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 31 毫秒
1.
Roitel O  Sergienko E  Branlant G 《Biochemistry》1999,38(49):16084-16091
Tetrameric phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from Bacillus stearothermophilus has been described as a "dimer of dimers" with three nonequivalent interfaces, P-axis (between subunits O and P and between subunits Q and R), Q-axis (between subunits O and Q and between subunits P and R), and R-axis interface (between subunits O and R and between subunits P and Q). O-P dimers, the most stable and the easiest to generate, have been created by selective disruption of hydrogen bonds across the R- and Q-axis interfaces by site-directed mutagenesis. Asp-186 and Ser-48, and Glu-276 and Tyr-46, which are hydrogen bond partners across the R- and Q-axis interfaces, respectively, have been replaced with glycine residues. All mutated residues are highly conserved among GAPDHs from different species and are located in loops. Both double mutants D186G/E276G and Y46G/S48G were dimeric, while all single mutants remained tetrameric. As previously described [Clermont, S., Corbier, C., Mely, Y., Gerard, D., Wonacott, A., and Branlant, G. (1993) Biochemistry 32, 10178-10184], NAD binding to wild type GAPDH (wtGAPDH) was interpreted according to the induced-fit model and exhibited negative cooperativity. However, NAD binding to wtGAPDH can be adequately described in terms of two independent dimers with two interacting binding sites in each dimer. Single mutants D186G, E276G, and Y46G exhibited behavior in NAD binding similar to that of the wild type, while both dimeric mutants D186G/E276G and Y46G/S48G exhibited positive cooperativity in binding the coenzyme NAD. The fact that O-P dimer mutants retained cooperative behavior shows that (1) the P-axis interface is important in transmitting the information induced upon NAD binding inside the O-P dimer from one subunit to the other and (2) the S-loop of the R-axis-related subunit is not directly involved in cooperative binding of NAD in the O-P dimer. In both O-P dimer mutants, the absorption band of the binary enzyme-NAD complex had a highly decreased intensity compared to that of the wild type and, in addition, totally disappeared in the presence of G3P or 1,3-dPG. However, no enzymatic activity was detected, indicating that the formed ternary enzyme-NAD-G3P or -1, 3-dPG complex was not catalytically efficient. In the O-P dimers, the interaction with the S-loop of the R-axis-related subunit is disrupted, and therefore, the S-loop should be less structured. This resulted in increased accessibility of the active site to the solvent, particularly for the adenosine-binding site of NAD. Thus, together, this is likely to explain both the lowered affinity of the dimeric enzyme for NAD and the absence of activity.  相似文献   

2.
Tetrameric phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from Bacillus stearothermophilus can be described as a dimer of dimers with three nonequivalent interfaces. To investigate the contribution of intra- and intersubunit interactions to GAPDH thermostability, 10 residues located either at the cofactor domain (amino acids 1-148 and 313-333) or at the catalytic domain (amino acids 149-312) were mutated and the thermal unfolding of the mutants was studied by differential scanning calorimetry in the absence and presence of saturating concentrations of NAD. Disruptions of intrasubunit interactions lead to a drastic decrease in thermostability of the N313T, Y283V, and W310F mutants. Moreover, for the N313T mutant, a weakening of cooperative interactions between the catalytic and the cofactor domains and an inefficient binding of NAD are observed. This is likely the consequences of modification or loss of the hydrogen bonding network associating N313 and residues 236-238 and N313 and the nicotinamide carboxyamide of NAD, respectively. For the residues Y283 and W310, which are involved in stacking hydrophobic interactions, mutating both positions does not affect the efficiency of NAD binding. This shows that the factors involved in the thermostability of the tetrameric apo GAPDH are then different from those induced by NAD binding. Disruption of intersubunit hydrogen bonds between the catalytic domain and the NAD-binding domain of a neighboring subunit also leads to a significant destabilization of the apo tetrameric form as observed for the D282G mutant. Moreover, no efficient binding of NAD is observed. Both results are likely the consequence of a loss of hydrogen bonds across the P-axis and the Q-axis between D282 and R197 and between D282 and R52, respectively. Similar results, i.e., a destabilizing effect and inefficient NAD binding, are observed with the T34Q/T39S/L43Q mutant in which steric hindrance is introduced at the S-loop of the R-axis-related subunit via mutations at the adenosine subsite. The dimeric form of the D282G mutant exhibits a single partial heat absorption peak, whereas the Y46G/R52G mutant which exists only as a dimer shows two peaks. Taking into account the recent small-angle X-ray scattering studies which suggested that the dimeric form of the D282G mutant and of the dimeric Y46G/R52G mutant are of the O-R and O-P types, respectively (Vachette, unpublished results), we propose that the presence of one or two peaks in thermal unfolding of dimers is a signature of the dimer type.  相似文献   

