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.
Aldo-keto reductases (AKRs) constitute a large protein superfamily of mainly NAD(P)-dependent oxidoreductases involved in carbonyl metabolism. Catalysis is promoted by a conserved tetrad of active site residues (Tyr, Lys, Asp and His). Recent results of structure-function relationship studies for xylose reductase (AKR2B5) require an update of the proposed catalytic mechanism. Electrostatic stabilization by the epsilon-NH3+ group of Lys is a key source of catalytic power of xylose reductase. A molecular-level analysis of the substrate binding pocket of xylose reductase provides a case of how a very broadly specific AKR achieves the requisite selectivity for its physiological substrate and could serve as the basis for the design of novel reductases with improved specificities for biocatalytic applications.  相似文献   

2.
The primary structure of the aldose xylose reductase from Candida tenuis (CtAR) is shown to be 39% identical to that of human aldose reductase (hAR). The catalytic tetrad of hAR is completely conserved in CtAR (Tyr51, Lys80, Asp46, His113). The amino acid residues involved in binding of NADPH by hAR (D.K. Wilson, et al., Science 257 (1992) 81-84) are 64% identical in CtAR. Like hAR the yeast enzyme is specific for transferring the 4-pro-R hydrogen of the coenzyme. These properties suggest that CtAR is a member of the aldo/keto reductase superfamily. Unlike hAR the enzyme from C. tenuis has a dual coenzyme specificity and shows similar specificity constants for NADPH and NADH. It binds NADP(+) approximately 250 times less tightly than hAR. Typical turnover numbers for aldehyde reduction by CtAR (15-20 s(-1)) are up to 100-fold higher than corresponding values for hAR, probably reflecting an overall faster dissociation of NAD(P)(+) in the reaction catalyzed by the yeast enzyme.  相似文献   

3.
Ye Q  Hyndman D  Li X  Flynn TG  Jia Z 《Proteins》2000,38(1):41-48
Chinese hamster ovary (CHO) reductase is an enzyme belonging to the aldo-keto reductase (AKR) superfamily that is induced by the aldehyde-containing protease inhibitor ALLN (Inoue, Sharma, Schimke, et al., J Biol Chem 1993;268: 5894). It shows 70% sequence identity to human aldose reductase (Hyndman, Takenoshita, Vera, et al., J Biol Chem 1997;272:13286), which is a target for drug design because of its implication in diabetic complications. We have determined the crystal structure of CHO reductase complexed with nicotinamide adenine dinucleotide phosphate (NADP)+ to 2.4 A resolution. Similar to aldose reductase and other AKRs, CHO reductase is an alpha/beta TIM barrel enzyme with cofactor bound in an extended conformation. All key residues involved in cofactor binding are conserved with respect to other AKR members. CHO reductase shows a high degree of sequence identity (91%) with another AKR member, FR-1 (mouse fibroblast growth factor-regulated protein), especially around the variable C-terminal end of the protein and has a similar substrate binding pocket that is larger than that of aldose reductase. However, there are distinct differences that can account for differences in substrate specificity. Trp111, which lies horizontal to the substrate pocket in all other AKR members is perpendicular in CHO reductase and is accompanied by movement of Leu300. This coupled with movement of loops A, B, and C away from the active site region accounts for the ability of CHO reductase to bind larger substrates. The position of Trp219 is significantly altered with respect to aldose reductase and appears to release Cys298 from steric constraints. These studies show that AKRs such as CHO reductase are excellent models for examining the effects of subtle changes in amino acid sequence and alignment on binding and catalysis.  相似文献   

4.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

5.
The interactions of transcobalamin II (TC II), intrinsic factor (IF) and R-type binding protein of cobalamin (Cb1, vitamin B12) with the hydrophobic chromatography matrix Phenyl-Sepharose CL-4B were investigated. IF-Cb1 and R-Cb1 complexes were not adsorbed on Phenyl-Sepharose at room temperature or at 4°C with buffer containing 50 mM sodium phosphate, pH 7.4 containing 150 mM sodium chloride. The TC II-Cb1 complex adsorbed and could be eluted with buffer containing 50% vv glycerol. IF without Cb1 adsorbed and was eluted with 50% glycerol at room temperature and 4°C. At room temperature, R binder without Cb1 eluted with buffer, but later than the R-Cb1 complex. At 4°C, R binder completely adsorbed to the matrix. TC II-without Cb1 bound to the matrix at 4°C and room temperature and could not be eluted with glycerol. These results suggest that Cb1 binding proteins can be separated and identified based on their hydrophobic properties. In addition, upon binding Cb1, TC II, IF and R-type binders undergo a conformational change such that the protein-Cb1 complex shows reduced hydrophobicity.  相似文献   

