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 191 毫秒
1.
2.
Glutamate dehydrogenase (L-glutamate:NAD+ oxidoreductase (deaminating); EC 1.4.1.2) has been purified from Peptostreptococcus asaccharolyticus in a single step using dye-ligand chromatography. The enzyme (GDH) was present in high yields and was stabilized in crude extracts. A subunit molecular weight of 49000 +/- 500 was determined by SDS polyacrylamide gel electrophoresis and six bands were obtained after cross-linking the subunits with dimethyl suberimidate. This bacterial GDH was predominantly NAD+-linked, but was able to utilize both NADP+ and NADPH at 4% of the rates with NAD+ and NADH, respectively. An investigation of the amino acid specificity revealed some similarities with GDH from mammalian sources and some clear differences. The values of apparent Km for the substrates ammonia, 2-oxoglutarate, NADH, NAD+ and glutamate were 18.4, 0.82, 0.066, 0.031 and 6 mM, respectively. The P. asaccharolyticus GDH was not regulated by purine nucleotides, but was subject to strong inhibition with increasing ionic strength.  相似文献   

3.
Markham GD  Bock CL  Schalk-Hihi C 《Biochemistry》1999,38(14):4433-4440
Inosine-5'-monophosphate dehydrogenase (IMPDH) catalyzes the K+-dependent reaction IMP + NAD + H2O --> XMP + NADH + H+ which is the rate-limiting step in guanine nucleotide biosynthesis. The catalytic mechanism of the human type-II IMPDH isozyme has been studied by measurement of the pH dependencies of the normal reaction, of the hydrolysis of 2-chloro-IMP (which yields XMP and Cl- in the absence of NAD), and of inactivation by the affinity label 6-chloro-purine-ribotide (6-Cl-PRT). The pH dependence of the IMPDH reaction shows bell-shaped profiles for kcat and the kcat/Km values for both IMP and NAD, illustrating the involvement of both acidic and basic groups in catalysis. Half-maximal kcat values occur at pH values of 7.2 and 9.8; similar pK values of 6.9 and 9.4 are seen in the kcat/Km profile for NAD. The kcat/Km profile for IMP, which binds first in the predominantly ordered kinetic mechanism, shows pK values of 8.1 and 7.3 for acidic and basic groups, respectively. None of the kinetic pK values correspond to ionizations of the free substrates and thus reflect ionization of the enzyme or enzyme-substrate complexes. The rate of inactivation by 6-Cl-PRT, which modifies the active site sulfhydryl of cysteine-331, increases with pH; the pK of 7.5 reflects the ionization of the sulfhydryl in the E.6-Cl-PRT complex. The pKs of the acids observed in the IMPDH reaction likely also reflect ionization of the cysteine-331 sulfhydryl which adds to C-2 of IMP prior to NAD reduction. The kcat and kcat/Km values for hydrolysis of 2-Cl-IMP show a pK value of 9.9 for a basic group, similar to that seen in the overall reaction, but do not exhibit the ionization of an acidic group. Surprisingly, the rates of 2-Cl-IMP hydrolysis and of inactivation by 6-Cl-PRT are not stimulated by K+, in contrast to the >100-fold K+ activation of the IMPDH reaction. Apparently the enigmatic role of K+ lies in the NAD(H)-dependent segment of the IMPDH reaction. To evaluate the importance of hydrogen bonding in substrate binding, several deamino- and deoxy-analogues of IMP were tested as substrates and inhibitors. Only 2'-deoxy-IMP was a substrate; the other compounds tested were competitive inhibitors with Ki values at most 10-fold greater than the KD for IMP, illustrating the greater importance of hydrogen-bonding interactions in the chemistry of the IMPDH reaction than simply in nucleotide binding.  相似文献   

