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 375 毫秒
1.
The native dimeric form of enolase from pig muscle was immobilized on Sepharose 4B activated with cyanogen bromide. The amount of matrix-bound enolase, its specific activity and kinetic properties depend on the extent of gel activation with CNBr. Only on the Sepharose activated with small quantities of CNBr the amount of protein which remained after treatment with Gdn.HCl was about 50% of the initially bound enolase, indicating that the enzyme was bound covalently to the matrix through a single subunit. The matrix-bound monomers obtained in this way were inactive and were unable to reassociate to dimers on addition of free subunits. The matrix-bound monomers obtained after KBr treatment were inactive but retained the ability to reassociate into active dimers after addition of free subunits. The results indicate that single matrix-bound subunits of pig muscle enolase are enzymatically inactive and dimeric structure is essential for catalytic activity.  相似文献   

2.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  相似文献   

3.
Limited proteolysis of DsbC with trypsin resulted in a compact and stable C-terminal fragment (residues 66-216), fDsbC, which retains the active site sequence, -Cys(98)-Gly-Tyr-Cys(101)-, and shows only minor differences in conformation compared with that of the intact molecule. The pK(a) of active site thiol and the K(SS) with glutathione are very close to that of DsbC, respectively; however, fDsbC is inactive as an isomerase in catalyzing the formation of correct disulfide bonds in scrambled RNase A and denatured and reduced bovine pancreatic trypsin inhibitor and shows only 13% thiol-protein oxidoreductase activity (TPOR) of DsbC. In contrast to the dimeric DsbC, fDsbC exists as a monomer and has no chaperone activity in assisting the reactivation of denatured d-glyceraldehyde-3-phosphate dehydrogenase. The heterodimer of DsbC with the inactive DsbC carboxymethylated at both active site thiols shows about 50% TPOR activity of DsbC but no isomerase activity, indicating that the DsbC subunit in the heterodimer displays full TPOR activity but little, if any, isomerase activity. It is concluded that the N-terminal sequence (residues 1-65) is essential for dimer formation and chaperone activity of DsbC. The active sites in both subunits of the dimeric DsbC appear to be essential for its isomerase activity.  相似文献   

4.
Gibson LM  Lovelace LL  Lebioda L 《Biochemistry》2008,47(16):4636-4643
Loop 181-197 of human thymidylate synthase (hTS) populates two conformational states. In the first state, Cys195, a residue crucial for catalytic activity, is in the active site (active conformer); in the other conformation, it is about 10 A away, outside the active site (inactive conformer). We have designed and expressed an hTS variant, R163K, in which the inactive conformation is destabilized. The activity of this mutant is 33% higher than that of wt hTS, suggesting that at least one-third of hTS populates the inactive conformer. Crystal structures of R163K in two different crystal forms, with six and two subunits per asymmetric part of the unit cells, have been determined. All subunits of this mutant are in the active conformation while wt hTS crystallizes as the inactive conformer in similar mother liquors. The structures show differences in the environment of catalytic Cys195, which correlate with Cys195 thiol reactivity, as judged by its oxidation state. Calculations show that the molecular electrostatic potential at Cys195 differs between the subunits of the dimer. One of the dimers is asymmetric with a phosphate ion bound in only one of the subunits. In the absence of the phosphate ion, that is in the inhibitor-free enzyme, the tip of loop 47-53 is about 11 A away from the active site.  相似文献   

5.
Human neuron-specific enolase (NSE) or isozyme gamma has been expressed with a C-terminal His-tag in Escherichia coli. The enzyme has been purified, crystallized and its crystal structure determined. In the crystals the enzyme forms the asymmetric complex NSE x Mg2 x SO4/NSE x Mg x Cl, where "/" separates the dimer subunits. The subunit that contains the sulfate (or phosphate) ion and two magnesium ions is in the closed conformation observed in enolase complexes with the substrate or its analogues; the other subunit is in the open conformation observed in enolase subunits without bound substrate or analogues. This indicates negative cooperativity for ligand binding between subunits. Electrostatic charge differences between isozymes alpha and gamma, -19 at physiological pH, are concentrated in the regions of the molecular surface that are negatively charged in alpha, i.e. surface areas negatively charged in alpha are more negatively charged in gamma, while areas that are neutral or positively charged tend to be charge-conserved.  相似文献   