3.
We studied the interaction of chaperonin GroEL with different misfolded forms of tetrameric phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPDH): (1) GAPDH from rabbit muscles with all SH-groups modified by 5,5'-dithiobis(2-nitrobenzoate); (2) O-R-type dimers of mutant GAPDH from Bacillus stearothermophilus with amino acid substitutions Y283V, D282G, and Y283V/W84F, and (3) O-P-type dimers of mutant GAPDH from B. stearothermophilus with amino acid substitutions Y46G/S48G and Y46G/R52G. It was shown that chemically modified GAPDH and the O-R-type mutant dimers bound to GroEL with 1:1 stoichiometry and dissociation constants K(d) of 0.4 and 0.9 muM, respectively. A striking feature of the resulting complexes with GroEL was their stability in the presence of Mg-ATP. Chemically modified GAPDH and the O-R-type mutant dimers inhibited GroEL-assisted refolding of urea-denatured wild-type GAPDH from B. stearothermophilus but did not affect its spontaneous reactivation. In contrast to the O-R-dimers, the O-P-type mutant dimers neither bound nor affected GroEL-assisted refolding of the wild-type GAPDH. Thus, we suggest that interaction of GroEL with certain types of misfolded proteins can result in the formation of stable complexes and the impairment of chaperonin activity.  相似文献   

4.
A single amino acid substitution (Y78R) at the dimer-dimer interface of homotetrameric single stranded DNA binding protein from E. coli (EcoSSB) renders the protein a stable dimer. This dimer can bind single-stranded DNA albeit with greatly reduced affinity. In vivo this dimeric SSB cannot replace homotetrameric EcoSSB. Amino acid changes at the rim of the dimer-dimer interface nearby (Q76K, Q76E) show an electrostatic interaction between a charged amino acid at position 76 and bound nucleic acid. In conclusion, nucleic acid binding to homotetrameric SSB must take place across both dimers to achieve functionally correct binding.  相似文献   

5.
A series of mutant lac repressor proteins at positions 281 or 282 was isolated for detailed characterization. Although Cys281 modification by sulfhydryl reagents abrogates pH effects on inducer binding and diminishes operator binding (Daly, T. J., Olson, J. S., and Matthews, K. S. (1986) Biochemistry 25, 5468-5474), substitution at this site by alanine, serine, phenylalanine, isoleucine, or methionine did not abolish completely the pH shift nor affect operator affinity. Thus, ionization of the sulfhydryl residue does not account fully for the alterations in inducer affinity and cooperativity of binding observed with elevated pH. Substitution for Cys281 did, however, alter the kinetic parameters for inducer association with the protein. The polarity of the side chain at 281 influenced the rates of sugar binding, presumably by altering the rate of opening/closing of the binding site. Furthermore, the presence of the branched side chain of isoleucine at position 281 disrupted oligomerization of the repressor. In contrast to the tolerance for substitution at 281, the only amino acid side chain exchanges for Tyr282 which yielded tetrameric protein with near normal operator binding characteristics were phenylalanine and leucine; this result is consistent with studies of suppressed nonsense mutations at position 282 which indicated repression occurred only for the corresponding substitutions (Kleina, L. G., and Miller, J. H. (1990) J. Mol. Biol. 221, 295-318). Despite the tetrameric character of the Y282F mutant protein, the pH dependence and cooperativity of inducer binding for this mutant protein were altered. All amino acid substitutions other than phenylalanine and leucine at this position resulted in either monomeric protein or no detectable repressor in the cell. Thus, the hydrophobic character of the side chain at position 282 is essential for tetramer formation, and the phenyl ring alone alters inducer binding parameters. The monomeric mutant proteins with substitutions for Tyr282 exhibited lower stability than their tetrameric counterparts, and the absence of dimer formation suggests alterations at this site affect both dimer and tetramer interfaces. Based on previous genetic studies and our detailed mutant characterization, the region encompassing 281 and 282, indicated by secondary structure prediction to be a turn or coil, is essential for oligomer formation and additionally exerts a strong influence on the dynamic properties of the protein, presumably mediated by interactions at the subunit interface which regulate the rate of opening and closing of the inducer binding cleft.  相似文献   

