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 78 毫秒
1.
Inosine 5'-monophosphate dehydrogenase (IMPDH) is the rate-limiting enzyme in de novo guanine nucleotide biosynthesis. IMPDH converts IMP to xanthosine 5'-monophosphate with concomitant conversion of NAD+ to NADH. All IMPDHs characterized to date contain a 130-residue "subdomain" that extends from an N-terminal loop of the alpha/beta barrel domain. The role of this subdomain is unknown. An IMPDH homolog has been cloned from Borrelia burgdorferi, the causative agent of Lyme disease (Margolis, N., Hogan, D., Tilly, K., and Rosa, P. A. (1994) J. Bacteriol. 176, 6427-6432). This homolog has replaced the subdomain with a 50-residue segment of unrelated sequence. We have expressed and characterized the B. burgdorferi IMPDH homolog. This protein has IMPDH activity, which unequivocally demonstrates that the subdomain is not required for catalytic activity. The monovalent cation and dinucleotide binding sites of B. burgdorferi IMPDH are significantly different from those of human IMPDH. Therefore, these sites are targets for the design of specific inhibitors for B. burgdorferi IMPDH. Such inhibitors might be new treatments for Lyme disease.  相似文献   

2.
Inhibitors of poly(ADP-ribose) polymerase (PARP, EC 2.4.2.30) are of clinical interest because they have potential for improving radiation therapy and chemotherapy of cancer. The refined binding structures of four such inhibitors are reported together with the refined structure of the unligated catalytic fragment of the enzyme. Following their design, all inhibitors bind at the position of the nicotinamide moiety of the substrate NAD+. The observed binding mode suggests inhibitor improvements that avoid other NAD(+)-binding enzymes. Because the binding pocket of NAD+ has been strongly conserved during evolution, the homology with ADP-ribosylating bacterial toxins could be used to extend the bound nicotinamide, which is marked by the inhibitors, to the full NAD+ molecule.  相似文献   

3.
The interaction of reduced nicotinamide mononucleotide (NMNH), constituting one half of NADH, with the wild-type and alphaD195E proton-pumping nicotinamide nucleotide transhydrogenase from Escherichia coli was investigated. Reduction of thio-NADP+ by NMNH was catalysed at approximately 30% of the rate with NADH. Other activities including proton pumping and the cyclic reduction of 3'-acetyl-pyridine-NAD+ by NMNH in the presence of NADP+ were more strongly inhibited. The alphaD195 residue is assumed to interact with the 2'-OH moiety of the adenosine-5'-phosphate, i.e., the second nucleotide of NADH. Mutation of this residue to alphaD195E resulted in a 90% decrease in activity with NMNH as well as NADH as substrate, suggesting that it produced global structural changes of the NAD(H) binding site. The results suggest that the NMN moiety of NADH is a substrate of transhydrogenase, and that the adenine nucleotide is not required for catalysis or proton pumping.  相似文献   

4.
BACKGROUND: The leucine biosynthetic enzyme 3-isopropylmalate dehydrogenase (IMDH) belongs to a unique class of bifunctional decarboxylating dehydrogenases. The two best-known members of this family, IMDH and isocitrate dehydrogenase (IDH), share a common structural framework and catalytic mechanism but have different substrate and cofactor specificities. IMDH is NAD(+)-dependent, while IDHs occur in both NAD(+)-dependent and NADP(+)-dependent forms. RESULTS: We have co-crystallized Thermus thermophilus IMDH with NAD+ and have determined the structure at 2.5 A resolution. NAD+ binds in an extended conformation. Comparisons with the structure in the absence of cofactor show that binding induces structural changes of up to 2.5 A in the five loops which form the dinucleotide-binding site. The adenine and diphosphate moieties of NAD+ are bound via interactions which are also present in the NADP(+)-IDH complex. Amino acids which interact with the NADP+ 2'-phosphate in IDH are substituted or absent in IMDH. The adenosine ribose forms two hydrogen bonds with Asp278, and the nicotinamide and nicotinamide ribose interact with Glu87 and Asp78, all unique to IMDH. CONCLUSIONS: NAD+ binding induces a conformational transition in IMDH, resulting in a structure that is intermediate between the most 'open' and 'closed' decarboxylating dehydrogenase conformations. Physiological specificity of IMDH for NAD+ versus NADP+ can be explained by the unique interaction between Asp278 and the free 2'-hydroxyl of the NAD+ adenosine, discrimination against the presence of the 2'-phosphate by the negative charge on Asp278, and the absence of potential favorable interactions with the 2'-phosphate of NADP+.  相似文献   

