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.
Amyes TL  Richard JP 《Biochemistry》2007,46(19):5841-5854
More than 80% of the rate acceleration for enzymatic catalysis of the aldose-ketose isomerization of (R)-glyceraldehyde 3-phosphate (GAP) by triosephosphate isomerase (TIM) can be attributed to the phosphodianion group of GAP [Amyes, T. L., O'Donoghue, A. C., and Richard, J. P. (2001) J. Am. Chem. Soc. 123, 11325-11326]. We examine here the necessity of the covalent connection between the phosphodianion and triose sugar portions of the substrate by "carving up" GAP into the minimal neutral two-carbon sugar glycolaldehyde and phosphite dianion pieces. This "two-part substrate" preserves both the alpha-hydroxycarbonyl and oxydianion portions of GAP. TIM catalyzes proton transfer from glycolaldehyde in D2O, resulting in deuterium incorporation that can be monitored by 1H NMR spectroscopy, with kcat/Km = 0.26 M-1 s-1. Exogenous phosphite dianion results in a very large increase in the observed second-order rate constant (kcat/Km)obsd for turnover of glycolaldehyde, and the dependence of (kcat/Km)obsd on [HPO32-] exhibits saturation. The data give kcat/Km = 185 M-1 s-1 for turnover of glycolaldehyde by TIM that is saturated with phosphite dianion so that the separate binding of phosphite dianion to TIM results in a 700-fold acceleration of proton transfer from carbon. The binding of phosphite dianion to the free enzyme (Kd = 38 mM) is 700-fold weaker than its binding to the fleeting complex of TIM with the altered substrate in the transition state (Kd = 53 muM); the total intrinsic binding energy of phosphite dianion in the transition state is 5.8 kcal/mol. We propose a physical model for catalysis by TIM in which the intrinsic binding energy of the substrate phosphodianion group is utilized to drive closing of the "mobile loop" and a protein conformational change that leads to formation of an active site environment that is optimally organized for stabilization of the transition state for proton transfer from alpha-carbonyl carbon.  相似文献   

2.
Triosephosphate isomerase (TIM), glycerol 3-phosphate dehydrogenase, and orotidine 5'-monophosphate decarboxylase each use the binding energy from the interaction of phosphite dianion with a flexible phosphate gripper loop to activate a second, phosphodianion-truncated, substrate towards enzyme-catalyzed proton transfer, hydride transfer, and decarboxylation, respectively. Studies on TIM suggest that the most important general effect of loop closure over the substrate phosphodianion, and the associated conformational changes, is to extrude water from the enzyme active site. This should cause a decrease in the effective active-site dielectric constant, and an increase in transition state stabilization from enhanced electrostatic interactions with polar amino acid side chains. The most important specific effect of these conformational changes is to increase the basicity of the carboxylate side chain of the active site glutamate base by its placement in a 'hydrophobic cage'.  相似文献   

3.
Phosphite is a stable phosphorus compound that, together with phosphate, made up a substantial part of the total phosphorus content of the prebiotic Earth's crust. Oxidation of phosphite to phosphate releases electrons at an unusually low redox potential (−690 mV at pH 7.0). Numerous aerobic and anaerobic bacteria use phosphite as a phosphorus source and oxidise it to phosphate for synthesis of nucleotides and other phosphorus-containing cell constituents. Only two pure cultures of strictly anaerobic bacteria have been isolated so far that use phosphite as an electron donor in their energy metabolism, the Gram-positive Phosphitispora fastidiosa and the Gram-negative Desulfotignum phosphitoxidans. The key enzyme of this metabolism is an NAD+-dependent phosphite dehydrogenase enzyme that phosphorylates AMP to ADP. These phosphorylating phosphite dehydrogenases were found to be related to nucleoside diphosphate sugar epimerases. The produced NADH is channelled into autotrophic CO2 fixation via the Wood-Ljungdahl (CO-DH) pathway, thus allowing for nearly complete assimilation of the substrate electrons into bacterial biomass. This extremely efficient type of electron flow connects energy and carbon metabolism directly through NADH and might have been important in the early evolution of life when phosphite was easily available on Earth.  相似文献   