6.
Canonical glutathione (GSH) transferases are dimeric proteins with subunits composed of an N-terminal GSH binding region (domain 1) and a C-terminal helical region (domain 2). The stabilities of several GSH transferase dimers are dependent upon two groups of interactions between domains 1 and 2 of opposing subunits: a hydrophobic ball-and-socket motif and a buried charge cluster motif. In rGSTM1-1, these motifs involve residues F56 and R81, respectively. The structural basis for the effects of mutating F56 to different residues on dimer stability and function has been reported (Codreanu et al. (2005) Biochemistry 44, 10605-10612). Here, we show that the simultaneous disruption of both motifs in the F56S/R81A mutant causes complete dissociation of the dimer to a monomeric protein on the basis of gel filtration chromatography and multiple-angle laser light scattering. The fluorescence and far-UV CD properties of the double mutant as well as the kinetics of amide H/D exchange along the polypeptide backbone suggest that the monomer has a globular structure that is similar to a single subunit in the native protein. However, the mutant monomer has severely impaired catalytic activity, suggesting that the dimer interface is vital for efficient catalysis. Backbone amide H/D exchange kinetics in the F56S and F56S/R81A mutants indicate that a reorganization of the loop structure between helix alpha2 and strand beta3 near the active site is responsible for the decreased catalytic activity of the monomer. In addition, the junction between the alpha4 and alpha5 helices in F56S/R81R shows decreased H/D exchange, indicating another structural change that may affect catalysis. Although the native subunit interface is important for dimer stability, urea-induced unfolding of the F56S/R81A mutant suggests that the interface is not essential for the thermodynamic stability of individual subunits. The H/D exchange data reveal a possible molecular basis for the folding cooperativity observed between domains 1 and 2.  相似文献   

7.
The binding of denatured B. stearothermophilus D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH) to the E. coli chaperonin GroEL was investigated in two systems: (1) GroEL immobilized on Sepharose via a single subunit was titrated with urea-denatured soluble GAPDH and (2) a Sepharose-bound denatured GAPDH monomer was titrated with soluble GroEL. Similar apparent K D values for the complex GroEL·GAPDH were obtained in both cases (0.04 and 0.03 M, respectively), the stoichiometry being 1.0 mol chaperonin per GAPDH subunit in the system with the immobilized GroEL and 0.2 mol chaperonin per Sepharose-bound GAPDH monomer. Addition of GroEL and Mg·ATP to a reactivation mixture increased the yield of reactivation of both E. coli and B. stearothermophilus GAPDHs. Incubation of the Sepharose-bound catalytically active tetrameric and dimeric GAPDH forms with the protein fraction of a wild-type E. coli cell extract resulted in the binding of GroEL to the dimer and no interaction with the tetrameric form. These data suggest that GroEL may be capable of interacting with the interdimeric contact regions of the folded GAPDH dimers.  相似文献   