5.
Enterobacter cloacae nitroreductase (NR) is a flavoprotein which catalyzes the pyridine nucleotide-dependent reduction of nitroaromatics. Initial velocity and inhibition studies have been performed which establish unambiguously a ping-pong kinetic mechanism. NADH oxidation proceeds stereospecifically with the transfer of the pro-R hydrogen to the enzyme and the amide moiety of the nicotinamide appears to be the principal mediator of the interaction between NR and NADH. 2,4-Dinitrotoluene is the most efficient oxidizing substrate examined, with a kcat/KM an order of magnitude higher than those of p-nitrobenzoate, FMN, FAD or riboflavin. Dicoumarol is a potent inhibitor competitive vs. NADH with a Ki of 62 nM. Several compounds containing a carboxyl group are also competitive inhibitors vs. NADH. Yonetani-Theorell analysis of dicoumarol and acetate inhibition indicates that their binding is mutually exclusive, which suggests that the two inhibitors bind to the same site on the enzyme. NAD+ does not exhibit product inhibition and in the absence of an electron acceptor, no isotope exchange between NADH and 32P-NAD+ could be detected. NR catalyzes the 4-electron reduction of nitrobenzene to hydroxylaminobenzene with no optically detectable net formation of the putative two-electron intermediate nitrosobenzene.  相似文献   

6.
Human recombinant CD38 catalyzes the formation of both cyclic ADP-ribose and ADP-ribose products from NAD+ and hydrolyzes cyclic ADP-ribose to ADP-ribose. The corresponding GDP products are formed from NGD+. The enzyme was characterized by substrate and inhibition kinetics, exchange studies, rapid-quench reactions, and stopped-flow-fluorescence spectroscopy to establish the reaction mechanism and energetics for individual steps. Noncyclizable substrates NMN+ and nicotinamide-7-deaza-hypoxanthine dinucleotide (7-deaza NHD+) were rapidly hydrolyzed by the enzyme. The kcat for NMN+ was 5-fold higher than that of NAD+ and has the greatest reported kcat of any substrate for CD38. 7-deaza-NHD+ was hydrolyzed at approximately one-third the rate of NHD+ but does not form a cyclic product. These results establish that a cyclic intermediate is not required for substrate hydrolysis. The ratio of methanolysis to hydrolysis for cADPR and NAD+ catalyzed by CD38 increases linearly with MeOH concentration. Both reactions produce predominantly the beta-methoxy riboside compound, with a relative nucleophilicity of MeOH to H2O of 11. These results indicate the existence of a stabilized cationic intermediate for all observed chemistries in the active site of CD38. The partitioning of this intermediate between cyclization, hydrolysis, and nicotinamide-exchange unites the mechanisms of CD38 chemistries. Steady-state and pre-steady-state parameters for the partition and exchange mechanisms allowed full characterization of the reaction coordinate. Stopped-flow methods indicate a burst of cGDPR formation followed by the steady-state reaction rate. A lag phase, which was NGD+ concentration dependent, was also observed. The burst size indicates that the dimeric enzyme has a single catalytic site formed by two subunits. Pre-steady-state quench experiments did not detect covalent intermediates. Nicotinamide hydrolysis of NGD+ precedes cyclization and the chemical quench decomposes the enzyme-bound species to a mixture of cyclic and hydrolysis products. The time dependence of this ratio indicated that nicotinamide bond-breakage occurs 4 times faster than the conversion of the intermediate to products. Product release is the overall rate-limiting step for enzyme reaction with NGD+.  相似文献   

