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.
Abstract Cell-free extracts of the photosynthetic eubacterium Rhodomicrobium vannielii contained both NADP and NAD-linked isocitrate dehydrogenase activities. Apparent K m values of 12 μM for NADP, 0.75 mM for NAD, 9.3 μM for isocitrate (NADP utilising) and 8.2 μM for isocitrate (NAD utilising) were determined in such extracts. Four lines of evidence indicated that one enzyme was responsible for the two activities; (i) non-additivity of reaction rates in the presence of both NADP and NAD (ii) the presence of one band which stained for activity with both cofactors on non-denaturing polyacrylamide gels (iii) identical heat inactivation kinetics for both activities (iv) co-elution of both activities after ion-exchange and hydrophobic interaction chromatography. This is the first report of a eubacterial isocitrate dehydrogenase with dual cofactor specificity.  相似文献   

2.
Lin AP  McAlister-Henn L 《Biochemistry》2011,50(38):8241-8250
Yeast NAD(+)-specific isocitrate dehydrogenase is an allosterically regulated octameric enzyme composed of four heterodimers of a catalytic IDH2 subunit and a regulatory IDH1 subunit. Despite structural predictions that the enzyme would contain eight isocitrate binding sites, four NAD(+) binding sites, and four AMP binding sites, only half of the sites for each ligand can be measured in binding assays. On the basis of a potential interaction between side chains of Cys-150 residues in IDH2 subunits in each tetramer of the enzyme, ligand binding assays of wild-type (IDH1/IDH2) and IDH1/IDH2(C150S) octameric enzymes were conducted in the presence of dithiothreitol. These assays demonstrated the presence of eight isocitrate and four AMP binding sites for the wild-type enzyme in the presence of dithiothreitol and for the IDH1/IDH2(C150S) enzyme in the absence or presence of this reagent, suggesting that interactions between sulfhydryl side chains of IDH2 Cys-150 residues limit access to these sites. However, only two NAD(+) sites could be measured for either enzyme. A tetrameric form of IDH (an IDH1(G15D)/IDH2 mutant enzyme) demonstrated half-site binding for isocitrate (two sites) in the absence of dithiothreitol and full-site binding (four sites) in the presence of dithiothreitol. Only one NAD(+) site could be measured for the tetramer under both conditions. In the context of the structure of the enzyme, these results suggest that an observed asymmetry between heterotetramers in the holoenzyme contributes to interactions between IDH2 Cys-150 residues and to half-site binding of isocitrate, but that a form of negative cooperativity may limit access to apparently equivalent NAD(+) binding sites.  相似文献   

3.
4.
K M Weeks  D M Crothers 《Biochemistry》1992,31(42):10281-10287
RNA recognition by the HIV Tat protein is mediated in part by an arginine- and lysine-rich basic subdomain implicated as a signature element in proteins that bind RNA. Relative RNA binding affinities for a 14-residue peptide derived from Tat that spans the basic region are determined using a competition protocol. Binding specificity is compared with complexation by a 38-residue model for the RNA binding domain of Tat using the same approach. Binding strength for the minimal (14 residue) peptide is correlated with that for the longer peptide: both peptides recognize a short, bulged duplex. However, the shorter peptide dissociates more rapidly from the wild-type site and discriminates less well between nonspecific (double-stranded RNA) and specific sites. Relative dissociation constants for 38-residue peptide determined from direct partition and competition assays differ; the former assay consistently predicts stronger discrimination against RNAs with mutations in the stems flanking the bulge. Differences between the two assays are reconciled in terms of contributions from labile binding which is unstable to native gel electrophoresis. Kinetic stability may constitute a major specificity determinant for basic subdomain-mediated recognition of RNA.  相似文献   