8.
Ornithine decarboxylase (ODC) catalyzes the decarboxylation of ornithine to putrescine and is the rate-limiting enzyme in the polyamine biosynthesis pathway. ODC is a dimeric enzyme, and the active sites of this enzyme reside at the dimer interface. Once the enzyme dissociates, the enzyme activity is lost. In this paper, we investigated the roles of amino acid residues at the dimer interface regarding the dimerization, protein stability and/or enzyme activity of ODC. A multiple sequence alignment of ODC and its homologous protein antizyme inhibitor revealed that 5 of 9 residues (residues 165, 277, 331, 332 and 389) are divergent, whereas 4 (134, 169, 294 and 322) are conserved. Analytical ultracentrifugation analysis suggested that some dimer-interface amino acid residues contribute to formation of the dimer of ODC and that this dimerization results from the cooperativity of these interface residues. The quaternary structure of the sextuple mutant Y331S/Y389D/R277S/D332E/V322D/D134A was changed to a monomer rather than a dimer, and the K d value of the mutant was 52.8 µM, which is over 500-fold greater than that of the wild-type ODC (ODC_WT). In addition, most interface mutants showed low but detectable or negligible enzyme activity. Therefore, the protein stability of these interface mutants was measured by differential scanning calorimetry. These results indicate that these dimer-interface residues are important for dimer formation and, as a consequence, are critical for enzyme catalysis.  相似文献   

9.
Thermal unfolding parameters were determined for a two-domain tetrameric enzyme, phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and for its isolated NAD(+)-binding domain. At pH 8.0, the transition temperatures (t(max)) for the apoforms of the native Bacillus stearothermophilus GAPDH and the isolated domain were 78.3 degrees C and 61.9 degrees C, with calorimetric enthalpies (DeltaH(cal)) of 4415 and 437 kJ/mol (or 30.7 and 22.1 J/g), respectively. In the presence of nearly saturating NAD(+) concentrations, the t(max) and the DeltaH(cal) increased by 13.6 degrees C and by 2365 kJ/mol, respectively, for the native apoenzyme, and by 2.8 degrees C and 109 kJ/mol for the isolated domain. These results indicate that interdomain interactions are essential for NAD(+) to produce its stabilizing effect on the structure of the native enzyme. The thermal stability of the isolated NAD(+)-binding domain increased considerably upon transition from pH 6.0 to 8.0. By contrast, native GAPDH exhibited greater stability at pH 6.0; similar pH-dependencies of thermal stability were displayed by GAPDHs isolated from rabbit muscle and Escherichia coli. The binding of NAD(+) to rabbit muscle apoenzyme increased t(max) and DeltaH(cal) and diminished the widths of the DSC curves; the effect was found to grow progressively with increasing coenzyme concentrations. Alkylation of the essential Cys149 with iodoacetamide destabilized the apoenzyme and altered the effect of NAD(+). Replacement of Cys149 by Ser or by Ala in the B. stearothermophilus GAPDH produced some stabilization, the effect of added NAD(+) being basically similar to that observed with the wild-type enzyme. These data indicate that neither the ion pairing between Cys149 and His176 nor the charge transfer interaction between Cys149 and NAD(+) make any significant contribution to the stabilization of the enzyme's native tertiary structure and the accomplishment of NAD(+)-induced conformational changes. The H176N mutant exhibited dramatically lower heat stability, as reflected in the values of both DeltaH(cal) and t(max). Interestingly, NAD(+) binding resulted in much wider heat capacity curves, suggesting diminished cooperativity of the unfolding transition.  相似文献   