4.
The amino acid sequence is reported for CNBr and tryptic peptide fragments of the NAD(+)-dependent glutamate dehydrogenase of Clostridium symbiosum. Together with the N-terminal sequence, these make up about 75% of the total sequence. The sequence shows extensive similarity with that of the NADP(+)-dependent glutamate dehydrogenase of Escherichia coli (52% identical residues out of the 332 compared) allowing confident placing of the peptide fragments within the overall sequence. This demonstrated sequence similarity with the E. coli enzyme, despite different coenzyme specificity, is much greater than the similarity (31% identities) between the GDH's of C. symbiosum and Peptostreptococcus asaccharolyticus, both NAD(+)-linked. The evolutionary implications are discussed. In the 'fingerprint' region of the nucleotide binding fold the sequence Gly X Gly X X Ala is found, rather than Gly X Gly X X Gly. The sequence found here has previously been associated with NADP+ specificity and its finding in a strictly NAD(+)-dependent enzyme requires closer examination of the function of this structural motif.  相似文献   

5.
An aldehyde derivative of riboflavin was covalently attached by reductive alkylation to soluble polycationic supports. The flavopolymers so obtained were stable under operational conditions. The catalytic efficiency towards oxidation of NADH by these flavopolymers was demonstrated, and the kinetic parameters (Km and kcat) revealed an overall catalytic efficiency (kcat/Km) 185-fold greater compared to riboflavin. Various factors affecting the chemical regeneration of NAD+ from NADH such as pH, ionic strength, nature of the buffer etc. were studied. The most interesting result was the highly favourable influence of borate ions which increased the reaction rate by a factor 2-4 compared to the other buffers. The flavopolymers are very effective for in situ recycling of NAD(P)+. With up to 300-fold NADH----NAD+ conversions for the system using yeast alcohol dehydrogenase and up to 1500-fold NADPH----NADP+ regenerations for the system using glucose-6-phosphate dehydrogenase. These flavopolymers are superior to previous chemical recycling systems.  相似文献   

6.
Substrate specificity of human mitochondrial low Km aldehyde dehydrogenase (EC 1.2.1.3) E2 isozyme has been investigated employing p-nitrophenyl esters of acyl groups of two to six carbon atoms and comparing with that of aldehydes of one to eight carbon atoms. The esterase reaction was studied under three conditions: in the absence of coenzyme, in the presence of NAD (1 mM), and in the presence of NADH (160 microM). The maximal velocity of the esterase reaction with p-nitrophenyl acetate and propionate as substrates in the presence of NAD was 3.9-4.7 times faster than that of the dehydrogenase reaction. Under all other conditions the velocities of dehydrogenase and esterase reactions were similar; the lowest kcat was for p-nitrophenyl butyrate in the presence of NAD. Stimulation of esterase activity by coenzymes was confined to esters of short acyl chain length; with longer acyl chain lengths or increased bulkiness (p-nitrophenyl guanidinobenzoate) no effect or even inhibition was observed. Comparison of kinetic constants for esters demonstrates that p-nitrophenyl butyrate is the worst substrate of all esters tested, suggesting that the active site topography is uniquely unfavorable for p-nitrophenyl butyrate. This fact is, however, not reflected in kinetic constants for butyraldehyde, which is a good substrate. The substrate specificity profile as determined by comparison of kcat/Km ratios was found to be quite different for aldehydes and esters. For aldehydes kcat/Km ratios increased with the increase of chain length; with esters under all three conditions, a V-shaped curve was produced with a minimum at p-nitrophenyl butyrate.  相似文献   

