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 0 毫秒
1.
Transhydrogenase couples the redox (hydride-transfer) reaction between NAD(H) and NADP(H) to proton translocation across a membrane. The redox reaction is catalyzed at the interface between two components (dI and dIII) which protrude from the membrane. A complex formed from recombinant dI and dIII (the dI(2)dIII(1) complex) from Rhodospirillum rubrum transhydrogenase catalyzes fast single-turnover hydride transfer between bound nucleotides. In this report we describe three new crystal structures of the dI(2)dIII(1) complex in different nucleotide-bound forms. The structures reveal an asymmetry in nucleotide binding that complements results from solution studies and supports the notion that intact transhydrogenase functions by an alternating site mechanism. In one structure, the redox site is occupied by NADH (on dI) and NADPH (on dIII). The dihydronicotinamide rings take up positions which may approximate to the ground state for hydride transfer: the redox-active C4(N) atoms are separated by only 3.6 A, and the perceived reaction stereochemistry matches that observed experimentally. The NADH conformation is different in the two dI polypeptides of this form of the dI(2)dIII(1) complex. Comparisons between a number of X-ray structures show that a conformational change in the NADH is driven by relative movement of the two domains which comprise dI. It is suggested that an equivalent conformational change in the intact enzyme is important in gating the hydride-transfer reaction. The observed nucleotide conformational change in the dI(2)dIII(1) complex is accompanied by rearrangements in the orientation of local amino acid side chains which may be responsible for sealing the site from the solvent and polarizing hydride transfer.  相似文献   

2.
Transhydrogenase couples the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. Coupling is achieved through changes in protein conformation. Upon mixing, the isolated nucleotide-binding components of transhydrogenase (dI, which binds NAD(H), and dIII, which binds NADP(H)) form a catalytic dI(2).dIII(1) complex, the structure of which was recently solved by x-ray crystallography. The fluorescence from an engineered Trp in dIII changes when bound NADP(+) is reduced. Using a continuous flow device, we have measured the Trp fluorescence change when dI(2).dIII(1) complexes catalyze reduction of NADP(+) by NADH on a sub-millisecond scale. At elevated NADH concentrations, the first-order rate constant of the reaction approaches 21,200 s(-1), which is larger than that measured for redox reactions of nicotinamide nucleotides in other, soluble enzymes. Rather high concentrations of NADH are required to saturate the reaction. The deuterium isotope effect is small. Comparison with the rate of the reverse reaction (oxidation of NADPH by NAD(+)) reveals that the equilibrium constant for the redox reaction on the complex is >36. This high value might be important in ensuring high turnover rates in the intact enzyme.  相似文献   

3.
Transhydrogenase, found in bacterial membranes and inner mitochondrial membranes of animal cells, couples the redox reaction between NAD(H) and NADP(H) to proton translocation. In this work, the invariant Gln132 in the NAD(H)-binding component (dI) of the Rhodospirillum rubrum transhydrogenase was substituted with Asn (to give dI.Q132N). Mixtures of the mutant protein and the NADP(H)-binding component (dIII) of the enzyme readily produced an asymmetric complex, (dI.Q132N)(2)dIII(1). The X-ray structure of the complex revealed specific changes in the interaction between bound nicotinamide nucleotides and the protein at the hydride transfer site. The first-order rate constant of the redox reaction between nucleotides bound to (dI.Q132N)(2)dIII(1) was <1% of that for the wild-type complex, and the deuterium isotope effect was significantly decreased. The nucleotide binding properties of the dI component in the complex were asymmetrically affected by the Gln-to-Asn mutation. In intact, membrane-bound transhydrogenase, the substitution completely abolished all catalytic activity. The results suggest that Gln132 in the wild-type enzyme behaves as a "tether" or a "tie" in the mutual positioning of the (dihydro)nicotinamide rings of NAD(H) and NADP(H) for hydride transfer during the conformational changes that are coupled to the translocation of protons across the membrane. This ensures that hydride transfer is properly gated and does not take place in the absence of proton translocation.  相似文献   

4.
5.
Chemical modification of amino acid residues with phenylglyoxal, diethylpyrocarbonate, and N-bromosuccinimide indicated that at least one residue each of arginine, histidine, and tryptophan were necessary for the activity of human liver serine hydroxymethyltransferase. Protection by substrates suggested that these residues might occur at the active site of the enzyme.  相似文献   