7.
IMP dehydrogenase (IMPDH) catalyzes the oxidation of IMP to XMP with the concomitant reduction of NAD+; the enzyme is activated by K+. This reaction is the rate-limiting step in de novo guanine nucleotide biosynthesis. In order to identify functionally important residues in IMPDH, including those involved in substrate and K+ binding, we have mutated 11 conserved Asp and Glu residues to Ala in Escherichia coli IMPDH. The values of kcat, Km, and Ki for GMP, XMP, mizoribine 5'-monophosphate (MMP), and beta-methylene-tiazofurin adenine dinucleotide (TAD) were determined. Five of these mutations caused a significant change (>/=10-fold) in one of these parameters. The Asp248 --> Ala mutation caused 100-fold decrease in the value of kcat and a 25-fold increase in the value of Kii for TAD; these observations suggest that Asp248 is in the NAD+ binding site. The Asp338 --> Ala mutation caused a 600-fold decrease in the value of kcat, but only a 5-10-fold increase in the values of Km for IMP and Kis for IMP analogs, suggesting that Asp338 may be involved in acid-base catalysis as well as IMP binding. The remaining three residues, Asp13, Asp50, and Glu469, appear to be involved in K+ activation; these residues may be ligands at one or more K+ binding sites. Interestingly, changes in the values of Ki for MMP correlate with changes in kcat/KmKm of IMPDH, while no such correlation is observed for GMP, XMP, and TAD. This observation indicates that MMP is a transition state analog for the IMPDH reaction.  相似文献   

8.
The NADH absorbance spectrum of nicotinoprotein (NADH-containing) alcohol dehydrogenase from Amycolatopsis methanolica has a maximum at 326 nm. Reduced enzyme-bound pyridine dinucleotide could be reversibly oxidized by acetaldehyde. The fluorescence excitation spectrum for NADH bound to the enzyme has a maximum at 325 nm. Upon excitation at 290 nm, energy transfer from tryptophan to enzyme-bound NADH was negligible. The fluorescence emission spectrum (excitation at 325 nm) for NADH bound to the enzyme has a maximum at 422 nm. The fluorescence intensity is enhanced by a factor of 3 upon binding of isobutyramide (Kd = 59 microM). Isobutyramide acts as competitive inhibitor (Ki = 46 microM) with respect to the electron acceptor NDMA (N,N-dimethyl-p-nitrosoaniline), which binds to the enzyme containing the reduced cofactor. The nonreactive substrate analogue trifluoroethanol acts as a competitive inhibitor with respect to the substrate ethanol (Ki = 1.6 microM), which binds to the enzyme containing the oxidized cofactor. Far-UV circular dichroism spectra of the enzyme containing NADH and the enzyme containing NAD+ were identical, indicating that no major conformational changes occur upon oxidation or reduction of the cofactor. Near-UV circular dichroism spectra of NADH bound to the enzyme have a minimum at 323 nm (Deltaepsilon = -8.6 M-1 cm-1). The fluorescence anisotropy decay of enzyme-bound NADH showed no rotational freedom of the NADH cofactor. This implies a rigid environment as well as lack of motion of the fluorophore. The average fluorescence lifetime of NADH bound to the enzyme is 0.29 ns at 20 degreesC and could be resolved into at least three components (in the range 0.13-0.96 ns). Upon binding of isobutyramide to the enzyme-containing NADH, the average excited-state lifetime increased to 1.02 ns and could be resolved into two components (0.37 and 1.11 ns). The optical spectra of NADH bound to nicotinoprotein alcohol dehydrogenase have blue-shifted maxima compared to other NADH-dehydrogenase complexes, but comparable to that observed for NADH bound to horse liver alcohol dehydrogenase. The fluorescence lifetime of NADH bound to the nicotinoprotein is very short compared to enzyme-bound NADH complexes, also compared to NADH bound to horse liver alcohol dehydrogenase. The cofactor-protein interaction in the nicotinoprotein alcohol dehydrogenase active site is more rigid and apolar than that in horse liver alcohol dehydrogenase. The optical properties of NADH bound to nicotinoprotein alcohol dehydrogenase differ considerably from NADH (tightly) bound to UDP-galactose epimerase from Escherichia coli. This indicates that although both enzymes have NAD(H) as nonexchangeable cofactor, the NADH binding sites are quite different.  相似文献   