7.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

8.
Carrigan JB  Engel PC 《The FEBS journal》2007,274(19):5167-5174
Glutamate dehydrogenase (EC 1.4.1.2-4) from Peptostreptococcus asaccharolyticus has a strong preference for NADH over NADPH as a coenzyme, over 1000-fold in terms of kcat/Km values. Sequence alignments across the wider family of NAD(P)-dependent dehydrogenases might suggest that this preference is mainly due to a negatively charged glutamate at position 243 (E243) in the adenine ribose-binding pocket. We have examined the possibility of altering coenzyme specificity of the Peptostreptococcus enzyme, and, more specifically, the role of residue 243 and neighbouring residues in coenzyme binding, by introducing a range of point mutations. Glutamate dehydrogenases are unusual among dehydrogenases in that NADPH-specific forms usually have aspartate at this position. However, replacement of E243 with aspartate led to only a nine-fold relaxation of the strong discrimination against NADPH. By contrast, replacement with a more positively charged lysine or arginine, as found in NADPH-dependent members of other dehydrogenase families, allows a more than 1000-fold shift toward NADPH, resulting in enzymes equally efficient with NADH or NADPH. Smaller shifts in the same direction were also observed in enzymes where a neighboring tryptophan, W244, was replaced by a smaller alanine (approximately six-fold) or Asp245 was changed to lysine (32-fold). Coenzyme binding studies confirm that the mutations result in the expected major changes in relative affinities for NADH and NADPH, and pH studies indicate that improved affinity for the extra phosphate of NADPH is the predominant reason for the increased catalytic efficiency with this coenzyme. The marked difference between the results of replacing E243 with aspartate and with positive residues implies that the mode of NADPH binding in naturally occurring NADPH-dependent glutamate dehydrogenases differs from that adopted in E243K or E243D and in other dehydrogenases.  相似文献   

9.
An isocitrate dehydrogenase able to function with either NADP or NAD as coenzyme was purified to homogeneity from cell-free extracts of the purple photosynthetic eubacterium Rhodomicrobium vannielii using a rapid two-step procedure involving dye-ligand affinity chromatography. The enzyme was obtained in 60% yield with specific activities of 23 U.mg protein-1 (NADP-linked reaction) and 18.5 U.mg protein-1 (NAD-linked reaction). The purified enzyme was monomeric and migrated with an approximate Mr of 75,000-80,000 on both SDS/PAGE and non-denaturing PAGE. Affinity constants (Km values) of 2.5 microM for NADP and 0.77 mM for NAD and values for kcat/Km of 981,200 min-1.mM-1 (NADP) and 2455 min-1.mM-1 (NAD) indicated a greater specificity for NADP compared to NAD. A number of metabolites were examined for possible differential regulatory effects on the NADP- and NAD-linked reactions, using a dual-wavelength assay. Oxaloacetate was found to be an effective inhibitor of both reactions and the enzyme was also sensitive to concerted inhibition by glyoxylate and oxaloacetate. The amino-acid composition and the identity of 39 residues at the N-terminus were determined and compared to other isocitrate dehydrogenases. The results suggested a relationship between the Rm. vannielii enzyme and the monomeric isocitrate dehydrogenase isoenzyme II from Vibrio ABE-1.  相似文献   

10.
W B Strong  R Cook  V Schirch 《Biochemistry》1989,28(1):106-114
The dissociation constants of tetrahydropteroylpolyglutamates, having from one to six glutamate residues, have been determined for the two mitochondrial enzymes serine hydroxymethyltransferase and dimethylglycine dehydrogenase. The ratios of the dissociation constants for the mono- and hexaglutamate forms of the coenzyme were 200 and less than 10 for serine hydroxymethyltransferase and dimethylglycine dehydrogenase, respectively. Km and kcat values were determined for the reversible interconversion of serine and glycine as a function of the number of glutamyl residues on the coenzyme. The values in the serine to glycine direction did not significantly change with the number of glutamyl residues, but in the glycine to serine direction, there was a 9-fold increase in the kcat/Km when the longer chain polyglutamates were used as the coenzyme substrate. A sensitive and rapid method for determining the dissociation constants of proteins which bind either tetrahydropteroylpolyglutamates or their 5-methyl and 5-formyl conjugates is described.  相似文献   

11.
The NAD+-specific glutamate dehydrogenase from Peptostreptococcus asaccharolyticus follows Michaelis-Menten kinetics in contrast to the enzyme from several other sources, and thus gives linear double-reciprocal plots of initial-rate data. The initial-rate parameters have been determined for this bacterial dehyrogenase in the direction of oxidative deamination. The use of alternative coenzymes leads to some conclusions about the order of substrate addition. An investigation of the pH dependence of this reaction reveals that the binding of oxidised coenzyme is independent of pH over the range 6-9. The kinetic data are consistent with an ordered addition of coenzyme prior to glutamate, the reverse of the mechanism derived with ox glutamate dehydrogenase in the presence of ADP.  相似文献   

