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 578 毫秒
1.
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   

2.
Transhydrogenase undergoes conformational changes to couple the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. The protein comprises three components: dI, which binds NAD(H); dIII, which binds NADP(H); and dII, which spans the membrane. Experiments using isothermal titration calorimetry, analytical ultracentrifugation, and small angle x-ray scattering show that, as in the crystalline state, a mixture of recombinant dI and dIII from Rhodospirillum rubrum transhydrogenase readily forms a dI(2)dIII(1) heterotrimer in solution, but we could find no evidence for the formation of a dI(2)dIII(2) tetramer using these techniques. The asymmetry of the complex suggests that there is an alternation of conformations at the nucleotide-binding sites during proton translocation by the complete enzyme. The characteristics of nucleotide interaction with the isolated dI and dIII components and with the dI(2)dIII(1) heterotrimer were investigated. (a) The rate of release of NADP(+) from dIII was decreased 5-fold when the component was incorporated into the heterotrimer. (b) The binding affinity of one of the two nucleotide-binding sites for NADH on the dI dimer was decreased about 17-fold in the dI(2)dIII(1) complex; the other binding site was unaffected. These observations lend strong support to the alternating-site mechanism.  相似文献   

3.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  相似文献   

4.
All transhydrogenases appear to have three components: dI, which binds NAD(H), and dIII, which binds NADP(H), protrude from the membrane, and dII spans the membrane. However, the polypeptide composition of the enzymes varies amongst species. The transhydrogenases of Mycobacterium tuberculosis and of Rhodospirillum rubrum have three polypeptides. Sequence analysis indicates that an ancestral three-polypeptide enzyme evolved into transhydrogenases with either two polypeptides (such as the Escherichia coli enzyme) or one polypeptide (such as the mitochondrial enzyme). The fusion steps in each case probably led to the development of an additional transmembrane helix. A hybrid transhydrogenase was constructed from the dI component of the M. tuberculosis enzyme and the dII and dIII components of the R. rubrum enzyme. The hybrid catalyses cyclic transhydrogenation but not the proton-translocating, reverse reaction. This shows that nucleotide-binding/release at the NAD(H) site, and hydride transfer, are fully functional but that events associated with NADP(H) binding/release are compromised. It is concluded that sequence mismatch in the hybrid prevents a conformational change between dI and dIII which is essential for the step accompanying proton translocation.  相似文献   

5.
Rosalind Wilson 《BBA》2006,1757(3):215-223
All transhydrogenases appear to have three components: dI, which binds NAD(H), and dIII, which binds NADP(H), protrude from the membrane, and dII spans the membrane. However, the polypeptide composition of the enzymes varies amongst species. The transhydrogenases of Mycobacterium tuberculosis and of Rhodospirillum rubrum have three polypeptides. Sequence analysis indicates that an ancestral three-polypeptide enzyme evolved into transhydrogenases with either two polypeptides (such as the Escherichia coli enzyme) or one polypeptide (such as the mitochondrial enzyme). The fusion steps in each case probably led to the development of an additional transmembrane helix. A hybrid transhydrogenase was constructed from the dI component of the M. tuberculosis enzyme and the dII and dIII components of the R. rubrum enzyme. The hybrid catalyses cyclic transhydrogenation but not the proton-translocating, reverse reaction. This shows that nucleotide-binding/release at the NAD(H) site, and hydride transfer, are fully functional but that events associated with NADP(H) binding/release are compromised. It is concluded that sequence mismatch in the hybrid prevents a conformational change between dI and dIII which is essential for the step accompanying proton translocation.  相似文献   

6.
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   

7.
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   

8.
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   

9.
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   

10.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The protein has three components: dI binds NADH, dIII binds NADP+, and dII spans the membrane. Transhydrogenase is a "dimer" of two dI-dII-dIII "monomers"; x-ray structures suggested that the two catalytic sites alternate during turnover. Invariant Tyr146 in recombinant dI of Rhodospirillum rubrum transhydrogenase was substituted with Phe and Ala (proteins designated dI.Y146F and dI.Y146A, respectively). Analytical ultracentrifuge experiments and differential scanning calorimetry show that dI.Y146A more readily dissociates into monomers than wild-type dI. Analytical ultracentrifuge and Trp fluorescence experiments indicate that the dI.Y146A monomers bind NADH much more weakly than dimers. Wild-type dI and dI.Y146F reconstituted activity to dI-depleted membranes with similar characteristics. However, dI.Y146A reconstituted activity in its dimeric form but not in its monomeric form, this despite monomers retaining their native fold and binding to the dI-depleted membranes. It is suggested that transhydrogenase reconstructed with monomers of dI.Y146A is catalytically compromised, at least partly as a consequence of the lowered affinity for NADH, and this results from lost interactions between the nucleotide binding site and the protein beta-hairpin upon dissociation of the dI dimer. The importance of these interactions and their coupling to dI domain rotation in the mechanism of action of transhydrogenase is emphasized. Two peaks in the 1H NMR spectrum of wild-type dI are broadened in dI.Y146A and are tentatively assigned to S-methyl groups of Met resonances in the beta-hairpin, consistent with the segmental mobility of this feature in the structure.  相似文献   