10.
The patterns of energetic response elicited by single-site hemoglobin mutations and chemical mocdifications have been determined in order to probe the dimer–dimer interface of the half-ligated tetramer (species[21]) that was previously shown to behave as allosterically distinct from both the unligated and fully ligated molecules1. In this study the free energies of quaternary assembly(dimers to tetramers) were determined for aseries of 24 tetrameric species in which one dimeric half-molecule is ligated while the adjacent αβ dimer is unligated and contains a single amino acid modification. Assembly energies have also been determined for tetramers bearing the same amino acid modifications but where the hemesites were completely vacant and additionally where they were fully occupied. A total of 72 molecular species were thus characterized. It was found that mutationally induced perturbations to the free energy of quaternary assembly were identical for the half-ligated tetramers and the unligated tetramers over the entire spatial distrubution of altered sites, but exhibited a radically different pattern from that of the fully ligated molecules. These results indicate that the dimer–dimer interface of the half-ligated tetramer(species[21]) has the same quaternary sturcture as that of the unligated molecule, i.e, “quaternary T.” This quaternary structure assignment of species [21] strongly supports the operation of a Symmetry Rule which translates changes in hemesite ligation into six T → R quaternary switchpoints2. Analysis of the observed Symmetry Rule behaviour in relation to the measured distribution of cooperative free energies for the partially ligated species reveals significant cooperativity between α and β subunits of the dimeric half-tetramer within quaternary T. The mutational results indicate that these interactions are not “paid for” by breaking or making noncovalent bonds at the dimer–dimer interface (α1β2). They arise from structural and energetic changes that are “internal” to the ligated dimer even though its association with the unligated dimer is required for the cooperativity to occur. Free energy of “tertiary constraint” is thus generated by the first binding step and is propagated to the second hemesite while the dimer–dimer interface α1β2serves as a constraint. The “sequential” cooperativity that occurs within the half-molecule is thus preconditioned by the constraint of a quaternary T interface; release of this constraint by dissociation produces only noncooperative dimers. When the constraint is released functionally by T to R dimer rearrangement (at each switch-point specified by the a Symmetry Rule) the alterations of interfacial bonds then dominate the energetics of cooperativity. © 1993 Wiley-Liss, Inc.  相似文献   

11.
Precisely how mutant p53 exerts a dominant negative effect over wild type p53 has been an enigma. To understand how wild type and mutant p53 form hetero-oligomers, we studied p53 biogenesis in vitro. We show here that p53 dimers are formed cotranslationally (on the polysome), whereas tetramers are formed posttranslationally (by the dimerization of dimers in solution). Coexpression of wild type and mutant p53 therefore results in 50% of the p53 generated being heterotetramers comprised of a single species: wild type dimer/mutant dimer. Using hot spot mutants of p53 and a variety of natural target sites, we show that all wild type/mutant heterotetramers manifest impaired DNA binding activity. This impairment is not due to the mutant dimeric subunit inhibiting association of the complex with DNA but rather due to the lack of significant contribution (positive cooperativity) from the mutant partner. For all heterotetramers, bias in binding is particularly pronounced against those sequences in genes responsible for apoptosis rather than cell growth arrest. These results explain the molecular basis of p53 dominant negative effect and suggest a functional role in the regulation of p53 tetramerization.  相似文献   

12.
The carboxyl-terminal sequence of the lac repressor protein contains heptad repeats of leucines at positions 342, 349, and 356 that are required for tetramer assembly, as substitution of these leucine residues yields solely dimeric species (Chakerian, A. E., Tesmer, V. M., Manly, S. P., Brackett, J. K., Lynch, M. J., Hoh, J. T., and Matthews, K. S. (1991) J. Biol. Chem. 266, 1371-1374; Alberti, S., Oehler, S., von Wilcken-Bergmann, B., Kr?mer, H., and Müller-Hill, B. (1991) New Biol. 3, 57-62). To further investigate this region, which may form a leucine zipper motif, a family of lac repressor carboxyl-terminal deletion mutants eliminating the last 4, 5, 11, 18, and 32 amino acids (aa) has been constructed. The -4 aa mutant, in which all of the leucines in the presumed leucine zipper are intact, is tetrameric and displays operator and inducer binding properties similar to wild-type repressor. The -5 aa, -11 aa, -18 aa, and -32 aa deletion mutants, depleted of 1, 2, or all 3 of the leucines in the heptad repeats, are all dimeric, as demonstrated by gel filtration chromatography. Circular dichroism spectra and protease digestion studies indicate similar secondary/tertiary structures for the mutant and wild-type proteins. Differences in reaction with a monoclonal antibody specific for a subunit interface are observed for the dimeric versus tetrameric proteins, indicative of exposure of the target epitope as a consequence of deletion. Inducer binding properties of the deletion mutants are similar to wild-type tetrameric repressor at neutral pH. Only small differences in affinity and cooperativity from wild-type are evident at elevated pH; thus, the cooperative unit within the tetramer appears to be the dimer. "Apparent" operator binding affinity for the dimeric proteins is diminished, although minimal change in operator dissociation rate constants was observed. The diminution in apparent operator affinity may therefore derive from either 1) dissociation of the dimeric mutants to monomer generating a linked equilibrium or 2) alterations in intrinsic operator affinity of the dimers; the former explanation is favored. This detailed characterization of the purified mutant proteins confirms that the carboxyl-terminal region is involved in the dimer-dimer interface and demonstrates that cooperativity for inducer binding is contained within the dimer unit of the tetramer structure.  相似文献   