5.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short chain dehydrogenase/reductase (SDR) family, is responsible for the biological inactivation of prostaglandins. Sequence alignment within SDR coupled with molecular modeling analysis has suggested that Gln-15, Asp-36, and Trp-37 of 15-PGDH may determine the coenzyme specificity of this enzyme. Site-directed mutagenesis was used to examine the important roles of these residues. Several single mutants (Q15K, Q15R, W37K, and W37R), double mutants (Q15K-W37K, Q15K-W37R, Q15R-W37K, and Q15R-W37R), and triple mutants (Q15K-D36A-W37R and Q15K-D36S-W37R) were prepared and expressed as glutathione S-transferase (GST) fusion proteins in Escherichia coli and purified by GSH-agarose affinity chromatography. Mutants Q15K, Q15R, W37K, W37R, Q15K-W37K, and Q15R-W37K were found to be inactive or almost inactive with NADP+ but still retained substantial activity with NAD+. Mutant Q15K-W37R and mutant Q15R-W37R showed comparable activity for NAD+ and NADP+ with an increase in activity nearly 3-fold over that of the wild type. However, approximately 30-fold higher in K(m) for NADP+ than that of the wild type enzyme for NAD+ was found for mutants Q15K-W37R and Q15R-W37R. Similarly, the K(m) values for PGE(2) of mutants were also shown to increase over that of the wild type. Further mutation of Asp-36 to either an alanine or a serine of the double mutant Q15K-W37R (i.e., triple mutants Q15K-D36A-W37R and Q15K-D36S-W37R) rendered the mutants exhibiting exclusive activity with NADP+ but not with NAD+. The triple mutants showed a decrease in K(m) for NADP+ but an increase in K(m) for PGE(2). Further mutation at Ala-14 to a serine of a triple mutant (Q15K-D36S-W37R) decreased the K(m) values for both NADP+ and PGE(2) to levels comparable to those of the wild type. These results indicate that the coenzyme specificity of 15-PGDH can be altered from NAD+ to NADP+ by changing a few critical residues near the N-terminal end.  相似文献   

6.
A series of NAD+ analogues, modified on the pyridinium ring, have been tested for their enzymic properties in reactions with D-glyceraldehyde-3-phosphate dehydrogenase form sturgeon muscle, rabbit muscle and Bacillus stearothermophilus. The observed activity, inhibition and binding data are correlated to the structure of the enzyme and coenzyme analogue by model building on a Vector General interactive graphic display system using coordinates from the B. stearothermophilus holoenzyme structure. Most of the analogues with substituents in the pyridinium-3 position could be bound to glyceraldehyde-3-phosphate dehydrogenase, either in manner similar to NAD+ or in a completely different way with the substituted pyridinium ring rotated 110 degrees or more around the glycosidic bond. This indicates different possible modes of binding of NAD+ analogues within the pyridinium binding subsite. Analogues with substituents in the pyridinium-4 position are shown to be weakly bound to glyceraldehyde-3-phosphate dehydrogenase. This is explained by a strong interaction of the substituent in the 4 position with the residues Asn-313 and Cys-149.  相似文献   

7.
8.
The fluorescence of the natural coenzyme, NADH, is used to monitor the environment of the nicotinamide moiety at the active centre of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12). Changes of the fluorescence quantum yield and polarization of a small amount of NADH, totally bound by an excess of enzyme, show that at half-saturation of the oligomer with NAD a conformational change is induced which affects the active centre regions of the remaining subunits. This conformational transition is not effected by adenosine diphosphoribose, suggesting that the binding of the nicotinamide moiety of NAD to two subunits is essential for the change of tertiary structure of the remaining subunits that causes the observed changes of the fluorescence properties of the ADH "tracer probe". It is suggested that this conformational transition of the oligomer is responsible for the major decrease of affinity for NAD which occurs at half-saturation, and possibly for the activation by NAD+ of the reductive dephosphorylation reaction catalysed by the enzyme. It is also suggested, by analogy with haemoglobin, that the molecular basis of the negative cooperativity may be the creation of additional intersubunit bonds during the binding of the first two NAD molecules to the tetramer, and a change from a "relaxed" quaternary structure to a "tense" structure at half-saturation.  相似文献   