11.
Recent developments have led to advances in our understanding of the structure and mechanism of action of proton-translocating (or AB) transhydrogenase. There is (a) a high-resolution crystal structure, and an NMR structure, of the NADP(H)-binding component (dIII), (b) a homology-based model of the NAD(H)-binding component (dI) and (c) an emerging consensus on the position of the transmembrane helices (in dII). The crystal structure of dIII, in particular, provides new insights into the mechanism by which the energy released in proton translocation across the membrane is coupled to changes in the binding affinities of NADP(+) and NADPH that drive the chemical reaction.  相似文献   

12.
Proton-translocating nicotinamide nucleotide transhydrogenases contain an NAD(H)-binding domain (dI), an NADP(H)-binding domain (dIII) and a membrane domain (dII) with the proton channel. Separately expressed and isolated dIII contains tightly bound NADP(H), predominantly in the oxidized form, possibly representing a so-called "occluded" intermediary state of the reaction cycle of the intact enzyme. Despite a K(d) in the micromolar to nanomolar range, this NADP(H) exchanges significantly with the bulk medium. Dissociated NADP(+) is thus accessible to added enzymes, such as NADP-isocitrate dehydrogenase, and can be reduced to NADPH. In the present investigation, dissociated NADP(H) was digested with alkaline phosphatase, removing the 2'-phosphate and generating NAD(H). Surprisingly, in the presence of dI, the resulting NADP(H)-free dIII catalyzed a rapid reduction of 3-acetylpyridine-NAD(+) by NADH, indicating that 3-acetylpyridine-NAD(+) and/or NADH interacts unspecifically with the NADP(H)-binding site. The corresponding reaction in the intact enzyme is not associated with proton pumping. It is concluded that there is a 2'-phosphate-binding region in dIII that controls tight binding of NADP(H) to dIII, which is not a required for fast hydride transfer. It is likely that this region is the Lys424-Arg425-Ser426 sequence and loops D and E. Further, in the intact enzyme, it is proposed that the same region/loops may be involved in the regulation of NADP(H) binding by an electrochemical proton gradent.  相似文献   

13.
14.
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   

15.
Transhydrogenase is a proton pump. It has three components: dI and dIII protrude from the membrane and contain the binding sites for NAD(H) and NADP(H), respectively, and dII spans the membrane. We have expressed dIII from Homo sapiens transhydrogenase (hsdIII) in Escherichia coli. The purified protein was associated with stoichiometric amounts of NADP(H) bound to the catalytic site. The NADP+ and NADPH were released only slowly from the protein, supporting the suggestion that nucleotide-binding by dIII is regulated by the membrane-spanning dII. HsdIII formed a catalytically active complex with recombinant dI from Rhodospirillum rubrum (rrdI), even in the absence of dII. The rates of forward and reverse transhydrogenation catalysed by this complex are probably limited by slow release from dIII of NADPH and NADP+, respectively. The hybrid complex also catalysed high rates of 'cyclic' transhydrogenation, indicating that hydride transfer, and exchange of nucleotides with dI, are rapid. Stopped-flow experiments revealed a rapid, monoexponential, single-turnover burst of reverse transhydrogenation in pre-steady-state. The apparent first-order rate constant of the burst increased with the concentration of rrdI. A deuterium isotope effect (kH/kD approximately 2 at 27 degrees C) was observed when [4B-1H]NADPH was replaced with [4B-2H]NADPH. The characteristics of the burst of transhydrogenation with rrdI:hsdIII differed from those previously reported for rrdI:rrdIII (J.D. Venning et al., Eur. J. Biochem. 257 (1998) 202-209), but the differences are readily explained by a greater dissociation constant of the hybrid complex. The steady-state rate of reverse transhydrogenation by the rrdI:hsdIII complex was almost independent of pH, but there was a single apparent pKa ( approximately 9.1) associated with the cyclic reaction. The reactions of the dI:dIII complex probably proceed independently of those protonation/deprotonation reactions which, in the complete enzyme, are associated with H+ translocation.  相似文献   