12.
T Sanner 《Biochemistry》1975,14(23):5094-5098
The reaction of glutamate dehydrogenase and glutamate (gl) with NAD+ and NADP+ has been studied with stopped-flow techniques. The enzyme was in all experiments present in excess of the coenzyme. The results indicate that the ternary complex (E-NAD(P)H-kg) is present as an intermediate in the formation of the stable complex (E-NAD(P)H-gl). The identification of the complexes is based on their absorption spectra. The binding of the coenzyme to (E-gl) is the rate-limiting step in the formation of (E-NAD(P)H-kg) while the dissociation of alpha-ketoglutarate (kg) from this complex is the rate-limiting step in the formation of (E-NAD(P)H-gl). The Km for glutamate was 20-25 mM in the first reaction and 3 mM in the formation of the stable complex. The Km values were independent of the coenzyme. The reaction rates with NAD+ were approximately 50% greater than those with NADP+. Furthermore, high glutamate concentration inhibited the formation of (E-NADH-kg) while no substrate inhibition was found with NADP+ as coenzyme. ADP enhanced while GTP reduced the rate of (E-NAD(P)H-gl) formation. The rate of formation of (E-NAD(P)H-kg) was inhibited by ADP, while it increased at high glutamate concentration when small amounts of GTP were added. The results show that the higher activity found with NAD+ compared to NADP+ under steady-state assay conditions do not necessarily involve binding of NAD+ to the ADP activating site of the enzyme. Moreover, the substrate inhibition found at high glutamate concentration under steady-state assay condition is not due to the formation of (E-NAD(P)H-gl) as this complex is formed with Km of 3 mM glutamate, and the substrate inhibition is only significant at 20-30 times this concentration.  相似文献   

13.
The proposal that the active site vacuole of NAD(+)-S-lactate dehydrogenase is unable to accommodate any imbalance in electrostatic charge was tested by genetically manipulating the cDNA coding for human muscle lactate dehydrogenase to make a protein with an aspartic acid introduced at position 140 instead of the wild-type asparagine. The Asn 140-Asp mutant enzyme has the same kcat as the wild type (Asn 140) at low pH (4.5), and at higher pH the Km for pyruvate increases 10-fold for each unit increase in pH up to pH 9. We conclude that the anion of Asp 140 is completely inactive and that it binds pyruvate with a Km that is over 1,000 times that of the Km of the neutral, protonated aspartic-140. Experimental results and molecular modeling studies indicate the pKa of the active site histidine-195 in the enzyme-NADH complex is raised to greater than 10 by the presence of the anion at position 140. Energy minimization and molecular dynamics studies over 36 ps suggest that the anion at position 140 promotes the opening of and the entry of mobile solvent beneath the polypeptide loop (98-110), which normally seals off the internal active site vacuole from external bulk solvent.  相似文献   

14.
The role of three histidine residues (His205, His296 and His303) and Asp259, important for the catalysis of NAD+-specific D-lactate dehydrogenase, was investigated using site-directed mutagenesis. None of these residues is presumed to be involved in coenzyme binding because Km for NADH remained essentially unchanged for all the mutant enzymes. Replacement of His205 with lysine resulted in a 125-fold reduction in kcat and a slight lowering of the Km value for pyruvate. D259N mutant showed a 56-fold reduction in kcat and a fivefold lowering of Km. The enzymatic activity profile shifted towards acidic pH by approximately 2 units. The H303K mutation produced no significant change in kcat values, although Km for pyruvate increased fourfold. Substitution of His296 with lysine produced no significant change in kcat values or in Km for substrate. The results obtained suggest that His205 and Asp259 play an important role in catalysis, whereas His303 does not. This corroborates structural information available for some members of the D-specific dehydrogenases family. The catalytic His296, proposed from structural studies to be the active site acid/base catalyst, is not invariant. Its function can be accomplished by lysine and this has significant implications for the enzymatic mechanism.  相似文献   