6.
The beta-ketoacyl synthetase site of eukaryotic fatty acid synthetases is comprised in part of a pantetheinyl residue on one subunit juxtapositioned with a cysteinyl residue on the adjacent subunit. The present study has confirmed this arrangement and has identified 2 additional residues in the site. The active site residues were identified as summarized below. Sodium borohydride reduction of the keto derivatives of the dibromopropanone cross-linked residues yielded the alcohol derivatives which were amenable to isolation in good yields. The active enzyme yielded primarily a cysteinecysteamine derivative of 2-propanol, demonstrating that a cystyl and the pantetheinyl residues were cross-linked by dibromopropanone. However, in the cold-inactivated enzyme, the primary product of the cross-linking reaction was the dicystyl derivative. In addition, cross-linking between the cystyl and pantetheinyl residues, but not the two cystyl residues, resulted in the cross-linking of the two subunits. Therefore, it is proposed that there are two cystyl residues on one subunit juxtapositioned with the pantetheinyl residue on the adjacent subunit. The cystyl residues are highly reactive toward alkylating agents at pH 6.5, suggesting the presence of a cationic residue interacting with the thiolate anion. This proposal was supported using the bifunctional reagent o-phthalaldehyde which was found to cross-link the epsilon-amino group of lysine with the pantetheinyl-SH or the cystyl-SH in the beta-ketoacyl synthetase site to form a thioisoindole ring. The dialdehyde inhibited the enzyme by inactivating the beta-ketoacyl synthetase activity, and the inhibition could be prevented by malonyl-CoA and to a lesser extent by acetyl-CoA. Blocking the reactive thiol groups with dibromopropanone or 5,5'-dithiobis(2-nitrobenzoic acid) reduced the formation of the fluorescent thioisoindole ring. The close arrangement of a cystyl-SH, the pantetheinyl-SH, and the epsilon-amino group of lysine led us to propose that the positive epsilon-amino group may serve as an electron sink in a general acid-catalyzed decarboxylation reaction.  相似文献   

7.
Proton-translocating transhydrogenase is found in the inner membranes of animal mitochondria, and in the cytoplasmic membranes of many bacteria. It catalyses hydride transfer from NADH to NADP(+) coupled to inward proton translocation. Evidence is reviewed suggesting the enzyme operates by a "binding-change" mechanism. Experiments with Escherichia coli transhydrogenase indicate the enzyme is driven between "open" and "occluded" states by protonation and deprotonation reactions associated with proton translocation. In the open states NADP(+)/NADPH can rapidly associate with, or dissociate from, the enzyme, and hydride transfer is prevented. In the occluded states bound NADP(+)/NADPH cannot dissociate, and hydride transfer is allowed. Crystal structures of a complex of the nucleotide-binding components of Rhodospirillum rubrum transhydrogenase show how hydride transfer is enabled and disabled at appropriate steps in catalysis, and how release of NADP(+)/NADPH is restricted in the occluded state. Thermodynamic and kinetic studies indicate that the equilibrium constant for hydride transfer on the enzyme is elevated as a consequence of the tight binding of NADPH relative to NADP(+). The protonation site in the translocation pathway must face the outside if NADP(+) is bound, the inside if NADPH is bound. Chemical shift changes detected by NMR may show where alterations in protein conformation resulting from NADP(+) reduction are initiated. This article is part of a Special Issue entitled: 17th European Bioenergetics Conference (EBEC 2012).  相似文献   