9.
cis-Biphenyl-2,3-dihydrodiol-2,3-dehydrogenase (BphB) is involved in the aerobic biodegradation of polychlorinated biphenyls (PCBs). The crystal structure of the NAD+-enzyme complex was determined by molecular replacement and refined to an R-value of 17.9% at 2.0 A. As a member of the short-chain alcohol dehydrogenase/reductase (SDR) family, the overall protein fold and positioning of the catalytic triad in BphB are very similar to those observed in other SDR enzymes, although small differences occur in the cofactor binding site. Modeling studies indicate that the substrate is bound in a deep hydrophobic cleft close to the nicotinamide moiety of the NAD+ cofactor. These studies further suggest that Asn143 is a key determinant of substrate specificity. A two-step reaction mechanism is proposed for cis-dihydrodiol dehydrogenases.  相似文献   

10.
This paper reports the crystal structures of free acid and ammonium salt of adenosine 2'-monophosphate (2'-AMP). 2'-AMP crystallizes in the hexagonal space group P6(5)22 with a = 9.530(3) A, c = 73.422(2) A, and Z= 12. 2'-AMP.NH4 crystallizes in the trigonal space group P3(1) with a = 9.003(2) A, c = 34.743(2) A and Z= 6. Both the structures were solved by direct methods and refined by full matrix least- squares method to final R factors of 0.080 and 0.038 for 2'-AMP and 2'-AMP.NH4 respectively. The adenine bases of both the structures are in syn conformation contrasting with the anti geometry in 3'-AMP, 5'-AMP and the enzyme bound state. Ribose moiety of 2'-AMP is in C2' -endo conformation. However, the ribose moieties of both the nucleotide molecules display C2'-endo-C3'-exo twist conformation in 2'- AMP.NH4 structure. Both structures demonstrate g+ conformation about C4' -C5' bond. 2'-AMP and one of the nucleotide molecules of 2'-AMP.NH4 are protonated at N1 and the ammonium ion is involved in a bifurcated hydrogen bond with O3' B and O3A atoms. A characteristic feature of both the structures is the intramolecular O5' -N3 hydrogen bond. Our crystallographic results on 2'-AMP corroborates the earlier conclusion that the enzyme-bound state is not the lowest energy state of this nucleotide. 2' -AMP displays base-ribose 04' stacking not seen in the 2'-AMP.NH4 structure. Theoretical and experimental studies on 2'-, 3'- and 5'-AMP structures have been discussed.  相似文献   

11.
In an earlier study, Liu et al. (Bioorg. Med. Chem. Lett. 1992, 2, 1741-1744) showed that both the E and Z isomers of 4',5'-didehydro-5'-fluoroaristeromycin were very potent irreversible inhibitors of S-adenosylhomocysteine (AdoHcy) hydrolase. However, it was unclear from a mechanistic standpoint whether these vinyl fluorides were themselves type-I mechanism-based inhibitors causing reduction of enzyme-bound NAD+ or whether they were prodrug for aristeromycin-5'-carboxaldehyde, which was the ultimate type-I inhibitor. To elucidate this mechanism of enzyme inhibition, (4'S)- and (4'R)-aristeromycin-5'-carboxaldehydes (1a,b) were synthesized in this study and shown to be potent type-I mechanism-based inhibitors of AdoHcy hydrolase with k2/Ki values of 4.4 x 10(6) and 8.2 x 10(4)M-1min-1, respectively. However, Using 19F NMR and HPLC, it was shown that (4'S)-4,5'-dedehydro-5'-fluoraristeromycin in the presence of AdoHcy hydrolase did not release fluoride ion or generate aristeromycin-5'-carboxaldehyde (1a,b). These results suggest that the E and Z isomers of 4',5'-didehydro-5'-fluoroaristeromycin are inactivating AdoHcy hydrolase by directly reducing NAD+ to NADH and not using the hydrolytic activity of the enzyme to generate aristeromycin-5'-carboxaldehyde.  相似文献   