15.
Recently, we amplified the Lactobacillus bulgaricus NAD(+)-dependent D-lactate dehydrogenase gene by the polymerase chain reaction, cloned and overexpressed it in Escherichia coli (Kochhar, S., Chuard, N., and Hottinger, H. (1992) Biochem. Biophys. Res. Commun. 185, 705-712). Polymerase chain reaction-amplified DNA fragments may contain base changes resulting in mutant gene products. A comparison of specific activities of D-lactate dehydrogenase in the crude extracts of 50 recombinant clones indicated that one of the clones had drastically reduced enzyme activity. Nucleotide sequence analysis of the insert DNA showed an exchange of A to G at position 795 resulting in substitution of Glu264 to Gly in the D-lactate dehydrogenase. The purified mutant D-lactate dehydrogenase showed a shift of 2 units in its optimum pH toward the acidic range. The dependence of kcat/Km on the pH of the mutant enzyme showed that the pKa of the free enzyme was around 4, at least 2 pH units lower than that of the wild-type enzyme. Both the wild-type and the mutant enzyme at their respective optimum pH values showed similar kcat and Km values. The data suggest that the highly conserved Glu264 is not critical for enzyme catalysis, but it must be situated within hydrogen bonding distance to amino acid residue(s) involved in substrate binding as well as in catalysis.  相似文献   

16.
Exposure of Neuro-2a and PC12 cells to micromolar concentrations of sulfite caused an increase in reactive oxygen species and a decrease in ATP. Likewise, the biosynthesis of ATP in intact rat brain mitochondria from the oxidation of glutamate was inhibited by micromolar sulfite. Glutamate-driven respiration increased the mitochondrial membrane potential (MMP), and this was abolished by sulfite but the MMP generated by oxidation of malate and succinate was not affected. The increased rate of production of NADH from exogenous NAD+ and glutamate added to rat brain mitochondrial extracts was inhibited by sulfite, and mitochondria preincubated with sulfite failed to reduce NAD+. Glutamate dehydrogenase (GDH) in rat brain mitochondrial extract was inhibited dose-dependently by sulfite as was the activity of a purified enzyme. An increase in the Km (glutamate) and a decrease in Vmax resulting in an attenuation in Vmax/Km (glutamate) at 100 microm sulfite suggest a mixed type of inhibition. However, uncompetitive inhibition was noted with decreases in both Km (NAD+) and Vmax, whereas Vmax/Km (NAD+) remained relatively constant. We propose that GDH is one target of action of sulfite, leading to a decrease in alpha-ketoglutarate and a diminished flux through the tricarboxylic acid cycle accompanied by a decrease in NADH through the mitochondrial electron transport chain, a decreased MMP, and a decrease in ATP synthesis. Because glutamate is a major metabolite in the brain, inhibition of GDH by sulfite could contribute to the severe phenotype of sulfite oxidase deficiency in human infants.  相似文献   

17.
Drosophila alcohol dehydrogenase (ADH), an NAD(+)-dependent dehydrogenase, shares little sequence similarity with horse liver ADH. However, these two enzymes do have substantial similarity in their secondary structure at the NAD(+)-binding domain [Benyajati, C., Place, A. P., Powers, D. A. & Sofer, W. (1981) Proc. Natl Acad. Sci. USA 78, 2717-2721]. Asp38, a conserved residue between Drosophila and horse liver ADH, appears to interact with the hydroxyl groups of the ribose moiety in the AMP portion of NAD+. A secondary-structure comparison between the nucleotide-binding domain of NAD(+)-dependent enzymes and that of NADP(+)-dependent enzymes also suggests that Asp38 could play an important role in cofactor specificity. Mutating Asp38 of Drosophila ADH into Asn38 decreases Km(app)NADP 62-fold and increases kcat/Km(app)NADP 590-fold at pH 9.8, when compared with wild-type ADH. These results suggest that Asp38 is in the NAD(+)-binding domain and its substituent, Asn38, allows Drosophila ADH to use both NAD+ and NADP+ as its cofactor. The observations from the experiments of thermal denaturation and kinetic measurement with pH also confirm that the repulsion between the negative charges of Asp38 and 2'-phosphate of NADP+ is the major energy barrier for NADP+ to serve as a cofactor for Drosophila ADH.  相似文献   