9.
Antibodies HyHEL8, HyHEL10, and HyHEL26 (HH8, HH10, and HH26, respectively) recognize highly overlapping epitopes on hen egg-white lysozyme (HEL) with similar affinities, but with different specificities. HH8 binding to HEL is least sensitive toward mutations in the epitope and thus is most cross-reactive, HH26 is most sensitive, whereas the sensitivity of HH10 lies in between HH8 and HH26. Here we have investigated intra- and intermolecular interactions in three antibody-protein complexes: theoretical models of HH8-HEL and HH26-HEL complexes, and the x-ray crystal structure of HH10-HEL complex. Our results show that HH8-HEL has the lowest number and HH26-HEL has the highest number of intra- and intermolecular hydrogen bonds. The number of salt bridges is lowest in HH8-HEL and highest in HH26-HEL. The binding site salt bridges in HH8-HEL are not networked, and are weak, whereas, in HH26-HEL, an intramolecular salt-bridge triad at the binding site is networked to an intermolecular triad to form a pentad. The pentad and each salt bridge of this pentad are exceptionally stabilizing. The number of binding-site salt bridges and their strengths are intermediate in HH10-HEL, with an intramolecular triad. Our further calculations show that the electrostatic component contributes the most to binding energy of HH26-HEL, whereas the hydrophobic component contributes the most in the case of HH8-HEL. A "hot-spot" epitope residue Lys-97 forms an intermolecular salt bridge in HH8-HEL, and participates in the intermolecular pentad in the HH26-HEL complex. Mutant modeling and surface plasmon resonance (SPR) studies show that this hot-spot epitope residue contributes significantly more to the binding than an adjacent epitope residue, Lys-96, which does not form a salt bridge in any of the three HH-HEL complexes. Furthermore, the effect of mutating Lys-97 is most severe in HH26-HEL. Lys-96, being a charged residue, also contributes the most in HH26-HEL among the three complexes. The SPR results on these mutants also highlight that the apparent "electrostatic steering" on net on rates actually act at post-collision level stabilization of the complex. The significance of this work is the observed variations in electrostatic interactions among the three complexes. Our work demonstrates that higher electrostatics, both as a number of short-range electrostatic interactions and their contributions, leads to higher binding specificity. Strong salt bridges, their networking, and electrostatically driven binding, limit flexibilities through geometric constrains. In contrast, hydrophobic driven binding and low levels of electrostatic interactions are associated with conformational flexibility and cross-reactivity.  相似文献   

10.
Changes induced on addition of the coenzyme, NADH or NAD+, to porcine lactic dehydrogenase isoenzymes H4 and M4 have been studied by hydrodynamic and spectroscopic methods. As shown by ultracentrifugal analysis, the native subunit structure remains unchanged on holoenzyme formation; a 5% increase of the sedimentation coefficient, parallelled by a slight decrease of the partial specific volume (<1%) indicate a significant change in the native tertiary and/or quaternary structure of the enzymes, corroborating earlier calorimetric data (Hinz and Jaenicke, 1975). The binding constant for the enzyme from skeletal muslce (M4) and NADH are found to be in agreement with K D-values obtained from equilibrium dialysis, as well as spectroscopic and thermal titration experiments (8 M). Far UV circular dichroism measurements do not show significant changes on ligand binding, indicating unchanged helicity or compensatory conformational effects. In the near UV, ligand binding is reflected by an extrinsic Cotton effect around 340 nm; in the range of aromatic absorption no changes are detectable.The experimental results suggest that there are gross structural changes on coenzyme binding to lactic dehydrogenase which do not affect the intrinsic spectral properties normally applied to analyze transconformation reactions in protein molecules.  相似文献   

11.
The kinetics of LDH-catalyzed reduction of pyruvate involving APADH were studied. It was shown that under conditions of a single turnover reaction the first order rate constant is equal to 37+/-4 sec-1. The reaction rate (vo) did not change when a deutero-coenzyme was used. The relationship between vo and pyruvate concentration is hyperbolic. It is concluded that isomerization of the ternary LDH-APADH-pyruvate complex limits the reaction rate. The spectral properties and the kinetics of formation and dissociation of abortive LDH complexes with pyruvate and NAD analogs (APAD and PAAD) were studied. The participation of the carboxamide group of NAD in conformational isomerization of the LDH-NADH-pyruvate and LDH-NAD-pyruvate complexes was studied.  相似文献   