12.
BACKGROUND: 3-Isopropylmalate dehydrogenase (IPMDH) and isocitrate dehydrogenase (ICDH) belong to a unique family of bifunctional decarboxylating dehydrogenases. Although the ICDH dimer catalyzes its reaction under a closed conformation, known structures of the IPMDH dimer (without substrate) adopt a fully open or a partially closed form. Considering the similarity in the catalytic mechanism, the IPMDH dimer must be in a fully closed conformation during the reaction. A large conformational change should therefore occur upon substrate binding. RESULTS: We have determined the crystal structure of IPMDH from Thiobacillus ferrooxidans (Tf) complexed with 3-isopropylmalate (IPM) at 2.0 A resolution by the molecular replacement method. The structure shows a fully closed conformation and the substrate-binding site is quite similar to that of ICDH except for a region around the gamma-isopropyl group. The gamma group is recognized by a unique hydrophobic pocket, which includes Glu88, Leu91 and Leu92 from subunit 1 and Val193' from subunit 2. CONCLUSIONS: A large movement of domain 1 is induced by substrate binding, which results in the formation of the hydrophobic pocket for the gamma-isopropyl moiety of IPM. A glutamic acid in domain 1, Glu88, participates in the formation of the hydrophobic pocket. The C beta and C gamma atoms of Glu88 interact with the gamma-isopropyl moiety of IPM and are central to the recognition of substrate. The acidic tip of Glu88 is likely to interact with the nicotinamide mononucleotide (NMN) ribose of NAD+ in the ternary complex. This structure clearly explains the substrate specificity of IPMDH.  相似文献   

13.
The oxidation of alcohol to aldehyde by horse liver alcohol dehydrogenase (LADH) requires the transfer of a hydride ion from the alcohol substrate to the cofactor nicotinamide adenine dinucleotide (NAD). A quantum mechanical tunneling contribution to this hydride transfer step has been demonstrated in a number of LADH mutants designed to enhance or diminish this effect [Bahnson, B. J., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 12797-12802]. The active site double mutant Phe93 --> Trp/Val203 --> Ala shows a 75-fold reduction in catalytic efficiency relative to that of the native enzyme, and reduced tunneling relative to that of either single mutant. We present here two crystal structures of the double mutant: a 2.0 A complex with NAD and the substrate analogue trifluoroethanol and a 2.6 A complex with the isosteric NAD analogue CPAD and ethanol. Changes at the active site observed in both complexes are consistent with reduced activity and tunneling. The NAD-trifluoroethanol complex crystallizes in the closed conformation characteristic of the active enzyme. However, the NAD nicotinamide ring rotates away from the substrate, toward the space vacated by replacement of Val203 with the smaller alanine. Replacement of Phe93 with the larger tryptophan also produces unfavorable steric contacts with the nicotinamide carboxamide group, potentially destabilizing hydrogen bonds required to maintain the closed conformation. These contacts are relieved in the second complex by rotation of the CPAD pyridine ring into an unusual syn orientation. The resulting loss of the carboxamide hydrogen bonds produces an open conformation characteristic of the apoenzyme.  相似文献   

14.
UDP-galactose 4-epimerase plays a critical role in sugar metabolism by catalyzing the interconversion of UDP-galactose and UDP-glucose. Originally, it was assumed that the enzyme contained a "traditional" catalytic base that served to abstract a proton from the 4'-hydroxyl group of the UDP-glucose or UDP-galactose substrates during the course of the reaction. However, recent high-resolution X-ray crystallographic analyses of the protein from Escherichia coli have demonstrated the lack of an aspartate, a glutamate, or a histidine residue properly oriented within the active site cleft for serving such a functional role. Rather, the X-ray crystallographic investigation of the epimerase.NADH.UDP-glucose abortive complex from this laboratory has shown that both Ser 124 and Tyr 149 are located within hydrogen bonding distance to the 4'- and 3'-hydroxyl groups of the sugar, respectively. To test the structural role of Ser 124 in the reaction mechanism of epimerase, three site-directed mutant proteins, namely S124A, S124T, and S124V, were constructed and crystals of the S124A.NADH.UDP, S124A.NADH.UDP-glucose, S124T. NADH.UDP-glucose, and S124V.NADH.UDP-glucose complexes were grown. All of the crystals employed in this investigation belonged to the space group P3221 with the following unit cell dimensions: a = b = 83.8 A, c = 108.4 A, and one subunit per asymmetric unit. X-ray data sets were collected to at least 2.15 A resolution, and each protein model was subsequently refined to an R value of lower than 19.0% for all measured X-ray data. The investigations described here demonstrate that the decreases in enzymatic activities observed for these mutant proteins are due to the loss of a properly positioned hydroxyl group at position 124 and not to major tertiary and quaternary structural perturbations. In addition, these structures demonstrate the importance of a hydroxyl group at position 124 in stabilizing the anti conformation of the nicotinamide ring as observed in the previous structural analysis of the epimerase.NADH. UDP complex.  相似文献   