13.
Glycolytic glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is a NAD-dependent oxidoreductase which catalyzes the oxidative phosphorylation of d-glyceraldehyde-3-phosphate (G3P) to form 1, 3-diphosphoglycerate. The currently accepted mechanism involves an oxidoreduction step followed by a phosphorylation. GAPDH is classified as a B-specific oxidoreductase. The inspection of several crystal structures of GAPDHs indicates that the efficient hydride transfer from the hemithioacetal intermediate to the C4 position of the pyridinium si face requires optimal nicotinamidium-protein contacts for a suitable pyridinium-ring orientation. In previous studies carried out on Escherichia coli GAPDH (C. Corbier, A. Mougin, Y. Mely, H. W. Adolph, M. Zeppezauer, D. Gerard, A. Wonacott, and G. Branlant, Biochimie 72, 545-554, 1990; J. Eyschen, C. Corbier, B. Vitoux, G. Branlant, and M. T. Cung, Protein Pept. Lett. 1, 19-24, 1994), the role of the invariant Asn 313 residue, as an anchor which favors the syn orientation of the nicotinamide ring, was examined. Here, we report further investigations on the molecular factors responsible for the cofactor stereospecificity. Two single [Gly317] and [Ala317] GAPDH mutants and one double [Thr313-Gly317] GAPDH mutant were constructed on the basis of a molecular modelling study from the crystal structure of holo GAPDH from E. coli (E. Duée, L. Olivier-Deyris, E. Fanchon, C. Corbier, G. Branlant, and O. Dideberg, J. Mol. Biol. 257, 814-838, 1996). The Kd constants of [Ala317], [Gly317], and [Thr313-Gly317] GAPDH mutants for NAD are 5, 13, and 300 times higher than that of wild-type GAPDH. Transferred nuclear Overhauser effect spectroscopy demonstrates that the wild-type syn orientation of bound nicotinamide remains unchanged in the [Gly317] and [Ala317] mutants, whereas a conformational equilibrium between the syn and anti forms occurs in the [Thr313-Gly317] double mutant with a preference for the anti conformer. Although the double mutant preferably binds the nicotinamide ring in an anti conformation, it still exhibits B hydride transfer stereospecificity. Yet, the catalytic efficiency is much less than that of the wild type. This indicates that the holo GAPDH mutant fraction with an anti orientation of bound NAD is not capable of forming the ternary complex with G3P which would be required for an efficient A-specific catalytic process. The reasons of this catalytic inefficiency are discussed in relation with the historical and functional models which were advanced to explain the stereospecificity of NAD(P)-dependent dehydrogenases.  相似文献   