6.
Lipoprotein lipase (LPL), a key enzyme in the metabolism of triglyceride-rich plasma lipoproteins, is a homodimer. Dissociation to monomers leads to loss of activity. Evidence that LPL dimers rapidly exchange subunits was demonstrated by fluorescence resonance energy transfer between lipase subunits labeled with Oregon Green and tetrametylrhodamine, respectively, and also by formation of heterodimers composed of radiolabeled and biotinylated lipase subunits captured on streptavidine-agarose. Compartmental modeling of the inactivation kinetics confirmed that rapid subunit exchange must occur. Studies of activity loss indicated the existence of a monomer that can form catalytically active dimers, but this intermediate state has not been possible to isolate and remains hypothetical. Differences in solution properties and conformation between the stable but catalytically inactive monomeric form of LPL and the active dimers were studied by static light scattering, intrinsic fluorescence, and probing with 4,4'-dianilino-1,1'-binaphtyl-5,5'-disulfonic acid and acrylamide. The catalytically inactive monomer appeared to have a more flexible and exposed structure than the dimers and to be more prone to aggregation. By limited proteolysis the conformational changes accompanying dissociation of the dimers to inactive monomers were localized mainly to the central part of the subunit, probably corresponding to the region for subunit interaction.  相似文献   

7.
Most enolases are homodimers. There are a few that are octamers, with the eight subunits arranged as a tetramer of dimers. These dimers have the same basic fold and same subunit interactions as are found in the dimeric enolases. The dissociation of the octameric enolase from S. pyogenes was examined, using NaClO4, a weak chaotrope, to perturb the quaternary structure. Dissociation was monitored by sedimentation velocity. NaClO4 dissociated the octamer into inactive monomers. There was no indication that dissociation of the octamer into monomers proceeded via formation of significant amounts of dimer or any other intermediate species. Two mutations at the dimer-dimer interface, F137L and E363G, were introduced in order to destabilize the octameric structure. The double mutant was more easily dissociated than was the wild type. Dissociation could also be produced by other salts, including tetramethylammonium chloride (TMACl) or by increasing pH. In all cases, no significant amounts of dimers or other intermediates were formed. Weakening one interface in this protein weakened the other interface as well. Although enolases from most organisms are dimers, the dimeric form of the S. pyogenes enzyme appears to be unstable.  相似文献   

8.
Thoresen T  Gelles J 《Biochemistry》2008,47(36):9514-9521
A single molecule of the motor enzyme kinesin-1 keeps a tight grip on its microtubule track, making tens or hundreds of discrete, unidirectional 8 nm steps before dissociating. This high duty ratio processive movement is thought to require a mechanism in which alternating stepping of the two head domains of the kinesin dimer is driven by alternating, overlapped cycles of ATP hydrolysis by the two heads. The R210K point mutation in Drosophila kinesin heavy chain was reported to disrupt the ability of the enzyme active site to catalyze ATP P-O bond cleavage. We expressed R210K homodimers as well as isolated R210K heads and confirmed that both are essentially inactive. We then coexpressed tagged R210K subunits with untagged wild-type subunits and affinity purified R210K/wild-type heterodimers together with the inactive R210K homodimers. In contrast to the R210K head or homodimer, the heterodimer was a highly active (>50% of wild-type) microtubule-stimulated ATPase, and the heterodimer displayed high duty ratio processive movement in single-molecule motility experiments. Thus, dimerization of a subunit containing the inactivating mutation with a functional subunit can complement the mutation; this must occur either by lowering or by bypassing kinetic barriers in the ATPase or mechanical cycles of the mutant head. The observations provide support for kinesin-1 gating mechanisms in which one head stimulates the rate of essential processes in the other.  相似文献   

9.
Wild-type thymidylate synthase (WT-TS) from Escherichia coli and several of its mutants showed varying degrees of susceptibility to trypsin. While WT-TS was resistant to trypsin as were the mutants C146S, K48E, and R126K, others such as Y94A, Y94F, C146W, and R126E were digested but at different rates from one another. The peptides released from the mutants were identified by mass spectrometry and Edman sequence analysis. The known crystal structures for WT-TS, Y94F, and R126E, surprisingly, showed no structural differences that could explain the difference in their susceptibility to trypsin. One explanation is that the mutations could perturb the dynamic equilibrium of the dimeric state of the mutants as to increase their dissociation to monomers, which being less structured than the dimer, would be hydrolyzed more readily by trypsin. Earlier studies appear to support this proposal since conditions that promote subunit dissociation in solutions of R126E with other inactive mutants, such as dilution, low concentrations of urea, and elevated pH, greatly enhance the rate of restoration of TS activity. Analytic ultracentrifuge studies with various TSs in urea, or at pH 9.0, or that have been highly diluted are, for the most part, in agreement with this thesis, since these conditions are associated with an increase in dissociation to monomers, particularly with the mutant TSs. However, these studies do not rule out the possibility that conformation differences among the various TS dimers are responsible for the differences in susceptibility to trypsin, particularly at high concentrations of protein where the WT-TS and mutants are mainly dimers.  相似文献   