15.
Transhydrogenase is a proton pump. It has separate binding sites for NAD+/NADH (on domain I of the protein) and for NADP+/NADPH (on domain III). Purified, detergent-dispersed transhydrogenase from Escherichia coli catalyses the reduction of the NAD+ analogue, acetylpyridine adenine dinucleotide (AcPdAD+), by NADH at a slow rate in the absence of added NADP+ or NADPH. Although it is slow, this reaction is surprising, since transhydrogenase is generally thought to catalyse hydride transfer between NAD(H)--or its analogues and NADP(H)--or its analogues, by a ternary complex mechanism. It is shown that hydride transfer occurs between the 4A position on the nicotinamide ring of NADH and the 4A position of AcPdAD+. On the basis of the known stereospecificity of the enzyme, this eliminates the possibilities of transhydrogenation(a) from NADH in domain I to AcPdAD+ wrongly located in domain III; and (b) from NADH wrongly located in domain III to AcPdAD+ in domain I. In the presence of low concentrations of added NADP+ or NADPH, detergent-dispersed E. coli transhydrogenase catalyses the very rapid reduction of AcPdAD+ by NADH. This reaction is cyclic; it takes place via the alternate oxidation of NADPH by AcPdAD+ and the reduction of NADP+ by NADH, while the NADPH and NADP+ remain tightly bound to the enzyme. In the present work, it is shown that the rate of the cyclic reaction and the rate of reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH, have similar dependences on pH and on MgSO4 concentration and that they have a similar kinetic character. It is therefore suggested that the reduction of AcPdAD+ by NADH is actually a cyclic reaction operating, either with tightly bound NADP+/NADPH on a small fraction (< 5%) of the enzyme, or with NAD+/NADH (or AcPdAD+/AcPdADH) unnaturally occluded within the domain III site. Transhydrogenase associated with membrane vesicles (chromatophores) of Rhodospirillum rubrum also catalyses the reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH. When the chromatophores were stripped of transhydrogenase domain I, that reaction was lost in parallel with 'normal reverse' transhydrogenation (e.g., the reduction of AcPdAD+ by NADPH). The two reactions were fully recovered upon reconstitution with recombinant domain I protein. However, after repeated washing of the domain I-depleted chromatophores, reverse transhydrogenation activity (when assayed in the presence of domain I) was retained, whereas the reduction of AcPdAD+ by NADH declined in activity. Addition of low concentrations of NADP+ or NADPH always supported the same high rate of the NADH-->AcPdAD+ reaction independently of how often the membranes were washed. It is concluded that, as with the purified E. coli enzyme, the reduction of AcPdAD+ by NADH in chromatophores is a cyclic reaction involving nucleotides that are tightly bound in the domain III site of transhydrogenase. However, in the case of R. rubrum membranes it can be shown with some certainty that the bound nucleotides are NADP+ or NADPH. The data are thus adequately explained without recourse to suggestions of multiple nucleotide-binding sites on transhydrogenase.  相似文献   