6.
The 2.2 Å X-ray crystal structure of Candida tenuis xylose reductase (AKR2B5) bound with NADP+ reveals that Phe-114 contributes to the substrate binding pocket of the enzyme. In the related human aldose reductase (AKR1B1), this phenylalanine is replaced by a tryptophan. The side chain of Trp was previously implicated in forming a hydrogen bond with bound substrate or inhibitor. The apparent Michaelis constant of AKR2B5 for xylose (Km≈90 mM) is 60 times that of AKR1B1, perhaps because critical enzyme–substrate interactions of Trp are not available to Phe-114. We, therefore, prepared a Phe-114→Trp mutant (F114W) of AKR2B5, to mimic the aldose reductase relationship in xylose reductase. Detailed analysis of the kinetic consequences in purified F114W revealed that the Km values for xylose and xylitol at pH 7.0 and 25°C were increased 5.1- and 4.4-fold, respectively, in the mutant compared with the wild-type. Turnover numbers (kcat) of F114W for xylose reduction and xylitol oxidation were half those of the wild-type. Apparent dissociation constants of NADH (KiNADH=44 µM) and NAD+ (KiNAD+=177 µM) were increased 1.6- and 1.4-fold in comparison with values of KiNADH and KiNAD+ for the wild-type, respectively. Catalytic efficiencies (kcat/Km) for NADH-dependent reduction of different aldehydes were between 3.1- and 31.5-fold lower than the corresponding kcat/Km values of the wild-type. Therefore, replacement of Phe-114 with Trp weakens rather than strengthens apparent substrate binding by AKR2B5, suggesting that xylose reductase exploits residue 114 in a different manner from aldose reductase.  相似文献   

7.
The industrial yeast Candida utilis can grow on media containing xylose as sole carbon source, but cannot ferment it to ethanol. The deficiency might be due to the low activity of NADPH-preferring xylose reductase (XR) and NAD(+)-dependent xylitol dehydogenase (XDH), which convert xylose to xylulose, because C. utilis can ferment xylulose. We introduced multiple site-directed mutations in the coenzyme binding sites of XR and XDH derived from the xylose-fermenting yeast Candida shehatae to alter their coenzyme specificities. Several combinations of recombinant and native XRs and XDHs were tested. Highest productivity was observed in a strain expressing CsheXR K275R/N277D (NADH-preferring) and native CsheXDH (NAD(+)-dependent), which produced 17.4 g/L of ethanol from 50 g/L of xylose in 20 h. Analysis of the genes responsible for ethanol production from the xylose capacity of C. utilis indicated that the introduction of CsheXDH was essential, while overexpression of CsheXR K275R/N277D improved efficiency of ethanol production.  相似文献   

8.
Apo and holo forms of lactoferrin (LF) from caprine and bovine species have been characterized and compared with regard to the structural stability determined by thermal denaturation temperature values (T m), at pH 2.0–8.0. The bovine lactoferrin (bLF) showed highest thermal stability with a T m of 90 ± 1°C at pH 7.0 whereas caprine lactoferrin (cLF) showed a lower T m value 68 ± 1°C. The holo form was much more stable than the apo form for the bLF as compared to cLF. When pH was gradually reduced to 3.0, the T m values of both holo bLF and holo cLF were reduced showing T m values of 49 ± 1 and 40 ± 1°C, respectively. Both apo and holo forms of cLF and bLF were found to be most stable at pH 7.0. A significant loss in the iron content of both holo and apo forms of the cLF and bLF was observed when pH was decreased from 7.0 to 2.0. At the same time a gradual unfolding of the apo and holo forms of both cLF and bLF was shown by maximum exposure of hydrophobic regions at pH 3.0. This was supported with a loss in α-helix structure together with an increase in the content of unordered (aperiodic) structure, while β structure seemed unchanged at all pH values. Since LF is used today as fortifier in many products, like infant formulas and exerts many biological functions in human, the structural changes, iron binding and release affected by pH and thermal denaturation temperature are important factors to be clarified for more than the bovine species.  相似文献   