8.
Aminopeptidases are major enzymes in the midgut microvillar membranes of most insects and are targets of insecticidal Bacillus thuringiensis crystal delta-endotoxins. Sequence analysis and substrate specificity studies showed that these enzymes resemble mammalian aminopeptidase N, although information on the organization of their active site is lacking. The effect of pH at different temperatures on the kinetic parameters of Tenebrio molitor (Coleoptera) larval aminopeptidase showed that enzyme catalysis depend on a deprotonated (pK 7.6; DeltaH degrees (ion), 7.6 kJ/mol) and a protonated (pK 8.2; DeltaH degrees (ion), 16.8 kJ/mol) group. 1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide and diethylpyrocarbonate inactivate the enzyme by modifying a pK 5.8 carboxylate and a imidazole group, respectively, with a reaction order around 1. Tetranitromethane changes the K(m) of the enzyme without affecting its V(max) by modifying a phenol group. The presence of a competitive inhibitor decrease the inactivation reaction rates in all these cases. EDTA inactivation of the aminopeptidase is affected by pH and temperature suggesting the involvement in metal binding of at least one deprotonated imidazole group (pK 5.8, DeltaH degrees (ion), 20 kJ/mol). The data support the hypothesis that T. molitor aminopeptidase catalysis depends on a catalytic metal and on a carboxylate and a protonated imidazole group, whereas substrate binding relies in one phenol and one carboxylate groups. The insect aminopeptidase shares common features with mammalian aminopeptidase N, although differing in details of substrate binding and in residues directly involved in catalysis.  相似文献   

9.
BACKGROUND: Membrane-bound ion pumps are involved in metabolic regulation, osmoregulation, cell signalling, nerve transmission and energy transduction. How the ion electrochemical gradient interacts with the scalar chemistry and how the catalytic machinery is gated to ensure high coupling efficiency are fundamental to the mechanism of action of such pumps. Transhydrogenase is a conformationally coupled proton pump linking a proton gradient to the redox reaction between NAD(H) and NADP(H). The enzyme has three components; dI binds NAD(H), dII spans the membrane and dIII binds NADP(H). RESULTS: The first crystal structure of a transhydrogenase dI component (from Rhodospirillum rubrum) has been determined at 2.0 A resolution. The monomer comprises two domains. Both are involved in dimer formation, and one has a Rossmann fold that binds NAD+ in a novel mode. The two domains can adopt different conformations. In the most closed conformation, the nicotinamide ring is expelled from the cleft between the two domains and is exposed on the outside of the protein. In this conformation it is possible to dock the structure of dI/NAD+ with that of a dIII/NADP+ complex to provide the first insights into the molecular basis of the hydride-transfer step. CONCLUSIONS: Analysis of the model of the dI/dIII complex identifies residues potentially involved in dI/dIII interaction and shows how domain motion in dI results in a shift in position of the nicotinamide ring of NAD+. We propose that this movement is responsible for switching between the forbidden and allowed states for hydride transfer during proton pumping.  相似文献   

10.
Transhydrogenase couples the transfer of hydride-ion equivalents between NAD(H) and NADP(H) to proton translocation across a membrane. The enzyme has three components: dI binds NAD(H), dIII binds NADP(H) and dII spans the membrane. Coupling between transhydrogenation and proton translocation involves changes in the binding of NADP(H). Mixtures of isolated dI and dIII from Rhodospirillum rubrum transhydrogenase catalyse a rapid, single-turnover burst of hydride transfer between bound nucleotides; subsequent turnover is limited by NADP(H) release. Stopped-flow experiments showed that the rate of the hydride transfer step is decreased at low pH. Single Trp residues were introduced into dIII by site-directed mutagenesis. Two mutants with similar catalytic properties to those of the wild-type protein were selected for a study of nucleotide release. The way in which Trp fluorescence was affected by nucleotide occupancy of dIII was different in the two mutants, and hence two different procedures for determining the rate of nucleotide release were developed. The apparent first-order rate constants for NADP(+) release and NADPH release from isolated dIII increased dramatically at low pH. It is concluded that a single ionisable group in dIII controls both the rate of hydride transfer and the rate of nucleotide release. The properties of the protonated and unprotonated forms of dIII are consistent with those expected of intermediates in the NADP(H)-binding-change mechanism. The ionisable group might be a component of the proton-translocation pathway in the complete enzyme.  相似文献   