4.
A general method for formulating complex thermodynamic systems in terms of hierarchical interactions has been developed, and has been applied in a previous analysis to hemoglobin oxygen binding data. Polyprotic acids can be considered a simple chemical model of thermodynamic interaction between ligand binding events. To further illustrate the hierarchical interaction approach it is applied to the analysis of the thermodynamic interactions between proton binding events in inorganic polyprotic acids. pK values for arsenate, carbonate, chromate, phosphate, phosphite, selenite, sulfide and sulfite were recast into hierarchical interaction terms. The intrinsic K(d,h) for protonation ranged from 8.8 x 10(-13) (M) for phosphate to 1.3 x 10(-6) (M) for chromate. Pairwise interactions (K(d,hh)) between protonation events ranged from 1.3 x 10(4) for phosphite to 9.4 x 10(5) for carbonate. Third order interactions (K(d,hhh)) were 0.91 and 0.51 for arsenate and phosphate, respectively, values relatively close to the no interaction value of 1. A principle feature of systems described by hierarchical interactions is that higher order interactions, representing more complex interactions, are less likely to be significant than lower order interactions, and this is further illustrated by these observations from polyprotic acids. The set of significant hierarchical interaction values can be used to predict values for as yet unobserved events, and projected pK values are made for all the polyprotic acids included in this study. Finally, application of this method to the protonation equilibria of water demonstrates a profound pairwise interaction between protonation events (K(d,hh) = 1.3 x 10(17)), which is attributed to oxygen's small size and lack of polarizability.  相似文献   

5.
Richard JP 《Biochemistry》2012,51(13):2652-2661
Triosephosphate isomerase (TIM) catalyzes the stereospecific 1,2-proton shift at dihydroxyacetone phosphate (DHAP) to give (R)-glyceraldehyde 3-phosphate through a pair of isomeric enzyme-bound cis-enediolate phosphate intermediates. The chemical transformations that occur at the active site of TIM were well understood by the early 1990s. The mechanism for enzyme-catalyzed isomerization is similar to that for the nonenzymatic reaction in water, but the origin of the catalytic rate acceleration is not understood. We review the results of experimental work that show that a substantial fraction of the large 12 kcal/mol intrinsic binding energy of the nonreacting phosphodianion fragment of TIM is utilized to activate the active site side chains for catalysis of proton transfer. Evidence is presented that this activation is due to a phosphodianion-driven conformational change, the most dramatic feature of which is closure of loop 6 over the dianion. The kinetic data are interpreted within the framework of a model in which activation is due to the stabilization by the phosphodianion of a rare, desolvated, loop-closed form of TIM. The dianion binding energy is proposed to drive the otherwise thermodynamically unfavorable desolvation of the solvent-exposed active site. This reduces the effective local dielectric constant of the active site, to enhance stabilizing electrostatic interactions between polar groups and the anionic transition state, and increases the basicity of the carboxylate side chain of Glu-165 that functions to deprotonate the bound carbon acid substrate. A rebuttal is presented to the recent proposal [Samanta, M., Murthy, M. R. N., Balaram, H., and Balaram, P. (2011) ChemBioChem 12, 1886-1895] that the cationic side chain of K12 functions as an active site electrophile to protonate the carbonyl oxygen of DHAP.  相似文献   

6.
The second order rate constant (k(cat)/K(m)) for decarboxylation of orotidine by yeast OMP decarboxylase (ODCase), measured by trapping (14)CO(2) released during the reaction, is 2 x 10(-4)M(-1)s(-1). This very low activity may be compared with a value of 3 x 10(7)M(-1)s(-1) for the action of yeast OMP decarboxylase on the normal substrate OMP. Both activities are strongly inhibited by 6-hydroxy UMP (BMP), and abrogated by mutation of Asp-96 to alanine. These results, in conjunction with the binding affinity of inorganic phosphate as a competitive inhibitor (K(i)=7 x 10(-4)M), imply an effective concentration of 1.1 x 10(9)M for the substrate phosphoryl group in stabilizing the transition state for enzymatic decarboxylation of OMP. The observed difference in rate (1.5 x 10(11)-fold) is the largest effect of a simple substituent that appears to have been reported for an enzyme reaction.  相似文献   