16.
NADP-dependent enzymes. I: Conserved stereochemistry of cofactor binding   总被引:2,自引:0,他引:2  
The ubiquitous redox cofactors nicotinamide adenine dinucleotides [NAD and NADP] are very similar molecules, despite their participation in substantially different biochemical processes. NADP differs from NAD in only the presence of an additional phosphate group esterified to the 2'-hydroxyl group of the ribose at the adenine end and yet NADP is confined with few exceptions to the reactions of reductive biosynthesis, whereas NAD is used almost exclusively in oxidative degradations. The discrimination between NAD and NADP is therefore an impressive example of the power of molecular recognition by proteins. The many known tertiary structures of NADP complexes affords the possibility for an analysis of their discrimination. A systematic analysis of several crystal structures of NAD(P)-protein complexes show that: 1) the NADP coenzymes are more flexible in conformation than those of NAD; 2) although the protein-cofactor interactions are largely conserved in the NAD complexes, they are quite variable in those of NADP; and 3) in both cases the pocket around the nicotinamide moiety is substrate dependent. The conserved and variable interactions between protein and cofactors in the respective binding pockets are reported in detail. Discrimination between NAD and NADP is essentially a consequence of the overall pocket and not of a few residues. A clear fingerprint in NAD complexes is a carboxylate side chain that chelates the diol group at the ribose near the adenine, whereas in NADP complexes an arginine side chain faces the adenine plane and interacts with the phosphomonoester. The latter type of interaction might be a general feature of recognition of nucleotides by proteins. Other features such as strand-like hydrogen bonding between the NADP diphosphate moieties and the protein are also significant. The NADP binding pocket properties should prove useful in protein engineering and design.  相似文献   

17.
Alanine dehydrogenase [EC 1. 4. 1. 1] was purified to homogeneity from a crude extract of Enterobacter aerogenes ICR 0220. The enzyme had a molecular mass of about 245 kDa and consisted of six identical subunits. The enzyme showed maximal activity at about pH 10.9 for the deamination of L-alanine and at about pH 8.7 for the amination of pyruvate. The enzyme required NAD+ as a coenzyme. Analogs of NAD+, deamino-NAD+ and nicotinamide guanine dinucleotide served as coenzymes. Initial-velocity and product inhibition studies suggested that the deamination of L-alanine proceeded through a sequential ordered binary-ternary mechanism. NAD+ bound first to the enzyme, followed by L-alanine, and the products were released in the order of ammonia, pyruvate, and NADH. The Km were 0.47 mM for L-alanine, 0.16 mM for NAD+, 0.22 mM for pyruvate, 0.067 mM for NADH, and 66.7 mM for ammonia. The Km for L-alanine was the smallest in the alanine dehydrogenases studied so far. The enzyme gene was cloned into Escherichia coli JM109 cells and the nucleotides were sequenced. The deduced amino acid sequence was very similar to that of the alanine dehydrogenase from Bacillus subtilis. However, the Enterobacter enzyme has no cysteine residue. In this respect, the Enterobacter enzyme is different from other alanine dehydrogenases.  相似文献   

18.
Rat liver mitochondria have a specific Ca2+ release pathway which operates when NAD+ is hydrolysed to nicotinamide and ADPribose. NAD+ hydrolysis is Ca(2+)-dependent and inhibited by cyclosporine A (CSA). Mitochondrial Ca2+ release can be activated by the prooxidant t-butylhydroperoxide (tbh) or by gliotoxin (GT), a fungal metabolite of the epipolythiodioxopiperazine group. Tbh oxidizes NADH to NAD+ through an enzyme cascade consisting of glutathione peroxidase, glutathione reductase, and the energy linked transhydrogenase, whereas GT oxidizes some vicinal thiols to the disulfide form, a prerequisite for NAD+ hydrolysis. We report now that rat skeletal muscle mitochondria also contain a specific Ca2+ release pathway activated by both tbh and GT. Ca2+ release increases with the mitochondrial Ca2+ load, is completely inhibited in the presence of CSA, and is paralleled by pyridine nucleotide oxidation. In the presence of tbh and GT, mitochondria do not lose their membrane potential and do not swell, provided continuous release and re-uptake of Ca2+ ('Ca2+ cycling') is prevented. These data support the notion that both tbh- and GT-induced Ca2+ release are not the consequence of an unspecific increase of the inner membrane permeability ('pore' formation). Tbh induces Ca2+ release from rat skeletal muscle less efficiently than from liver mitochondria indicating that the coupling between tbh and NADH oxidation is much weaker in skeletal muscle mitochondria. This conclusion is corroborated by a much lower glutathione peroxidase activity in skeletal muscle than in liver mitochondria. The prooxidant-dependent pathway promotes, under drastic conditions (high mitochondrial Ca2+ loads and high tbh concentrations), Ca2+ release to about the same extent and rate as the Na+/Ca2+ exchanger. This renders the prooxidant-dependent pathway relevant in the pathophysiology of mitochondrial myopathies where its activation by an increased generation of reactive oxygen species probably results in excessive Ca2+ cycling and damage to mitochondria.  相似文献   