11.
Conformational changes in proton pumping transhydrogenases have been suggested to be dependent on binding of NADP(H) and the redox state of this substrate. Based on a detailed amino acid sequence analysis, it is argued that a classical betaalphabetaalphabeta dinucleotide binding fold is responsible for binding NADP(H). A model defining betaA, alphaB, betaB, betaD, and betaE of this domain is presented. To test this model, four single cysteine mutants (cfbetaA348C, cfbetaA390C, cfbetaK424C, and cfbetaR425C) were introduced into a functional cysteine-free transhydrogenase. Also, five cysteine mutants were constructed in the isolated domain III of Escherichia coli transhydrogenase (ecIIIH345C, ecIIIA348C, ecIIIR350C, ecIIID392C, and ecIIIK424C). In addition to kinetic characterizations, effects of sulfhydryl-specific labeling with N-ethylmaleimide, 2-(4'-maleimidylanilino)naphthalene-6-sulfonic acid, and diazotized 3-aminopyridine adenine dinucleotide (phosphate) were examined. The results are consistent with the view that, in agreement with the model, beta-Ala348, beta-Arg350, beta-Ala390, beta-Asp392, and beta-Lys424 are located in or close to the NADP(H) site. More specifically, beta-Ala348 succeeds betaB. The remarkable reactivity of betaR350C toward NNADP suggests that this residue is close to the nicotinamide moiety of NADP(H). beta-Ala390 and beta-Asp392 terminate or succeed betaD, and are thus, together with the region following betaA, creating the switch point crevice where NADP(H) binds. beta-Asp392 is particularly important for the substrate affinity, but it could also have a more complex role in the coupling mechanism for transhydrogenase.  相似文献   

12.
Transhydrogenase (E.C. 1.6.1.1) couples the redox reaction between NAD(H) and NADP(H) to the transport of protons across a membrane. The enzyme is composed of three components. The dI and dIII components, which house the binding site for NAD(H) and NADP(H), respectively, are peripheral to the membrane, and dII spans the membrane. We have estimated dissociation constants (K(d) values) for NADPH (0.87 microM), NADP(+) (16 microM), NADH (50 microM), and NAD(+) (100-500 microM) for intact, detergent-dispersed transhydrogenase from Escherichia coli using micro-calorimetry. This is the first complete set of dissociation constants of the physiological nucleotides for any intact transhydrogenase. The K(d) values for NAD(+) and NADH are similar to those previously reported with isolated dI, but the K(d) values for NADP(+) and NADPH are much larger than those previously reported with isolated dIII. There is negative co-operativity between the binding sites of the intact, detergent-dispersed transhydrogenase when both nucleotides are reduced or both are oxidized.  相似文献   

13.
Murine A/J anti-p-azophenylarsonate (Ars) antibodies sharing a predominant idiotype are encoded by a single combination of germ-line V region gene segments. The dominance of this idiotype among secondary immune response anti-Ars antibodies has been explained by the Ag-driven selection of favorable somatic mutants of this gene segment combination, associated with an intrinsic Ars-affinity of the germ-line V region higher than that of other possible combinations. To determine the effect of junctional diversity upon affinity for Ag, independently of somatic mutation, we determined the V region sequences and affinity for Ars of five primary response antibodies. These antibodies share identical unmutated V regions but differ only at the D gene junctions. Among the five antibodies, Ars-affinity differed up to 10-fold depending upon the identity of the amino acid residues at the VH-D and the D-JH junctions. The combination of junctional residues observed in two primary response antibodies with relatively low Ars-affinity has not been observed among secondary response antibodies. Thus the identity of junctional residues resulting from gene rearrangement prior to antigen stimulation must be taken into account in hypotheses which account for idiotype dominance by selection on the basis of affinity.  相似文献   

14.
15.
Proton-translocating nicotinamide nucleotide transhydrogenase is a conformationally driven pump which catalyzes the reversibel reduction of NADP(+) by NADH. Transhydrogenases contain three domains, i.e., the hydrophilic NAD(H)-binding domain I and the NADP(H)-binding domain III, and the hydrophobic domain II containing the proton channel. Domains I and III have been separately expressed and characterized structurally by, e.g. X-ray crystallography and NMR. These domains catalyze transhydrogenation in the absence of domain II. However, due to the absence of the latter domain, the reactions catalyzed by domains I and III differ significantly from those catalyzed by the intact enzyme. Mutagenesis of residues in domain II markedly affects the activity of the intact enzyme. In order to resolve the structure-function relationships of the intact enzyme, and the molecular mechanism of proton translocation, it is therefore essential to establish the structure and function of domain II and its interactions with domains I and III. This review describes some relevant recent results in this field of research.  相似文献   