7.
The ptxD gene from Pseudomonas stutzeri WM88 encoding the novel phosphorus oxidizing enzyme NAD:phosphite oxidoreductase (trivial name phosphite dehydrogenase, PtxD) was cloned into an expression vector and overproduced in Escherichia coli. The heterologously produced enzyme is indistinguishable from the native enzyme based on mass spectrometry, amino-terminal sequencing, and specific activity analyses. Recombinant PtxD was purified to homogeneity via a two-step affinity protocol and characterized. The enzyme stoichiometrically produces NADH and phosphate from NAD and phosphite. The reverse reaction was not observed. Gel filtration analysis of the purified protein is consistent with PtxD acting as a homodimer. PtxD has a high affinity for its substrates with Km values of 53.1 +/- 6.7 microm and 54.6 +/- 6.7 microm, for phosphite and NAD, respectively. Vmax and kcat were determined to be 12.2 +/- 0.3 micromol x min(-1) x mg(-1) and 440 min(-1). NADP can substitute poorly for NAD; however, none of the numerous compounds examined were able to substitute for phosphite. Initial rate studies in the absence or presence of products and in the presence of the dead end inhibitor sulfite are most consistent with a sequential ordered mechanism for the PtxD reaction, with NAD binding first and NADH being released last. Amino acid sequence comparisons place PtxD as a new member of the d-2-hydroxyacid NAD-dependent dehydrogenases, the only one to have an inorganic substrate. To our knowledge, this is the first detailed biochemical study on an enzyme capable of direct oxidation of a reduced phosphorus compound.  相似文献   

8.
1H-3-Hydroxy-4-oxoquinaldine 2,4-dioxygenase (Hod) is a cofactor-less dioxygenase belonging to the alpha/beta hydrolase fold family, catalyzing the cleavage of 1H-3-hydroxy-4-oxoquinaldine (I) and 1H-3-hydroxy-4-oxoquinoline (II) to N-acetyl- and N-formylanthranilate, respectively, and carbon monoxide. Bisubstrate steady-state kinetics and product inhibition patterns of HodC, the C69A protein variant of Hod, suggested a compulsory-order ternary-complex mechanism, in which binding of the organic substrate precedes dioxygen binding, and carbon monoxide is released first. The specificity constants, k(cat)/K(m,A) and k(cat)/K(m,O)()2, were 1.4 x 10(8) and 3.0 x 10(5) M(-1) s(-1) with I and 1.2 x 10(5) and 0.41 x 10(5) M(-1) s(-1) with II, respectively. Whereas HodC catalyzes formation of the dianion of its organic substrate prior to dioxygen binding, HodC-H251A does not, suggesting that H251, which aligns with the histidine of the catalytic triad of the alpha/beta hydrolases, acts as general base in catalysis. Investigation of base-catalyzed dioxygenolysis of I by electron paramagnetic resonance (EPR) spectroscopy revealed formation of a resonance-stabilized radical upon exposure to dioxygen. Since in D(2)O spectral properties are not affected, exchangeable protons are not involved, confirming that the dianion is the reactive intermediate that undergoes single-electron oxidation. We suggest that in the ternary complex of the enzyme, direct single-electron transfer from the substrate dianion to dioxygen may occur, resulting in a radical pair. Based on the estimated spin distribution within the radical anion (observed in the model reaction of I), radical recombination may produce a C4- or C2-hydroperoxy(di)anion. Subsequent intramolecular attack would result in the 2,4-endoperoxy (di)anion that may collapse to the reaction products.  相似文献   