19.
Lactate dehydrogenase from the malarial parasite Plasmodium falciparum has many amino acid residues that are unique compared to any other known lactate dehydrogenase. This includes residues that define the substrate and cofactor binding sites. Nevertheless, parasite lactate dehydrogenase exhibits high specificity for pyruvic acid, even more restricted than the specificity of human lactate dehydrogenases M4 and H4. Parasite lactate dehydrogenase exhibits high catalytic efficiency in the reduction of pyruvate, kcat/Km = 9.0 x 10(8) min(-1) M(-1). Parasite lactate dehydrogenase also exhibits similar cofactor specificity to the human isoforms in the oxidation of L-lactate with NAD+ and with a series of NAD+ analogs, suggesting a similar cofactor binding environment in spite of the numerous amino acid differences. Parasite lactate dehydrogenase exhibits an enhanced kcat with the analog 3-acetylpyridine adenine dinucleotide (APAD+) whereas the human isoforms exhibit a lower kcat. This differential response to APAD+ provides the kinetic basis for the enzyme-based detection of malarial parasites. A series of inhibitors structurally related to the natural product gossypol were shown to be competitive inhibitors of the binding of NADH. Slight changes in structure produced marked changes in selectivity of inhibition of lactate dehydrogenase. 7-p-Trifluoromethylbenzyl-8-deoxyhemigossylic acid inhibited parasite lactate dehydrogenase, Ki = 0.2 microM, which was 65- and 400-fold tighter binding compared to the M4 and H4 isoforms of human lactate dehydrogenase. The results suggest that the cofactor site of parasite lactate dehydrogenase may be a potential target for structure-based drug design.  相似文献   

20.
Many Oriental people possess a liver mitochondrial aldehyde dehydrogenase where glutamate at position 487 has been replaced by a lysine, and they have very low levels of mitochondrial aldehyde dehydrogenase activity. To investigate the cause of the lack of activity of this aldehyde dehydrogenase, we mutated residue 487 of rat and human liver mitochondrial aldehyde dehydrogenase to a lysine and expressed the mutant and native enzyme forms in Escherichia coli. Both rat and human recombinant aldehyde dehydrogenases showed the same molecular and kinetic properties as the enzyme isolated from liver mitochondria. The E487K mutants were found to be active but possessed altered kinetic properties when compared to the glutamate enzyme. The Km for NAD+ at pH 7.4 increased more than 150-fold, whereas kcat decreased 2-10-fold with respect to the recombinant native enzymes. Detailed steady-state kinetic analysis showed that the binding of NAD+ to the mutant enzyme was impaired, and it could be calculated that this resulted in a decreased nucleophilicity of the active site cysteine residue. The rate-limiting step for the rat E487K mutant was also different from that of the recombinant rat liver aldehyde dehydrogenase in that no pre-steady-state burst of NADH formation was found with the mutant enzyme. Both the rat native enzyme and the E487K mutant oxidized chloroacetaldehyde twice as fast as acetaldehyde, indicating that the rate-limiting step was not hydride transfer or coenzyme dissociation but depended upon nucleophilic attack. Each enzyme form showed a 2-fold activation upon the addition of Mg2+ ions. Substituting a glutamine for the glutamate did not grossly affect the properties of the enzyme. Glutamate 487 may interact directly with the positive nicotinamide ring of NAD+ for the Ki of NADH was the same in the lysine enzyme as it was in the glutamate form. Because of the altered NAD+ binding properties and kcat of the E487K variant, it is assumed that people possessing this form will not have a functional mitochondrial aldehyde dehydrogenase.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号