18.
Coenzyme specificity of mammalian liver D-glycerate dehydrogenase   总被引:1,自引:0,他引:1  
D-Glycerate dehydrogenase (glyoxylate reductase) was partially purified from rat liver by anion- and cation-exchange chromatography. When assayed in the direction of D-glycerate or glycolate formation, the enzyme was inhibited by high (greater than or equal to 0.5 mM), unphysiological concentrations of hydroxypyruvate or glyoxylate much more potently in the presence of NADPH than in the presence of NADH. However, the dehydrogenase displayed a much greater affinity for NADPH (Km less than 1 microM) than for NADH (Km = 48-153 microM). Furthermore, NADP was over 1000-fold more potent than NAD in inhibiting the enzyme competitively with respect to NADH. NADP also inhibited the reaction competitively with respect to NADPH whereas NAD, at concentrations of up to 10 mM had no inhibitory effect. When measured by the formation of hydroxypyruvate from D-glycerate, the enzyme also displayed a much greater affinity for NADP than for NAD. These properties indicate that liver D-glycerate dehydrogenase functions physiologically as an NADPH-specific reductase. In agreement with this conclusion, the addition of hydroxypyruvate or glyoxylate to suspensions of rat hepatocytes stimulated the pentose-phosphate pathway. The coenzyme specificity of D-glycerate dehydrogenase is discussed in relation to the biochemical findings made in D-glyceric aciduria and in primary hyperoxaluria type II (L-glyceric aciduria).  相似文献   

19.
Mitochondrial aldehyde dehydrogenase (ALDH2) is the major enzyme that oxidizes ethanol-derived acetaldehyde. A nearly inactive form of the enzyme, ALDH2*2, is found in about 40% of the East Asian population. This variant enzyme is defined by a glutamate to lysine substitution at residue 487 located within the oligomerization domain. ALDH2*2 has an increased Km for its coenzyme, NAD+, and a decreased kcat, which lead to low activity in vivo. Here we report the 2.1 A crystal structure of ALDH2*2. The structure shows a large disordered region located at the dimer interface that includes much of the coenzyme binding cleft and a loop of residues that form the base of the active site. As a consequence of these structural changes, the variant enzyme exhibits rigid body rotations of its catalytic and coenzyme-binding domains relative to the oligomerization domain. These structural perturbations are the direct result of the inability of lysine 487 to form important stabilizing hydrogen bonds with arginines 264 and 475. Thus, the elevated Km for coenzyme exhibited by this variant probably reflects the energetic penalty for reestablishing this site for productive coenzyme binding, whereas the structural alterations near the active site are consistent with the lowered Vmax.  相似文献   

20.
A circularly permuted (cp) variant of the phosphorylating NAD-dependent glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from Bacillus stearothermophilus has been constructed with N- and C-termini created within the coenzyme binding domain. The cp variant has a kcat value equal to 40% of the wild-type value, whereas Km and KD values for NAD show a threefold decrease compared to wild type. These results indicate that the folding process and the conformational changes that accompany NAD binding during the catalytic event occur efficiently in the permuted variant and that NAD binding is tighter. Reversible denaturation experiments show that the stability of the variant is only reduced by 0.7 kcal/mol compared to the wild-type enzyme. These experiments confirm and extend results obtained recently on other permuted proteins. For multimeric proteins, such as GAPDH, which harbor subunits with two structural domains, the natural location of the N- and C-termini is not a prerequisite for optimal folding and biological activity.  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