9.
10.
D-Xylose is a major component of the carbohydrates derived from agricultural residues and forest products. Among more than two hundred known xylose-utilizing yeasts, only a few species are known to be able to ferment xylose anaerobically. Candida shehatae is one of such xylose-fermenting yeasts. Xylose reductase (E.C. 1.1.1.21) is a key enzyme responsible for xylose metabolism in xylose-utilizing as well as xylose-fermenting yeasts. In this paper, we report the development of a convenient and reliable procedure for the purification of xylose reductase from C. shehatae to near homogeneity. The amino acid composition and N-terminal sequence of the enzyme have also been analyzed. C. shehatae seems to contain only a single xylose reductase, but the enzyme has a dual coenzyme specificity for both NADPH and NADH. The enzyme is remarkably stable at room temperature and 4 degrees C.  相似文献   

11.
Xylose reductases catalyse the initial reaction in the xylose utilisation pathway, the NAD(P)H+H+ dependent reduction of xylose to xylitol. In this work, the xylose reductase gene from Candida tenuis CBS 4435 was cloned and successfully expressed in E. coli. From the purified and partially sequenced protein primers were deduced for PCR. The fragment obtained was used for Southern blot analysis and screening of a subgenomic library. The clone containing the open reading frame was sequenced; the gene consisted of 969 nucleotides coding for a 322 amino acids protein with a molecular mass of 36 kDa. Putative regulatory signals were identified with the help of a Saccharomyces cerevisiae regulatory sequence database. In order to express the xylose reductase in E. coli, the gene was placed under positive and negative control. At low temperatures, the xylose reductase was expressed in soluble and active form up to about 10% of the soluble protein; with rising temperatures formation of visible inclusion bodies occurred. In refolding experiments we were able to recover the major portion of xylose reductase activity from the pellet fraction.  相似文献   

12.
The structure of the rat liver aflatoxin dialdehyde reductase (AKR7A1) has been solved to 1.38-A resolution. Although it shares a similar alpha/beta-barrel structure with other members of the aldo-keto reductase superfamily, AKR7A1 is the first dimeric member to be crystallized. The crystal structure also reveals details of the ternary complex as one subunit of the dimer contains NADP(+) and the inhibitor citrate. Although the underlying catalytic mechanism appears similar to other aldo-keto reductases, the substrate-binding pocket contains several charged amino acids (Arg-231 and Arg-327) that distinguish it from previously characterized aldo-keto reductases with respect to size and charge. These differences account for the substrate specificity for 4-carbon acid-aldehydes such as succinic semialdehyde and 2-carboxybenzaldehyde as well as for the idiosyncratic substrate aflatoxin B(1) dialdehyde of this subfamily of enzymes. Structural differences between the AKR7A1 ternary complex and apoenzyme reveal a significant hinged movement of the enzyme involving not only the loops of the structure but also parts of the alpha/beta-barrel most intimately involved in cofactor binding.  相似文献   

13.
Molecular dynamic (MD) simulations have been performed on Tth-MCO, a hyperthermophilic multicopper oxidase from thermus thermophilus HB27, in the apo as well as the holo form, with the aim of exploring the structural dynamic properties common to the two conformational states. According to structural comparison between this enzyme and other MCOs, the substrate in process to electron transfer in an outer-sphere event seems to transiently occupy a shallow and overall hydrophobic cavity near the Cu type 1 (T1Cu). The linker connecting the β-strands 21 and 24 of the second domain (loop (β21-β24)(D2)) has the same conformation in both states, forming a flexible lid at the entrance of the electron-transfer cavity. Loop (β21-β24)(D2) has been tentatively assigned a role occluding the access to the electron-transfer site. The dynamic of the loop (β21-β24)(D2) has been investigated by MD simulation, and results show that the structures of both species have the same secondary and tertiary structure during almost all the MD simulations. In the simulation, loop (β21-β24)(D2) of the holo form undergoes a higher mobility than in the apo form. In fact, loop (β21-β24)(D2) of the holo form experiences a conformational change which enables exposure to the electron-transfer site (open conformation), while in the apo form the opposite effect takes place (closed conformation). To confirm the hypothesis that the open conformation might facilitate the transient electron-donor molecule occupation of the site, the simulation was extended another 40 ns with the electron-donor molecule docked into the protein cavity. Upon electron-donor molecule stabilization, loops near the cavity reduce their mobility. These findings show that coordination between the copper and the protein might play an important role in the general mobility of the enzyme, and that the open conformation seems to be required for the electron transfer process to T1Cu.  相似文献   