10.
Soluble guanylate cyclase (sGC) catalyses the synthesis of cyclic GMP in response to nitric oxide. The enzyme is a heterodimer of homologous α and β subunits, each of which is composed of multiple domains. We present here crystal structures of a heterodimer of the catalytic domains of the α and β subunits, as well as an inactive homodimer of β subunits. This first structure of a metazoan, heteromeric cyclase provides several observations. First, the structures resemble known structures of adenylate cyclases and other guanylate cyclases in overall fold and in the arrangement of conserved active-site residues, which are contributed by both subunits at the interface. Second, the subunit interaction surface is promiscuous, allowing both homodimeric and heteromeric association; the preference of the full-length enzyme for heterodimer formation must derive from the combined contribution of other interaction interfaces. Third, the heterodimeric structure is in an inactive conformation, but can be superposed onto an active conformation of adenylate cyclase by a structural transition involving a 26° rigid-body rotation of the α subunit. In the modelled active conformation, most active site residues in the subunit interface are precisely aligned with those of adenylate cyclase. Finally, the modelled active conformation also reveals a cavity related to the active site by pseudo-symmetry. The pseudosymmetric site lacks key active site residues, but may bind allosteric regulators in a manner analogous to the binding of forskolin to adenylate cyclase. This indicates the possibility of developing a new class of small-molecule modulators of guanylate cyclase activity targeting the catalytic domain.  相似文献   

11.
Purified enolase from Bacillus subtilis has a native mass of approximately 370 kDa. Since B. subtilis enolase was found to have a subunit mass of 46.58 kDa, the quaternary structure of B. subtilis is octameric. The pl for B. subtilis enolase is 6.1, the pH optimum (pHo) for activity is 8.1–8.2, and the K m for 2-PGA is approximately 0.67 mM. Using the dimeric Cα structure of yeast dimeric enolase as a guide, these dimers were arranged as a tetramer of dimers to simulate the electron microscopy image processing obtained for the octameric enolase purified from Thermotoga maritima. This arrangement allowed identification of helix J of one dimer (residues 86–96) and the loop between helix L and strand 1 (HL–S1 loop) of another dimer as possible subunit interaction regions. Alignment of available enolase amino acid sequences revealed that in 16 there are two tandem glycines at the C-terminal end of helix L and the HL–S1 loop is truncated by 4–6 residues relative to the yeast polypeptide, two structural features absent in enolases known to be dimers. From these arrangements and alignments it is proposed that the GG tandem at the C-terminal end of helix L and truncation of the HL–S1 loop may play a critical role in octamer formation of enolases. Interestingly, the sequence features associated with dimeric quaternary structure are found in three phylogenetically disparate groups, suggesting that the ancestral enolase was an octamer and that the dimeric structure has arisen independently multiple times through evolutionary history.  相似文献   

12.
The isolated subunites of Eurypelma californicum hemocyanin were studied by aid of antibodies raised against whole, dissociated hemocyanin. The proportion of impurities was found to be low in almost all subunits. There was no cross reaction between the individual chains, and the total number of antigenically different subunits was found to be seven, confirming results obtained by different methods. If an artificial mixture prepared from purified subunits is compared to whole, dissociated hemocyanin, an overall very similar pattern is obtained but differences appear which are due to specific interaction.--The dimeric subunit 4D was shown to be a heterodimer (asymmetric dimer) composed of chains b and c4.  相似文献   

13.
The crystal structure of Escherichia coli enolase (EC 4.2.1.11, phosphopyruvate hydratase), which is a component of the RNA degradosome, has been determined at 2.5 A. There are four molecules in the asymmetric unit of the C2 cell, and in one of the molecules, flexible loops close onto the active site. This closure mimics the conformation of the substrate-bound intermediate. A comparison of the structure of the E. coli enolase with the eukaryotic enolase structures available (lobster and yeast) indicates a high degree of conservation of the hydrophobic core and the subunit interface of this homodimeric enzyme. The dimer interface is enriched in charged residues compared with other protein homodimers, which may explain our observations from analytical ultracentrifugation that dimerisation is affected by ionic strength. The putative role of enolase in the RNA degradosome is discussed; although it was not possible to ascribe a specific role to it, a structural role is possible.  相似文献   