14.
The cytosolic NADP+-dependent malic enzyme (c-NADP-ME) has a dimer-dimer quaternary structure in which the dimer interface associates more tightly than the tetramer interface. In this study, the urea-induced unfolding process of the c-NADP-ME interface mutants was monitored using fluorescence and circular dichroism spectroscopy, analytical ultracentrifugation and enzyme activities. Here, we demonstrate the differential protein stability between dimer and tetramer interface interactions of human c-NADP-ME. Our data clearly demonstrate that the protein stability of c-NADP-ME is affected predominantly by disruptions at the dimer interface rather than at the tetramer interface. First, during thermal stability experiments, the melting temperatures of the wild-type and tetramer interface mutants are 8–10°C higher than those of the dimer interface mutants. Second, during urea denaturation experiments, the thermodynamic parameters of the wild-type and tetramer interface mutants are almost identical. However, for the dimer interface mutants, the first transition of the urea unfolding curves shift towards a lower urea concentration, and the unfolding intermediate exist at a lower urea concentration. Third, for tetrameric WT c-NADP-ME, the enzyme is first dissociated from a tetramer to dimers before the 2 M urea treatment, and the dimers then dissociated into monomers before the 2.5 M urea treatment. With a dimeric tetramer interface mutant (H142A/D568A), the dimer completely dissociated into monomers after a 2.5 M urea treatment, while for a dimeric dimer interface mutant (H51A/D90A), the dimer completely dissociated into monomers after a 1.5 M urea treatment, indicating that the interactions of c-NADP-ME at the dimer interface are truly stronger than at the tetramer interface. Thus, this study provides a reasonable explanation for why malic enzymes need to assemble as a dimer of dimers.  相似文献   

15.
To explore the structural mechanisms underlying the assembly and activation of family A GPCR dimers, we used the rat M(3) muscarinic acetylcholine receptor (M3R) as a model system. Studies with Cys-substituted mutant M3Rs expressed in COS-7 cells led to the identification of several mutant M3Rs that exclusively existed as cross-linked dimers under oxidizing conditions. The cross-linked residues were located at the bottom of transmembrane domain 5 (TM5) and within the N-terminal portion of the third intracellular loop (i3 loop). Studies with urea-stripped membranes demonstrated that M3R disulfide cross-linking did not require the presence of heterotrimeric G proteins. Molecular modeling studies indicated that the cross-linking data were in excellent agreement with the existence of a low-energy M3R dimer characterized by a TM5-TM5 interface. [(35)S]GTPγS binding/Gα(q/11) immunoprecipitation assays revealed that an M3R dimer that was cross-linked within the N-terminal portion of the i3 loop (264C) was functionally severely impaired (~50% reduction in receptor-G-protein coupling, as compared to control M3R). These data support the novel concept that agonist-induced activation of M3R dimers requires a conformational change of the N-terminal segment of the i3 loop. Given the high degree of structural homology among family A GPCRs, these findings should be of broad significance.  相似文献   

16.
The three-dimensional crystal structure of the (R207S, R292S) mutant of malate dehydrogenase from Haloarcula marismortui was solved at 1.95A resolution in order to determine the role of salt bridges and solvent ions in halophilic adaptation and quaternary structure stability. The mutations, located at the dimer-dimer interface, disrupt two inter-dimeric salt bridge clusters that are essential for wild-type tetramer stabilisation. Previous experiments in solution, performed on the double mutant, had shown a tetrameric structure in 4M NaCl, which dissociated into active dimers in 2M NaCl. In order to establish if the active dimeric form is a product of the mutation, or if it also exists in the wild-type protein, complementary studies were performed on the wild-type enzyme by analytical centrifugation and small angle neutron scattering experiments. They showed the existence of active dimers in NaF, KF, Na(2)SO(4), even in the absence of NADH, and in the presence of NADH at concentrations of NaCl below 0.3M. The crystal structure shows a tetramer that, in the absence of the salt bridge clusters, appears to be stabilized by a network of ordered water molecules and by Cl(-) binding at the dimer-dimer interface. The double mutant and wild-type dimer folds are essentially identical (the r.m.s. deviation between equivalent C(alpha) positions is 0.39A). Chloride ions are also observed at the monomer-monomer interfaces of the mutant, contributing to the stability of each dimer against low salt dissociation. Our results support the hypothesis that extensive binding of water and salt is an important feature of adaptation to a halophilic environment.  相似文献   