14.
Xylose reductase from the yeast Candida tenuis (CtXR) is a family 2 member of the aldo-keto reductase (AKR) superfamily of proteins and enzymes. Active site His-113 is conserved among AKRs, but a unified mechanism of how it affects catalytic activity is outstanding. We have replaced His-113 by alanine using site-directed mutagenesis, determined a 2.2 A structure of H113A mutant bound to NADP(+), and compared catalytic reaction profiles of NADH-dependent reduction of different aldehydes catalyzed by the wild type and the mutant. Deuterium kinetic isotope effects (KIEs) on k(cat) and k(cat)/K(m xylose) show that, relative to the wild type, the hydride transfer rate constant (k(7) approximately 0.16 s(-1)) has decreased about 1000-fold in H113A whereas xylose binding was not strongly affected. No solvent isotope effect was seen on k(cat) and k(cat)/K(m xylose) for H113A, suggesting that proton transfer has not become rate-limiting as a result of the mutation. The pH profiles of log(k(cat)/K(m xylose)) for the wild type and H113A decreased above apparent pK(a) values of 8.85 and 7.63, respectively. The DeltapK(a) of -1.2 pH units likely reflects a proximally disruptive character of the mutation, affecting the position of Asp-50. A steady-state kinetic analysis for H113A-catalyzed reduction of a homologous series of meta-substituted benzaldehyde derivatives was carried out, and quantitative structure-reactivity correlations were used to factor the observed kinetic substituent effect on k(cat) and k(cat)/K(m aldehyde) into an electronic effect and bonding effects (which are lacking in the wild type). Using the Hammett sigma scale, electronic parameter coefficients (rho) of +0.64 (k(cat)) and +0.78 (k(cat)/K(m aldehyde)) were calculated and clearly differ from rho(k(cat)/K(aldehyde)) and rho(k(cat)) values of +1.67 and approximately 0.0, respectively, for the wild-type enzyme. Hydride transfer rate constants of H113A, calculated from kinetic parameters and KIE data, display a substituent dependence not seen in the corresponding wild-type enzyme rate constants. An enzymic mechanism is proposed in which His-113, through a hydrogen bond from Nepsilon2 to aldehyde O1, assists in catalysis by optimizing the C=O bond charge separation and orbital alignment in the ternary complex.  相似文献   

15.
Cellular retinol-binding proteins (CRBPs) are cytoplasmic retinol-specific binding proteins. Mammalian CRBPs have been thoroughly characterised previously. Here we report on the identification and X-ray structural analysis of the apo (1.7A resolution) and holo (1.4A resolution) forms of a zebrafish CRBP. According to amino acid sequence and structure analyses, the zebrafish CRBP that we have identified resembles closely mammalian CRBP II, suggesting that it is the zebrafish orthologue of this mammalian CRBP type. Zebrafish CRBP forms a tight complex with all-trans retinol, producing an absorption spectrum similar to those of mammalian holo-CRBPs, albeit slightly blue-shifted. The superposition of the alpha-carbon atoms of the liganded (complexed with retinol) and unliganded forms of zebrafish CRBP shows significant differences in correspondence of the betaC-betaD (residues 55-58) and betaE-betaF (residues 74-77) turns, providing evidence for the occurrence of conformational changes accompanying retinol binding/release. Remarkable and well-defined ligand-dependent conformational changes in the protein region comprising the two beta-turns affect both the main chain and the side-chains of several residues. The two beta-turns project towards the interior of the cavity devoid of ligand of the apoprotein. The side-chains of F57, Y60 and L77 change substantially their orientation and position in the apoprotein relative to the holoprotein. In the beta-barrel internal cavity of apo-CRBP they occupy some of the space that is otherwise occupied by bound retinol in holo-CRBP, and are displaced from these positions on ligand binding. These results indicate that a flexible area encompassing the betaC-betaD and betaE-betaF turns may serve as the ligand portal and that these turns undergo conformational changes associated with the not yet clarified mechanism of retinol binding and release in CRBPs.  相似文献   