9.
The early steps of the proposed mechanistic pathway for dehydroquinate synthase have been probed with a series of substrate analogues. These analogues, 3-9, are structurally prohibited from undergoing the beta-elimination of inorganic phosphate that represents the committed step in the conversion of the substrate 3-deoxy-D-arabino-heptulosonate 7-phosphate (1) to dehydroquinate (2). In agreement with previous observations, the analogues that possess shortened side chains (3,5, and 6) bind more tightly to the enzyme than those (4 and 7-9) that are more nearly isosteric with the substrate. Two hitherto unrecognized factors that influence binding have been identified: (i) carbacylic analogues bind 25-100 times more tightly than the corresponding oxacyclic materials (indeed, the carbacyclic phosphonate 5 has a Ki value of 8 x 10(-10)M) and (ii) the side chain appears to be bound in a gauche conformation similar to the most stable conformation of the cis-vinylhomophosphonate 8. These trends in binding can be rationalized by considering the behavior of the analogues in the first two chemical steps of the mechanism: NAD+-mediated oxidation at C-5 and enolization at C-6 (the first part of the E1cB elimination of inorganic phosphate). Direct spectrophotometric determination of the equilibrium level of enzyme-bound NADH indicates that the carbacyclic analogues are more readily oxidized than the oxacyclic compounds, and this predictable difference in redox behavior is reflected in the observed differences in binding. The gauche conformation of the C-7 side chain appears to be required for proton abstraction from C-6, since only those analogues that can adopt this conformation undergo enzyme-catalyzed exchange of the C-6 proton with the solvent. This conformation positions one of the peripheral oxygens of the phosphate (or phosphonate) group close to the C-6 proton. Taken together with other data, these results suggest that the enzyme exploits this substrate base in the enolization, which occurs through an intramolecular proton transfer. The loss of Pi then completes the beta-elimination.  相似文献   

10.
Rabbit muscle lactate dehydrogenase was subjected to frontal affinity chromatography on Sepharose-oxamate in the presence of various concentrations of NADH and sodium phosphate buffer (0.05 M, pH 6.8) containing 0.5 M-NaCl. Quantitative interpretation of the results yields an intrinsic association constant of 9.0 x 10 (4)M-1 for the interaction of enzyme with NADH at 5 degrees C, a value that is confirmed by equilibrium-binding measurements. In a second series of experiments, zonal affinity chromatography of a mouse tissue extract under the same conditions was used to evaluate assoication constants of the order 2 x 10(5)M-1, 3 x 10(5)M-1, 4 x 10(5)M-1, 7 x 10(5)M-1 and 2 x 10(6)M-1 for the interaction of NADH with the M4, M3H, M2H2, MH3 and H4 isoenzymes respectively of lactate dehydrogenase.  相似文献   

11.
Phosphite dehydrogenase (PTDH) catalyzes the unusual oxidation of phosphite to phosphate with the concomitant reduction of NAD(+) to NADH. PTDH shares significant amino acid sequence similarity with D-hydroxy acid dehydrogenases (DHs), including strongly conserved catalytic residues His292, Glu266, and Arg237. Site-directed mutagenesis studies corroborate the essential role of His292 as all mutants of this residue were completely inactive. Histidine-selective inactivation studies with diethyl pyrocarbonate provide further evidence regarding the importance of His292. This residue is most likely the active site base that deprotonates the water nucleophile. Kinetic analysis of mutants in which Arg237 was changed to Leu, Lys, His, and Gln revealed that Arg237 is involved in substrate binding. These results agree with the typical role of this residue in D-hydroxy acid DHs. However, Glu266 does not play the typical role of increasing the pK(a) of His292 to enhance substrate binding and catalysis as the Glu266Gln mutant displayed an increased k(cat) and unchanged pH-rate profile compared to those of wild-type PTDH. The role of Glu266 is likely the positioning of His292 and Arg237 with which it forms hydrogen bonds in a homology model. Homology modeling suggests that Lys76 may also be involved in substrate binding, and this postulate is supported by mutagenesis studies. All mutants of Lys76 display reduced activity with large effects on the K(m) for phosphite, and Lys76Cys could be chemically rescued by alkylation with 2-bromoethylamine. Whereas a positively charged residue is absolutely essential for activity at the position of Arg237, Lys76 mutants that lacked a positively charged side chain still had activity, indicating that it is less important for binding and catalysis. These results highlight the versatility of nature's catalytic scaffolds, as a common framework with modest changes allows PTDH to catalyze its unusual nucleophilic displacement reaction and d-hydroxy acid DHs to oxidize alcohols to ketones.  相似文献   