16.
17.
BACKGROUND: Membrane-bound ion translocators have important functions in biology, but their mechanisms of action are often poorly understood. Transhydrogenase, found in animal mitochondria and bacteria, links the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. Linkage is achieved through changes in protein conformation at the nucleotide binding sites. The redox reaction takes place between two protein components located on the membrane surface: dI, which binds NAD(H), and dIII, which binds NADP(H). A third component, dII, provides a proton channel through the membrane. Intact membrane-located transhydrogenase is probably a dimer (two copies each of dI, dII, and dIII). RESULTS: We have solved the high-resolution crystal structure of a dI:dIII complex of transhydrogenase from Rhodospirillum rubrum-the first from a transhydrogenase of any species. It is a heterotrimer, having two polypeptides of dI and one of dIII. The dI polypeptides fold into a dimer. The loop on dIII, which binds the nicotinamide ring of NADP(H), is inserted into the NAD(H) binding cleft of one of the dI polypeptides. The cleft of the other dI is not occupied by a corresponding dIII component. CONCLUSIONS: The redox step in the transhydrogenase reaction is readily visualized; the NC4 atoms of the nicotinamide rings of the bound nucleotides are brought together to facilitate direct hydride transfer with A-B stereochemistry. The asymmetry of the dI:dIII complex suggests that in the intact enzyme there is an alternation of conformation at the catalytic sites associated with changes in nucleotide binding during proton translocation.  相似文献   

18.
Modification of maize δ-aminolevulinic acid dehydratase (ALAD) by diethylpyrocarbonate (DEP) caused rapid and complete inactivation of the enzyme. The inactivation showed saturation kinetics with a half inactivation time at saturating DEP equal to 0.3 min and KDEP  0.3 mM. Substrate δ-aminolevulinic acid (ALA) and competitive inhibitor levulinic acid protected against inactivation, thereby indicating that DEP modifies the active site. The modified enzyme showed an increase in absorbance at 240 nm which was lost upon treatment with 0.8 M hydroxylamine. Most of the activity lost by DEP treatment could be restored after treatment with 0.8 M hydroxylamine. The results suggest that DEP modifies 7.4 residues/mole of the enzyme. These histidine residues are essential for catalysis by ALAD.  相似文献   

19.
Tania Bizouarn  Tina Bhakta 《BBA》2005,1708(3):404-410
Transhydrogenase (E.C. 1.6.1.1) couples the redox reaction between NAD(H) and NADP(H) to the transport of protons across a membrane. The enzyme is composed of three components. The dI and dIII components, which house the binding site for NAD(H) and NADP(H), respectively, are peripheral to the membrane, and dII spans the membrane. We have estimated dissociation constants (Kd values) for NADPH (0.87 μM), NADP+ (16 μM), NADH (50 μM), and NAD+ (100-500 μM) for intact, detergent-dispersed transhydrogenase from Escherichia coli using micro-calorimetry. This is the first complete set of dissociation constants of the physiological nucleotides for any intact transhydrogenase. The Kd values for NAD+ and NADH are similar to those previously reported with isolated dI, but the Kd values for NADP+ and NADPH are much larger than those previously reported with isolated dIII. There is negative co-operativity between the binding sites of the intact, detergent-dispersed transhydrogenase when both nucleotides are reduced or both are oxidised.  相似文献   

20.
To elucidate the role of amino acid residues adjacent to the catalytic site of pepsin-like enzymes, we analyzed and compared the crystal structures of these enzymes, their complexes with inhibitors, and zymogens in the active site area (a total of 82 structures). In addition to the water molecule (W1) located between the active carboxyls and playing a role of the nucleophile during catalytic reaction, another water molecule (W2) at the vicinity of the active groups was found to be completely conserved. This water molecule plays an essential role in formation of a chain of hydrogen-bonded residues between the active site flap and the active carboxyls on ligand binding. These data suggest a new approach to understanding the role of residues around the catalytic site, which can assist the development of the catalytic reaction. The influence of groups adjacent to the active carboxyls is manifested by pepsin activity at pH 1.0. Some features of pepsin-like enzymes and their mutants are discussed in the framework of the approach.  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