12.
Transhydrogenase (TH) couples direct and stereospecific hydride transfer between NAD(H) and NADP(H), bound within soluble domains I and III, respectively, to proton translocation across membrane bound domain II. The cocrystal structure of Rhodospirillum rubrum TH domains I and III has been determined in the presence of limiting NADH, under conditions in which the subunits reach equilibrium during crystallization. The crystals contain three heterotrimeric complexes, dI(2)dIII, in the asymmetric unit. Multiple conformations of loops and side-chains, and NAD(H) cofactors, are observed in domain I pertaining to substrate/product exchange, and highlighting electrostatic interactions during the hydride transfer. Two interacting NAD(H)-NADPH pairs are observed where alternate conformations of the NAD(H) phosphodiester and conserved arginine side-chains are correlated. In addition, the stereochemistry of one NAD(H)-NADPH pair approaches that expected for nicotinamide hydride transfer reactions. The cocrystal structure exhibits non-crystallographic symmetry that implies another orientation for domain III, which could occur in dimeric TH. Superposition of the "closed" form of domain III (PDB 1PNO, chain A) onto the dI(2)dIII complex reveals a severe steric conflict of highly conserved loops in domains I and III. This overlap, and the overlap with a 2-fold related domain III, suggests that motions of loop D within domain III and of the entire domain are correlated during turnover. The results support the concept that proton pumping in TH is driven by the difference in binding affinity for oxidized and reduced nicotinamide cofactors, and in the absence of a difference in redox potential, must occur through conformational effects.  相似文献   

13.
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   

14.
R D LaReau  W Wan  V E Anderson 《Biochemistry》1989,28(8):3619-3624
The isotope effect on binding [4-2H]NAD+ and [4-3H]NAD+ to lactate dehydrogenase has been shown to be 1.10 +/- 0.03 by whole molecule isotope ratio mass spectrometry and 1.085 +/- 0.01 by 3H/14C scintillation counting. These values demonstrate that specific interactions of the nicotinamide ring with the enzyme make the C-H bond at C-4 less stiff in the binary complex.  相似文献   

15.
Methylmalonate-semialdehyde dehydrogenase (MSDH) belongs to the CoA-dependent aldehyde dehydrogenase subfamily. It catalyzes the NAD-dependent oxidation of methylmalonate semialdehyde (MMSA) to propionyl-CoA via the acylation and deacylation steps. MSDH is the only member of the aldehyde dehydrogenase superfamily that catalyzes a β-decarboxylation process in the deacylation step. Recently, we demonstrated that the β-decarboxylation is rate-limiting and occurs before CoA attack on the thiopropionyl enzyme intermediate. Thus, this prevented determination of the transthioesterification kinetic parameters. Here, we have addressed two key aspects of the mechanism as follows: 1) the molecular basis for recognition of the carboxylate of MMSA; and 2) how CoA binding modulates its reactivity. We substituted two invariant arginines, Arg-124 and Arg-301, by Leu. The second-order rate constant for the acylation step for both mutants was decreased by at least 50-fold, indicating that both arginines are essential for efficient MMSA binding through interactions with the carboxylate group. To gain insight into the transthioesterification, we substituted MMSA with propionaldehyde, as both substrates lead to the same thiopropionyl enzyme intermediate. This allowed us to show the following: 1) the pK(app) of CoA decreases by ~3 units upon binding to MSDH in the deacylation step; and 2) the catalytic efficiency of the transthioesterification is increased by at least 10(4)-fold relative to a chemical model. Moreover, we observed binding of CoA to the acylation complex, supporting a CoA-binding site distinct from that of NAD(H).  相似文献   

16.
Huber T  Menon S  Sakmar TP 《Biochemistry》2008,47(42):11013-11023
Crystal structures of engineered human beta 2-adrenergic receptors (ARs) in complex with an inverse agonist ligand, carazolol, provide three-dimensional snapshots of the disposition of seven transmembrane helices and the ligand-binding site of an important G protein-coupled receptor (GPCR). As expected, beta 2-AR shares substantial structural similarities with rhodopsin, the dim-light photoreceptor of the rod cell. However, although carazolol and the 11- cis-retinylidene moiety of rhodopsin are situated in the same general binding pocket, the second extracellular (E2) loop structures are quite distinct. E2 in rhodopsin shows beta-sheet structure and forms part of the chromophore-binding site. In the beta 2-AR, E2 is alpha-helical and seems to be distinct from the receptor's active site, allowing a potential entry pathway for diffusible ligands. The structures, together with extensive structure-activity relationship (SAR) data from earlier studies, provide insight about possible structural determinants of ligand specificity and how the binding of agonist ligands might alter receptor conformation. We review key features of the new beta 2-AR structures in the context of recent complementary work on the conformational dynamics of GPCRs. We also report 600 ns molecular dynamics simulations that quantified beta 2-AR receptor mobility in a membrane bilayer environment and show how the binding of an agonist ligand, adrenaline (epinephrine), causes conformational changes to the ligand-binding pocket and neighboring helices.  相似文献   