12.
Lad L  Mewies M  Raven EL 《Biochemistry》2002,41(46):13774-13781
The catalytic mechanism of recombinant soybean cytosolic ascorbate peroxidase (rsAPX) and a derivative of rsAPX in which a cysteine residue (Cys32) located close to the substrate (L-ascorbic acid) binding site has been modified to preclude binding of ascorbate [Mandelman, D., Jamal, J., and Poulos, T. L. (1998) Biochemistry 37, 17610-17617] has been examined using pre-steady-state and steady-state kinetic techniques. Formation (k1 = 3.3 +/- 0.1 x 10(7) M(-1) s(-1)) of Compound I and reduction (k(2) = 5.2 +/- 0.3 x 10(6) M(-1) s(-1)) of Compound I by substrate are fast. Wavelength maxima for Compound I of rsAPX (lambda(max) (nm) = 409, 530, 569, 655) are consistent with a porphyrin pi-cation radical. Reduction of Compound II by L-ascorbate is rate-limiting: at low substrate concentration (0-500 microM), kinetic traces were monophasic but above approximately 500 microM were biphasic. Observed rate constants for the fast phase overlaid with observed rate constants extracted from the (monophasic) dependence observed below 500 microM and showed saturation kinetics; rate constants for the slow phase were linearly dependent on substrate concentration (k(3-slow)) = 3.1 +/- 0.1 x 10(3) M(-1) s(-1)). Kinetic transients for reduction of Compound II by L-ascorbic acid for Cys32-modified rsAPX are monophasic at all substrate concentrations, and the second-order rate constant (k(3) = 0.9 +/- 0.1 x 10(3) M(-1) s(-1)) is similar to that obtained from the slow phase of Compound II reduction for unmodified rsAPX. Steady-state oxidation of L-ascorbate by rsAPX showed a sigmoidal dependence on substrate concentration and data were satisfactorily rationalized using the Hill equation; oxidation of L-ascorbic acid by Cys32-modified rsAPX showed no evidence of sigmoidal behavior. The data are consistent with the presence of two kinetically competent binding sites for ascorbate in APX.  相似文献   

13.
The mechanism of action of the flavoprotein D-aspartate oxidase (EC 1.4.3.1) has been investigated by steady-state and stopped flow kinetic studies using D-aspartate and O2 as substrates in 50 mM KPi, 0.3 mM EDTA, pH 7.4, 4 degrees C. Steady-state results indicate that a ternary complex containing enzyme, O2, and substrate (or product) is an obligatory intermediate in catalysis. The kinetic parameters are turnover number = 11.1 s-1, Km(D-Asp) = 2.2 x 10(-3) M, Km(O2) = 1.7 x 10(-4) M. Rapid reaction studies show that 1) the reductive half reaction is essentially irreversible with a maximum rate of reduction of 180 s-1; 2) the free reduced enzyme cannot be the species which is reoxidized during turnover since its reoxidation by oxygen (second order rate constant equal to 5.3 x 10(2) M-1 s-1) is too slow to be of relevance in catalysis; 3) reduced enzyme can bind a ligand rapidly and be reoxidized as a complex at a rate faster than that observed for the free reduced enzyme; 4) the rate of reoxidation of reduced enzyme by oxygen during turnover is dependent on both O2 and D-aspartate concentrations (second order rate constant of reaction between O2 and reduced enzyme-substrate complex equal to 6.2 x 10(4) M-1 s-1); and 5) the rate-limiting step in catalysis occurs after reoxidation of the enzyme and before its reduction in the following turnover. A mechanism involving reduction of enzyme by substrate, dissociation of product from reduced enzyme, binding of a second molecule of substrate to the reduced enzyme, and reoxidation of the reduced enzyme-substrate complex is proposed for the enzyme-catalyzed oxidation of D-aspartate.  相似文献   