16.
Proton-pumping nicotinamide nucleotide transhydrogenases are composed of three main domains, the NAD(H)-binding and NADP(H)-binding hydrophilic domains I (dI) and III (dIII), respectively, and the hydrophobic domain II (dII) containing the assumed proton channel. dII in the Escherichia coli enzyme has recently been characterised with regard to topology and a packing model of the helix bundle in dII is proposed. Extensive mutagenesis of conserved charged residues of this domain showed that important residues are betaHis91 and betaAsn222. The pH dependence of betaH91D, as well as betaH91C (unpublished), when compared to that of wild type shows that reduction of 3-acetylpyridine-NAD(+) by NADPH, i.e., the reverse reaction, is optimal at a pH essentially coinciding with the pK(a) of the residue in the beta91 position. It is therefore concluded that the wild-type transhydrogenase is regulated by the degree of protonation of betaHis91. The mechanisms of the interactions between dI+dIII and dII are suggested to involve pronounced conformational changes in a 'hinge' region around betaR265.  相似文献   

17.
Proton-translocating nicotinamide nucleotide transhydrogenases contain an NAD(H)-binding domain (dI), an NADP(H)-binding domain (dIII) and a membrane domain (dII) with the proton channel. Separately expressed and isolated dIII contains tightly bound NADP(H), predominantly in the oxidized form, possibly representing a so-called “occluded” intermediary state of the reaction cycle of the intact enzyme. Despite a Kd in the micromolar to nanomolar range, this NADP(H) exchanges significantly with the bulk medium. Dissociated NADP+ is thus accessible to added enzymes, such as NADP-isocitrate dehydrogenase, and can be reduced to NADPH. In the present investigation, dissociated NADP(H) was digested with alkaline phosphatase, removing the 2′-phosphate and generating NAD(H). Surprisingly, in the presence of dI, the resulting NADP(H)-free dIII catalyzed a rapid reduction of 3-acetylpyridine-NAD+ by NADH, indicating that 3-acetylpyridine-NAD+ and/or NADH interacts unspecifically with the NADP(H)-binding site. The corresponding reaction in the intact enzyme is not associated with proton pumping. It is concluded that there is a 2′-phosphate-binding region in dIII that controls tight binding of NADP(H) to dIII, which is not a required for fast hydride transfer. It is likely that this region is the Lys424-Arg425-Ser426 sequence and loops D and E. Further, in the intact enzyme, it is proposed that the same region/loops may be involved in the regulation of NADP(H) binding by an electrochemical proton gradent.  相似文献   

18.
Transhydrogenase couples proton translocation across a membrane to hydride transfer between NADH and NADP+. Previous x-ray structures of complexes of the nucleotide-binding components of transhydrogenase ("dI2dIII1" complexes) indicate that the dihydronicotinamide ring of NADH can move from a distal position relative to the nicotinamide ring of NADP+ to a proximal position. The movement might be responsible for gating hydride transfer during proton translocation. We have mutated three invariant amino acids, Arg-127, Asp-135, and Ser-138, in the NAD(H)-binding site of Rhodospirillum rubrum transhydrogenase. In each mutant, turnover by the intact enzyme is strongly inhibited. Stopped-flow experiments using dI2dIII1 complexes show that inhibition results from a block in the steps associated with hydride transfer. Mutation of Asp-135 and Ser-138 had no effect on the binding affinity of either NAD+ or NADH, but mutation of Arg-127 led to much weaker binding of NADH and slightly weaker binding of NAD+. X-ray structures of dI2dIII1 complexes carrying the mutations showed that their effects were restricted to the locality of the bound NAD(H). The results are consistent with the suggestion that in wild-type protein movement of the Arg-127 side chain, and its hydrogen bonding to Asp-135 and Ser-138, stabilizes the dihydronicotinamide of NADH in the proximal position for hydride transfer.  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.
Proton-translocating transhydrogenase (TH) couples direct and stereospecific hydride transfer between NAD(H) and NADP(H), bound to soluble domains dI and dIII, respectively, to proton translocation across a membrane bound domain, dII. The reaction occurs with proton-gradient coupled conformational changes, which affect the energetics of substrate binding and interdomain interactions. The crystal structure of TH dIII from Rhodospirillum rubrum has been determined in the presence of NADPH (2.4 A) and NADP (2.1 A) (space group P6(1)22). Each structure has two molecules in the asymmetric unit, differing in the conformation of the NADP(H) binding loop D. In one molecule, loop D has an open conformation, with the B face of (dihydro)nicotinamide exposed to solvent. In the other molecule, loop D adopts a hitherto unobserved closed conformation, resulting in close interactions between NADP(H) and side chains of the highly conserved residues, betaSer405, betaPro406, and betaIle407. The conformational change shields the B face of (dihydro)nicotinamide from solvent, which would block hydride transfer in the intact enzyme. It also alters the environments of invariant residues betaHis346 and betaAsp393. However, there is little difference in either the open or the closed conformation upon change in oxidation state of nicotinamide, i.e., for NADP vs. NADPH. Consequently, the occurrence of two loop D conformations for both substrate oxidation states gives rise to four states: NADP-open, NADP-closed, NADPH-open, and NADPH-closed. Because these states are distinguished by protein conformation and by net charge they may be important in the proton translocating mechanism of intact TH.  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