17.
The enzyme phenylalanine hydroxylase (PAH) is defective in the inherited disorder phenylketonuria. PAH, a tetrameric enzyme, is highly regulated and displays positive cooperativity for its substrate, Phe. Whether Phe binds to an allosteric site is a matter of debate, despite several studies worldwide. To address this issue, we generated a dimeric model for Phe–PAH interactions, by performing molecular docking combined with molecular dynamics simulations on human and rat wild-type sequences and also on a human G46S mutant. Our results suggest that the allosteric Phe-binding site lies at the dimeric interface between the regulatory and the catalytic domains of two adjacent subunits. The structural and dynamical features of the site were characterized in depth and described. Interestingly, our findings provide evidence for lower allosteric Phe-binding ability of the G46S mutant than the human wild-type enzyme. This also explains the disease-causing nature of this mutant.  相似文献   

18.
A mutation at the dimer interface of Plasmodium falciparum triosephosphate isomerase (PfTIM) was created by mutating a tyrosine residue at position 74, at the subunit interface, to glycine. Tyr74 is a critical residue, forming a part of an aromatic cluster at the interface. The resultant mutant, Y74G, was found to have considerably reduced stability compared with the wild-type protein (TIMWT). The mutant was found to be much less stable to denaturing agents such as urea and guanidinium chloride. Fluorescence and circular dichroism studies revealed that the Y74G mutant and TIMWT have similar spectroscopic properties, suggestive of similar folded structures. Further, the Y74G mutant also exhibited a concentration-dependent loss of enzymatic activity over the range 0.1-10 microM. In contrast, the wild-type enzyme did not show a concentration dependence of activity in this range. Fluorescence quenching of intrinsic tryptophan emission was much more efficient in case of Y74G than TIMWT, suggestive of greater exposure of Trp11, which lies adjacent to the dimer interface. Analytical gel filtration studies revealed that in Y74G, monomeric and dimeric species are in dynamic equilibrium, with the former predominating at low protein concentration. Spectroscopic studies established that the monomeric form of the mutant is largely folded. Low concentrations of urea also drive the equilibrium towards the monomeric form. These studies suggest that the replacement of tyrosine with a small residue at the interface of triosephosphate isomerase weakens the subunit-subunit interactions, giving rise to structured, but enzymatically inactive, monomers at low protein concentration.  相似文献   

19.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

20.
Here, we report the first crystal structure of a photosynthetic glyceraldehyde-3-phosphate dehydrogenase (GAPDH) complexed with NADP. The enzyme, purified from spinach chloroplasts, is constituted of a single type of subunit (A) arranged in homotetramers. It shows non-regulated NADP-dependent and NAD-dependent activities, with a preference for NADP. The structure has been solved to 3.0 A resolution by molecular replacement. The crystals belong to space group C222 with three monomers in the asymmetric unit. One of the three monomers generates a tetramer using the space group 222 point symmetry and a very similar tetramer is generated by the other two monomers, related by a non-crystallographic symmetry, using a crystallographic 2-fold axis.The protein reveals a large structural homology with known GAPDHs both in the cofactor-binding domain and in regions of the catalytic domain. Like all other GAPDHs investigated so far, the A(4)-GAPDH belongs to the Rossmann fold family of dehydrogenases. However, unlike most dehydrogenases of this family, the adenosine 2'-phosphate group of NADP does not form a salt-bridge with any positively charged residue in its surroundings, being instead set in place by hydrogen bonds with a threonine residue belonging to the Rossmann fold and a serine residue located in the S-loop of a symmetry-related monomer. While increasing our knowledge of an important photosynthetic enzyme, these results contribute to a general understanding of NADP versus NAD recognition in pyridine nucleotide-dependent enzymes.Although the overall structure of A(4)-GAPDH is similar to that of the cytosolic GAPDH from bacteria and eukaryotes, the chloroplast tetramer is peculiar, in that it can actually be considered a dimer of dimers, since monomers are bound in pairs by a disulphide bridge formed across Cys200 residues. This bridge is not found in other cytosolic or chloroplast GAPDHs from animals, bacteria, or plants other than spinach.  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