14.
The kinetics of tissue plasminogen activator (t-PA) and DSPAalpha1-catalyzed plasminogen activation using untreated and TAFIa-treated fibrin degradation products (FDPs), ranging in weight average molecular weight (M(w)) from 0.48 x 10(6) to 4.94 x 10(6) g/mol, were modeled according to the steady-state template model. The FDPs served as effective cofactors for both activators. The intrinsic catalytic efficiencies of both t-PA (17.4 x 10(5) m(-1) s(-1)) and DSPAalpha1 (6.0 x 10(5) m(-1) s(-1)) were independent of FDP M(w). The intrinsic catalytic efficiency of t-PA was 12-fold higher than that measured under identical conditions with intact fibrin as the cofactor. At sub-saturating levels of cofactor and substrate, rates were strongly dependent on FDP M(w) with DSPAalpha1 but not t-PA. Loss of activity with decreasing FDP M(w) correlated with loss of finger-dependent binding of the activators to the FDPs. TAFIa treatment of the FDPs resulted in 90- and 215-fold decreases in the catalytic efficiencies of t-PA (0.20 x 10(5) m(-)(1) s(-1)) and DSPAalpha1 (0.028 x 10(5) m(-1) s(-1)), yielding cofactors that were still 30- and 50-fold better than fibrinogen with t-PA and DSPAalpha1, respectively. Our results show that for both activators the products released during fibrinolysis are very effective cofactors for plasminogen activation, and both t-PA and DSPAalpha1 cofactor activity are strongly down-regulated by TAFIa.  相似文献   

15.
The binding of heparin to human antithrombin III (ATIII) was investigated by titration calorimetry (TC) and differential scanning calorimetry (DSC). TC measurements of homogeneous high-affinity pentasaccharide and octasaccharide fragments of heparin in 0.02 M phosphate buffer and 0.15 M sodium chloride (pH 7.3) yielded binding constants of (7.1 +/- 1.3) x 10(5) M-1 and (6.7 +/- 1.2) x 10(6) M-1, respectively, and corresponding binding enthalpies of -48.3 +/- 0.7 and -54.4 +/- 5.4 kJ mol-1. The binding enthalpy of heparin in phosphate buffer (0.02 M, 0.15 M NaCl, pH 7.3) was estimated from TC measurements to be -55 +/- 10 kJ mol-1, while the enthalpy in Tris buffer (0.02 M, 0.15 M NaCl, pH 7.3) was -18 +/- 2 kJ mol-1. The heparin-binding affinity was shown by fluorescence measurements not to change under these conditions. The 3-fold lower binding enthalpy in Tris can be attributed to the transfer of a proton from the buffer to the heparin-ATIII complex. DSC measurements of the ATIII unfolding transition exhibited a sharp denaturation peak at 329 +/- 1 K with a van 't Hoff enthalpy of 951 +/- 89 kJ mol-1, based on a two-state transition model and a much broader transition from 333 to 366 K. The transition peak at 329 K accounted for 9-18% of the total ATIII. At sub-saturate heparin concentrations, the lower temperature peak became bimodal with the appearance of a second transition peak at 336 K. At saturate heparin concentration only the 336 K peak was observed. This supports a two domain model of ATIII folding in which the lower stability domain (329 K) binds and is stabilized by heparin.  相似文献   