14.
Purified enolase from Bacillus subtilis has a native mass of approximately 370 kDa. Since B. subtilis enolase was found to have a subunit mass of 46.58 kDa, the quaternary structure of B. subtilis is octameric. The pl for B. subtilis enolase is 6.1, the pH optimum (pHo) for activity is 8.1–8.2, and the K m for 2-PGA is approximately 0.67 mM. Using the dimeric C structure of yeast dimeric enolase as a guide, these dimers were arranged as a tetramer of dimers to simulate the electron microscopy image processing obtained for the octameric enolase purified from Thermotoga maritima. This arrangement allowed identification of helix J of one dimer (residues 86–96) and the loop between helix L and strand 1 (HL–S1 loop) of another dimer as possible subunit interaction regions. Alignment of available enolase amino acid sequences revealed that in 16 there are two tandem glycines at the C-terminal end of helix L and the HL–S1 loop is truncated by 4–6 residues relative to the yeast polypeptide, two structural features absent in enolases known to be dimers. From these arrangements and alignments it is proposed that the GG tandem at the C-terminal end of helix L and truncation of the HL–S1 loop may play a critical role in octamer formation of enolases. Interestingly, the sequence features associated with dimeric quaternary structure are found in three phylogenetically disparate groups, suggesting that the ancestral enolase was an octamer and that the dimeric structure has arisen independently multiple times through evolutionary history.  相似文献   

15.
Bovine seminal ribonuclease (BS-RNase) is a dimeric protein with two identical subunits linked by two disulfide bridges, each subunit showing 80% of sequence identity with pancreatic RNase A. BS-RNase exists in two different quaternary conformations in solution: the MxM form, in which each subunit exchanges its alpha-helical N-terminal segment with its partner, and the M=M form with no exchange. By differential scanning microcalorimetry (DSC), the denaturation of the two dimeric forms of BS-RNase was found to be more complex than a simple two-state process. Monomeric derivatives of the dimeric protein follow instead a simple two-state mechanism, but are distinctly less stable than RNase A. The three-state N if I if D denaturation process of the two quaternary isoforms was interpreted by identifying in the dimers a central highly structured core, enclosing the covalently bonded subunit interface, which unfolds only after the periphery (mainly the N-terminal peptide) unfolds. Circular dichroism spectra of the two forms in the far-ultraviolet region show large differences between the secondary structure of the isoforms and that of the native BS-RNase mixture at equilibrium. This has been attributed to the presence in the equilibrium mixture of intermediate forms with displaced and disordered N-terminal alpha-helical segments.  相似文献   

16.
E M Reimann 《Biochemistry》1986,25(1):119-125
The type II adenosine cyclic 3',5'-phosphate (cAMP) dependent protein kinase from bovine heart, consisting of a dimeric regulatory subunit and two catalytic subunits, was converted to a heterodimer by limited tryptic digestion. Loss of the tetrameric structure was accompanied by proteolysis of the regulatory subunit to a form with an apparent molecular weight of 45 000 vs. 52 000 for the native subunit. The proteolyzed subunit behaved as a monomer, in contrast to the dimeric native subunit. Amino acid sequence analysis established that proteolysis removed 45 residues at the N-terminus, indicating that these 45 residues constitute the dimerizing domain of this protein. The kinetic properties of this heterodimer were indistinguishable from those of the native tetramer: half-maximal kinase activation occurred at 48 nM cAMP with a Hill coefficient of 1.45, the regulatory subunit bound 1.5 equiv of cAMP with half-maximal binding occurring at 33 nM, and kinetics for dissociation of bound cAMP were biphasic, indicating the presence of two different binding sites. These observations suggest that residues 1-45 function only in the formation of dimers and that dimerization has little influence on other functional properties of the regulatory subunit. More extensive proteolysis cleaved the monomeric fragment at Lys-311. The fragments resulting from this second cleavage did not dissociate, and the complex inhibited the catalytic subunit in a cAMP-dependent manner.  相似文献   

17.
Sulfonylurea receptors (SURs) associate with Kir6.x subunits to form tetradimeric K(ATP) channel complexes. SUR1 and SUR2 confer differential channel sensitivities to nucleotides and pharmacological agents, and are expressed in specific, but overlapping, tissues. This raises the question of whether these different SUR subtypes can assemble in the same channel complex and generate channels with hybrid properties. To test this, we engineered dimeric constructs of wild type or N160D mutant Kir6.2 fused to SUR1 or SUR2A. Dimeric fusions formed functional, ATP-sensitive, channels. Coexpression of weakly rectifying SUR1-Kir6.2 (WTF-1) with strongly rectifying SUR1-Kir6.2[N160D] (NDF-1) in COSm6 cells results in mixed subunit complexes that exhibit unique rectification properties. Coexpression of NDF-1 and SUR2A-Kir6.2 (WTF-2) results in similar complex rectification, reflecting the presence of SUR1- and SUR2A-containing dimers in the same channel. The data demonstrate clearly that SUR1 and SUR2A subunits associate randomly, and suggest that heteromeric channels will occur in native tissues.  相似文献   

