes.  相似文献   

18.
It has been established that phosphate analogues can activate glycogen phosphorylase reconstituted with pyridoxal in place of the natural cofactor pyridoxal 5'-phosphate (Change YC. McCalmont T, Graves DJ. 1983. Biochemistry 22:4987-4993). Pyridoxal phosphorylase b has been studied by kinetic, ultracentrifugation, and X-ray crystallographic experiments. In solution, the catalytically active species of pyridoxal phosphorylase b adopts a conformation that is more R-state-like than that of native phosphorylase b, but an inactive dimeric species of the enzyme can be stabilized by activator phosphite in combination with the T-state inhibitor glucose. Co-crystals of pyridoxal phosphorylase b complexed with either phosphite, phosphate, or fluorophosphate, the inhibitor glucose, and the weak activator IMP were grown in space group P4(3)2(1)2, with native-like unit cell dimensions, and the structures of the complexes have been refined to give crystallographic R factors of 18.5-19.2%, for data between 8 and 2.4 A resolution. The anions bind tightly at the catalytic site in a similar but not identical position to that occupied by the cofactor 5'-phosphate group in the native enzyme (phosphorus to phosphorus atoms distance = 1.2 A). The structural results show that the structures of the pyridoxal phosphorylase b-anion-glucose-IMP complexes are overall similar to the glucose complex of native T-state phosphorylase b. Structural comparisons suggest that the bound anions, in the position observed in the crystal, might have a structural role for effective catalysis.  相似文献   

19.
Li L  Zhang L  Cook PF 《Biochemistry》2006,45(42):12680-12686
Crystal structures of 6-phosphogluconate dehydrogenase (6PGDH) from sheep liver indicate that S128 and N187 are within hydrogen-bonding distance of 6PG in the E:6PG binary complex and NADPH in the E:NADPH binary complex. In addition, H186 is also within hydrogen-bonding distance of NADPH in the E:NADPH binary complex, while in the E:6PG binary complex it is within hydrogen-bonding distance of S128 and close to N187. The structures suggest that this triad of residues may play a dual role during the catalytic reaction. Site-directed mutagenesis has been performed to mutate each of the three residues to alanine. All mutant enzymes exhibit a decrease in V/E(t) (the turnover number), ranging from 7- to 67-fold. An increase in the Km for 6PG (K(6PG)) was observed for S128A and H187A mutant enzymes, while for the H186A mutation, K(6PG) is decreased by a factor of 2. K(NADP) remains the same as the wild type enzyme for the S128A and H186A mutant enzyme, while it increases by 6-fold in the N187A mutant enzyme. An increased K(iNADPH) was measured for all of the mutant enzymes. The primary kinetic 13C-isotope effect is increased, while the primary deuterium kinetic isotope effect is decreased, indicating that the decarboxylation step has become more rate limiting under conditions where substrate is limiting. A quantitative analysis of the data suggests that the S128, H186, and N187 triad is multifunctional in the 6PGDH reaction and contributes as follows. The triad (1) participates in the precatalytic conformational change; (2) provides ground state binding affinity for 6PG and NADPH; and (3) affects the relative rates of reduction or decarboxylation of the 3-keto-6PG intermediate by anchoring the cofactor after hydride transfer, which is accompanied by the rotation of the nicotinamide ring around the N-glycosidic bond and displacement of C1 of 6PG, facilitating decarboxylation.  相似文献   

20.
Daum S  Tai CH  Cook PF 《Biochemistry》2003,42(1):106-113
O-Acetylserine sulfhydrylase catalyzes the synthesis of l-cysteine from O-acetyl-l-serine (OAS) and inorganic bisulfide. An anti-E2 mechanism has been proposed for the OASS-catalyzed elimination of acetate from OAS (Tai, C.-H., and Cook, P. F. (2001) Acc. Chem. Res. 34, 49-59). Site-directed mutagenesis was used to change S272 to alanine, which would be expected to eliminate the hydrogen bond to N1 of PLP or to aspartate, which would be expected to enhance the hydrogen-bonding interaction. Both mutant enzymes catalyze the overall reaction and are in fact still good enzymes, consistent with the proposed anti-E2 mechanism. Data suggest that hydrogen bonding to the pyridine ring does not play a significant role in the alpha,beta-elimination reaction catalyzed by OASS-A. The V/K(OAS), which reflects the first half-reaction, is identical to the wild-type enzyme in the case of the S272D mutant enzyme and is decreased by only a factor of 3 in the case of the S272A mutant enzyme. In the case of the alanine mutation, and to a lesser extent the aspartate mutation, a decrease in the rate of the elimination is compensated by an increase in affinity for OAS, leading to the observed second-order rate constant, V/K. The decrease in the rate of the elimination is proposed to result from a change in the orientation of the bound cofactor, as might be expected since one of the ligands that determines the position of the bound PLP has been changed. Consistent with a change in the orientation of the cofactor are the results from a number of the spectral probes. The visible CD data for the internal Schiff base have a molar ellipticity 50% that of wild-type enzyme, and the alpha-aminoacrylate intermediate has a molar ellipticity 25% that of wild-type enzyme. The alpha-aminoacrylate intermediate can be formed from l-cysteine and l-serine with the S272A,D mutant enzymes, but not with the wild-type enzyme, and taken together with the increased K(d) for the serine external Schiff base is consistent with a change in cofactor orientation in the active site. The long wavelength fluorescence emission for the S272A mutant enzyme, attributed to F?rster resonance energy transfer (McClure, G. D., Jr., and Cook, P. F. (1994) Biochemistry 33, 1647-1683) has an intensity near zero, as compared to significant fluorescence for the wild-type enzyme.  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