16.
Glucose-6-phosphate dehydrogenase (G6PD) catalyses the first step of the pentose phosphate pathway which generates NADPH for anabolic pathways and protection systems in liver. G6PD was purified from dog liver with a specific activity of 130 U x mg(-1) and a yield of 18%. PAGE showed two bands on protein staining; only the slower moving band had G6PD activity. The observation of one band on SDS/PAGE with M(r) of 52.5 kDa suggested the faster moving band on native protein staining was the monomeric form of the enzyme.Dog liver G6PD had a pH optimum of 7.8. The activation energy, activation enthalpy, and Q10, for the enzymatic reaction were calculated to be 8.96, 8.34 kcal x mol(-1), and 1.62, respectively.The enzyme obeyed "Rapid Equilibrium Random Bi Bi" kinetic model with Km values of 122 +/- 18 microM for glucose-6-phosphate (G6P) and 10 +/- 1 microM for NADP. G6P and 2-deoxyglucose-6-phosphate were used with catalytic efficiencies (kcat/Km) of 1.86 x 10(6) and 5.55 x 10(6) M(-1) x s(-1), respectively. The intrinsic Km value for 2-deoxyglucose-6-phosphate was 24 +/- 4mM. Deamino-NADP (d-NADP) could replace NADP as coenzyme. With G6P as cosubstrate, Km d-ANADP was 23 +/- 3mM; Km for G6P remained the same as with NADP as coenzyme (122 +/- 18 microM). The catalytic efficiencies of NADP and d-ANADP (G6P as substrate) were 2.28 x 10(7) and 6.76 x 10(6) M(-1) x s(-1), respectively. Dog liver G6PD was inhibited competitively by NADPH (K(i)=12.0 +/- 7.0 microM). Low K(i) indicates tight enzyme:NADPH binding and the importance of NADPH in the regulation of the pentose phosphate pathway.  相似文献   

17.
Phosphite dehydrogenase (PTDH) from Pseudomonas stutzeri catalyzes the nicotinamide adenine dinucleotide-dependent oxidation of phosphite to phosphate. The enzyme belongs to the family of D-hydroxy acid dehydrogenases (DHDHs). A search of the protein databases uncovered many additional putative phosphite dehydrogenases. The genes encoding four diverse candidates were cloned and expressed, and the enzymes were purified and characterized. All oxidized phosphite to phosphate and had similar kinetic parameters despite a low level of pairwise sequence identity (39-72%). A recent crystal structure identified Arg301 as a residue in the active site that has not been investigated previously. Arg301 is fully conserved in the enzymes shown here to be PTDHs, but the residue is not conserved in other DHDHs. Kinetic analysis of site-directed mutants of this residue shows that it is important for efficient catalysis, with an ~100-fold decrease in k(cat) and an almost 700-fold increase in K(m,phosphite) for the R301A mutant. Interestingly, the R301K mutant displayed a slightly higher k(cat) than the parent PTDH, and a more modest increase in K(m) for phosphite (nearly 40-fold). Given these results, Arg301 may be involved in the binding and orientation of the phosphite substrate and/or play a catalytic role via electrostatic interactions. Three other residues in the active site region that are conserved in the PTDH orthologs but not DHDHs were identified (Trp134, Tyr139, and Ser295). The importance of these residues was also investigated by site-directed mutagenesis. All of the mutants had k(cat) values similar to that of the wild-type enzyme, indicating these residues are not important for catalysis.  相似文献   