17.
Mutations in human coagulation factor IX cause an X-linked bleeding disorder Hemophilia B, which can be classified as severe, moderately severe and mild based on the plasma levels of factor IX among affected individuals with respect to normal factor IX activity assayed in the patients' plasma (<1%, 2-5%, 6-30%, respectively). Recently, we identified hemophilia B to be a disease with mutations showing clinical variation and speculated that this phenotypic heterogeneity might be a replacement-specific property. Here, we have analyzed the differences in sequence and structural properties among identical mutations with varying phenotypes (IMVPs) by comparing with mutations with uniform phenotypes (MUPs), with recurring reports in Haemophilia B mutation database. Classification of mutations into IMVPs and MUPs has been done based on rigorous systematic evaluation of the clotting activity each mutation is associated with. IMVPs (n = 51) occur in less conserved mutant sites with more tolerated substitutions compared to MUPs (n = 100). A preponderance of CpG site mutations and Arg as the mutated residue in IMVPs compared to Cys in MUPs was observed. Hence, a CpG site substitution at less conserved Arg site might have an increased propensity of expressing variable phenotypes. The changes in intrinsic properties associated with the mutation are less drastic for IMVPs than for MUPs, though no significant differences were observed in structural properties. Based on this study and available literature we speculate that modifier genes at other loci, epigenetic interactions and environment may serve individually or cumulatively to bring about the clinical variation implicating hemophilia B to be deviation from classical Mendelian disorder with complete penetrance. We demonstrate that phenotypic heterogeneity appears to be site-specific also owing to the lesser conservation of the mutant site.  相似文献   

18.
Changes induced on addition of the coenzyme, NADH or NAD+, to porcine lactic dehydrogenase isoenzymes H4 and M4 have been studied by hydrodynamic and spectroscopic methods. As shown by ultracentrifugal analysis, the native subunit structure remains unchanged on holoenzyme formation; an approximately 5% increase of the sedimentation coefficient, parallelled by a slight decrease of the partial specific volume (less than 1%) indicate a significant change in the native tertiary and/or quaternary structure of the enzymes, corroborating earlier calorimetric data (Hinz and Jaenicke, 1975). The binding constant for the enzyme from skeletal muscle (M4) and NADH are found to be in agreement with KD-values obtained from equilibrium dialysis, as well as spectroscopic and thermal titration experiments (8 muM). Far UV circular dichroism measurements do not show significant changes on ligand binding, indicating unchanged helicity or compensatory conformational effects. In the near UV, ligand binding is reflected by an extrinsic Cotton effect around 340 nm; in the range of aromatic absorption no changes are detectable. The experimental results suggest that there are gross structural changes on coenzyme binding to lactic dehydrogenase which do not affect the intrinsic spectral properties normally applied to analyze transconformation reactions in protein molecules.  相似文献   

19.
The structure of the N62D mutant of the junction-resolving endonuclease VII (EndoVII) from phage T4 has been refined at 1.3 A, and a second wild-type crystal form solved and refined at 2.8 A resolution. Comparison of the mutant with the wild-type protein structure in two different crystal environments reveals considerable conformational flexibility at the dimer level affecting the substrate-binding cleft, the dimerization interface and the orientation of the C-terminal domains. The opening of the DNA-binding cleft, the orientation of the C-terminal domains relative to the central dimerization domain as well as the relative positioning of helices in the dimerization interface appear to be sensitive to the crystal packing environment. The highly unexpected rearrangement within the extended hydrophobic interface does change the contact surface area but keeps the number of hydrophobic contacts about the same and will therefore not require significant energy input. The conformational flexibility most likely is of functional significance for the broad substrate specificity of EndoVII. Binding of sulphate ions in the mutant structure and their positions relative to the active-site metal ions and residues known to be essential for catalysis allows us to propose a possible catalytic mechanism. A comparison with the active-site geometries of other magnesium-dependent nucleases, among them the homing endonuclease I-PpoI and Serratia endonuclease, shows common features, suggesting related catalytic mechanisms.  相似文献   

20.
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