18.
Koyama T  Gotoh Y  Nishino T 《Biochemistry》2000,39(2):463-469
Farnesyl diphosphate synthase is a homodimer of subunits having typically two aspartate-rich motifs with two sets of substrate binding sites for an allylic diphosphate and isopentenyl diphosphate per molecule of a homodimeric enzyme. To determine whether each subunit contains an independent active site or whether the active sites are created by intersubunit interaction, we constructed several expression plasmids that overproduce hybrid-type heterodimers of Bacillus stearothermophilus FPP synthases constituting different types of mutated monomers, which exhibit little catalytic activity as homodimers, by combining two tandem fps genes for the manipulated monomer subunit with a highly efficient promoter trc within an overexpression pTrc99A plasmid. A heterodimer of a combination of subunits of the wild type and of R98E, a mutant subunit which exhibits little enzymatic activity as a dimer form (R98E)(2), exhibited 78% of the activity of the wild-type homodimer enzyme, (WT)(2). Moreover, when a hybrid-type heterodimeric dimer of FPP synthase mutant subunits (R98E/F220A) was prepared, the FPP synthase activity was 18- and 390-fold of that of each of the almost inactive mutants as a dimeric enzymes, (R98E)(2) and (F220A)(2) [Koyama, T., et al. (1995) Biochem. Biophys. Res. Commun. 212, 681-686], respectively. These results suggest that the subunits of the FPP synthase interact with each other to form a shared active site in the homodimer structure rather than an independent active site in each subunit.  相似文献   

19.
The three-dimensional structure of an R-state conformer of glycogen phosphorylase containing the coenzyme-substrate analog pyridoxal-5'-diphosphate at the catalytic site (PLPP-GPb) has been refined by X-ray crystallography to a resolution of 2.87 A. The molecule comprises four subunits of phosphorylase related by approximate 222 symmetry. Whereas the quaternary structure of R-state PLPP-GPb is similar to that of phosphorylase crystallized in the presence of ammonium sulfate (Barford, D. & Johnson, L.N., 1989, Nature 340, 609-616), the tertiary structures differ in that the two domains of the PLPP-GPb subunits are rotated apart by 5 degrees relative to the T-state conformation. Global differences among the four subunits suggest that the major domains of the phosphorylase subunit are connected by a flexible hinge. The two different positions observed for the terminal phosphate of the PLPP are interpreted as distinct phosphate subsites that may be occupied at different points along the reaction pathway. The structural basis for the unique ability of R-state dimers to form tetramers results from the orientation of subunits with respect to the dyad axis of the dimer. Residues in opposing dimers are in proper registration to form tetramers only in the R-state.  相似文献   

20.
Each of the two active sites of thymidylate synthase contains amino acid residues contributed by the other subunit. For example, Arg-178 of one monomer binds the phosphate group of the substrate dUMP in the active site of the other monomer [Hardy et al. (1987) Science 235, 448-455]. Inactive mutants of such residues should combine with subunits of other inactive mutants to form heterodimeric hybrids with one functional active site. In vivo and in vitro approaches were used to test this hypothesis. In vivo complementation was accomplished by cotransforming plasmid mixtures encoding pools of inactive Arg-178 mutants and pools of inactive Cys-198 mutants into a host strain deficient in thymidylate synthase. Individual inactive mutants of Arg-178 were also cotransformed with the C198A mutant. Subunit complementation was detected by selection or screening for transformants which grew in the absence of thymidine, and hence produced active enzyme. Many mutants at each position representing a wide variety of size and charge supported subunit complementation. In vitro complementation was accomplished by reversible dissociation and unfolding of mixtures of purified individual inactive Arg-178 and Cys-198 mutant proteins. With the R178F + C198A heterodimer, the Km values for dUMP and CH2H4folate were similar to those of the wild-type enzyme. By titrating C198A with R178F under unfolding-refolding conditions, we were able to calculate the kcat value for the active heterodimer. The catalytic efficiency of the single wild-type active site of the C198A + R178F heterodimer approaches that of the wild-type enzyme.  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