18.
Malabanan MM  Go MK  Amyes TL  Richard JP 《Biochemistry》2011,50(25):5767-5779
Product yields for the reactions of (R)-glyceraldehyde 3-phosphate (GAP) in D2O at pD 7.9 catalyzed by wildtype triosephosphate isomerase from Trypanosoma brucei brucei (Tbb TIM) and a monomeric variant (monoTIM) of this wildtype enzyme were determined by (1)H NMR spectroscopy and were compared with the yields determined in earlier work for the reactions catalyzed by TIM from rabbit and chicken muscle [O'Donoghue, A. C., Amyes, T. L., and Richard, J. P. (2005), Biochemistry 44, 2610 - 2621]. Three products were observed from the reactions catalyzed by TIM: dihydroxyacetone phosphate (DHAP) from isomerization with intramolecular transfer of hydrogen, d-DHAP from isomerization with incorporation of deuterium from D2O into C-1 of DHAP, and d-GAP from incorporation of deuterium from D2O into C-2 of GAP. The yield of DHAP formed by intramolecular transfer of hydrogen decreases from 49% for the muscle enzymes to 40% for wildtype Tbb TIM to 34% for monoTIM. There is no significant difference in the ratio of the yields of d-DHAP and d-GAP for wildtype TIM from muscle sources and Trypanosoma brucei brucei, but partitioning of the enediolate intermediate of the monoTIM reaction to form d-DHAP is less favorable ((k(C1))(D)/(k(C2))(D) = 1.1) than for the wildtype enzyme ((k(C1))(D)/(k(C2))(D) = 1.7). Product yields for the wildtype Tbb TIM and monoTIM-catalyzed reactions of glycolaldehyde labeled with carbon-13 at the carbonyl carbon ([1-(13)C]-GA) at pD 7.0 in the presence of phosphite dianion and in its absence were determined by (1)H NMR spectroscopy [Go, M. K., Amyes, T. L., and Richard, J. P. (2009) Biochemistry 48, 5769-5778]. There is no detectable difference in the yields of the products of wildtype muscle and Tbb TIM-catalyzed reactions of [1-(13)C]-GA in D2O. The kinetic parameters for phosphite dianion activation of the reactions of [1-(13)C]-GA catalyzed by wildtype Tbb TIM are similar to those reported for the enzyme from rabbit muscle [Amyes, T. L. and Richard, J. P. (2007) Biochemistry 46, 5841-5854], but there is no detectable dianion activation of the reaction catalyzed by monoTIM. The engineered disruption of subunit contacts at monoTIM causes movement of the essential side chains of Lys-13 and His-95 away from the catalytic active positions. We suggest that this places an increased demand that the intrinsic binding energy of phosphite dianion be utilized to drive the change in the conformation of monoTIM back to the active structure for wildtype TIM.  相似文献   

19.
The values of K(i) for methylphosphate dianion (MP(2-)) inhibition of the cleavage of 2-hydroxypropyl-4-nitrophenyl phosphate (HpPNP) catalyzed by 1,3-bis(1,4,7-triazacyclonon-1-yl)-2-hydroxypropane (Zn(2)(1)(H(2)O)) approach a small limiting value of 6 microM at pH相似文献   

20.
Escherichia coli flavorubredoxin (FlRd) belongs to the family of flavodiiron proteins (FDPs), microbial enzymes that are expressed to scavenge nitric oxide (NO) under anaerobic conditions. To degrade NO, FlRd has to be reduced by NADH via the FAD-binding protein flavorubredoxin reductase, thus the kinetics of electron transfer along this pathway was investigated by stopped-flow absorption spectroscopy. We found that NADH, but not NADPH, quickly reduces the FlRd-reductase (k = 5.5 +/- 2.2 x 10(6) M(-1).s(-1) at 5 degrees C), with a limiting rate of 255 +/- 17 s(-1). The reductase in turn quickly reduces the rubredoxin (Rd) center of FlRd, as assessed at 5 degrees C working with the native FlRd enzyme (k = 2.4 +/- 0.1 x 10(6) m(-1).s(-1)) and with its isolated Rd-domain (k approximately 1 x 10(7) M(-1).s(-1)); in both cases the reaction was found to be dependent on pH and ionic strength. In FlRd the fast reduction of the Rd center occurs synchronously with the formation of flavin mononucleotide semiquinone. Our data provide evidence that (a) FlRd-reductase rapidly shuttles electrons between NADH and FlRd, a prerequisite for NO reduction in this detoxification pathway, and (b) the electron accepting site in FlRd, the Rd center, is in very fast redox equilibrium with the flavin mononucleotide.  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