16.
Sicinska W  Westler WM  DeLuca HF 《Proteins》2005,61(3):461-467
Binding sites in the full-length, ligand-binding domain of rat vitamin D receptor (LBD-rVDR) for an active hormone derived from vitamin D (1alpha,25-dihydroxyvitamin D(3)) and three of its C-2 substituted analogs were compared by nuclear magnetic resonance (NMR) spectroscopy. Specific residue labeled with [UL]-(15)N(2) Trp allowed assignment of the side-chain H(epsilon1) and N(epsilon1) resonances of the single tryptophan residue at position 282 in LBD-rVDR. Comparison of (1)H[(15)N] Heteronuclear Single Quantum Correlation (HSQC) spectra of apo and holo LBD-rVDR revealed that the position of the Trp282 H(epsilon1) and N(epsilon1) signals are sensitive to the presence of the ligand in the receptor cavity. Binding of the ligands to LBD-rVDR results in a shift of both Trp H(epsilon1) and N(epsilon1) resonances to lower frequencies. The results indicate that the interaction between the ligands and Trp282 is not responsible for differences in calcemic activity observed in vitamin D analogs.  相似文献   

17.
Acyl carrier protein (ACP) is a chloroplast-localized cofactor of fatty acid synthesis, desaturation, and acyl transfer. We have transformed tobacco with a chimeric gene consisting of the tobacco ribulose-1,5-bisphosphate carboxylase promoter and transit peptide and the sequence encoding the mature spinach ACP-I. Spinach ACP-I was expressed in the transformed plants at levels twofold to threefold higher than the endogenous tobacco ACPs as determined by protein immunoblots and assays of ACP in leaf extracts. In addition to these elevated levels of the holo form, there were high levels of apoACP-I, a form lacking the 4'-phosphopantetheine prosthetic group and not previously detected in vivo. The mature forms of both apoACP-I and holoACP-I were located in the chloroplasts, indicating that the transit peptide was cleaved and that attachment of the prosthetic group was not required for uptake into the plastid. There were also significant levels of spinach acyl-ACP-I, demonstrating that spinach ACP-I participated in tobacco fatty acid metabolism. Lipid analyses of the transformed plants indicated that the increased ACP levels caused no significant alterations in leaf lipid biosynthesis.  相似文献   

18.
R plasmid dihydrofolate reductase with a dimeric subunit structure   总被引:5,自引:0,他引:5  
Dihydrofolate reductase specified by plasmid R483 from a trimethoprim-resistant strain of Escherichia coli has been purified 2,000-fold to homogeneity using dye-ligand chromatography, gel filtration, and polyacrylamide gel electrophoresis. The protein migrated as a single band on nondenaturing polyacrylamide gel electrophoresis and had a specific activity of 250 mumol/mg min(-1). The molecular weight was estimated to be 32,000 by gel filtration and 39,000 by Ferguson analysis of polyacrylamide gel electrophoresis. When subjected to electrophoresis in the presence of sodium dodecyl sulfate, the protein migrated as a single 19,000-molecular weight species, a fact that suggests that the native enzyme is a dimer of similar or identical subunits. Antibody specific for R483-encoded dihydrofolate reductase did not cross-react with dihydrofolate reductase encoded by plasmid R67, T4 phage, E. coli RT500, or mouse L1210 leukemia cells. The amino acid sequence of the first 34 NH2-terminal residues suggests that the R483 plasmid dihydrofolate reductase is more closely related to the chromosomal dihydrofolate reductase than is the enzyme coded by plasmid R67.  相似文献   

19.
In this study the ability of various sugars and sugar alcohols to induce aldose reductase (xylose reductase) and xylitol dehydrogenase (xylulose reductase) activities in the yeast Candida tenuis was investigated. Both enzyme activities were induced when the organism was grown on d-xylose or l-arabinose as well as on the structurally related sugars d-arabinose or d-lyxose. Mixtures of d-xylose with the more rapidly metabolizable sugar d-glucose resulted in a decrease in the levels of both enzymes formed. These results show that the utilization of d-xylose by C. tenuis is regulated by induction and catabolite repression. Furthermore, the different patterns of induction on distinct sugars suggest that the synthesis of both enzymes is not under coordinate control.  相似文献   

20.
Abstract Cell-free extracts of xylose-grown Pachysolen tannophilus exhibited xylose reductase activity with both NADPH and NADH. The ratio of the NADPH- and NADH-dependent activities varied with growth conditions. Affinity chromatography of cell-free extracts resulted in a separation of two xylose reductases. One was active with both NADPH and NADH, the other was specific for NADPH. Apart from this coenzyme specificity, the two enzymes also differed in their affinities for xylose and NADPH. The role of the two enzymes in xylose metabolism is discussed in relation to attempts to use P. tannophilus for the alcoholic fermentation of wood sugars.  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