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.
BACKGROUND: Oligomeric proteins may have been selected for in hyperthermophiles because subunit association provides extra stabilization. Phosphoribosylanthranilate isomerase (PRAI) is monomeric and labile in most mesophilic microorganisms, but dimeric and stable in the hyperthermophile Thermotoga maritima (tPRAI). The two subunits of tPRAI are associated back-to-back and are locked together by a hydrophobic loop. The hypothesis that dimerization is important for thermostability has been tested by rationally designing monomeric variants of tPRAI. RESULTS: The comparison of tPRAI and PRAI from Escherichia coli (ePRAI) suggested that levelling the nonplanar dimer interface would weaken the association. The deletion of two residues in the loop loosened the dimer. Subsequent filling of the adjacent pocket and the exchange of polar for apolar residues yielded a weakly associating and a nonassociating monomeric variant. Both variants are as active as the parental dimer but far more thermolabile. The thermostability of the weakly associating monomer increased significantly with increasing protein concentration. The X-ray structure of the nonassociating monomer differed from that of the parental subunit only in the restructured interface. The orientation of the original subunits was maintained in a crystal contact between two monomers. CONCLUSIONS: tPRAI is dimeric for reasons of stability. The clearly separated responsibilities of the betaalpha loops, which are involved in activity, and the alphabeta loops, which are involved in protein stability, has permitted the evolution of dimers without compromising their activity. The preserved interaction in the crystal contacts suggests the most likely model for dimer evolution.  相似文献   

2.
A central theme in prion protein research is the detection of the process that underlies the conformational transition from the normal cellular prion form (PrP(C)) to its pathogenic isoform (PrP(Sc)). Although the three-dimensional structures of monomeric and dimeric human prion protein (HuPrP) have been revealed by NMR spectroscopy and x-ray crystallography, the process underlying the conformational change from PrP(C) to PrP(Sc) and the dynamics and functions of PrP(C) remain unknown. The dimeric form is thought to play an important role in the conformational transition. In this study, we performed molecular dynamics (MD) simulations on monomeric and dimeric HuPrP at 300 K and 500 K for 10 ns to investigate the differences in the properties of the monomer and the dimer from the perspective of dynamic and structural behaviors. Simulations were also undertaken with Asp178Asn and acidic pH, which is known as a disease-associated factor. Our results indicate that the dynamics of the dimer and monomer were similar (e.g., denaturation of helices and elongation of the beta-sheet). However, additional secondary structure elements formed in the dimer might result in showing the differences in dynamics and properties between the monomer and dimer (e.g., the greater retention of dimeric than monomeric tertiary structure).  相似文献   

3.
Dimeric form of diphtheria toxin: purification and characterization   总被引:10,自引:0,他引:10  
Many preparations of diphtheria toxin were found to contain dimeric and multimeric toxin forms. The monomeric and dimeric forms were fractionated to greater than 98% purity, and their properties were compared. Dimeric toxin slowly dissociated to native monomers in solution at neutral pH and could be rapidly dissociated with dimethyl sulfoxide. In cell culture assays and rabbit skin tests, the dimer exhibited no significant toxic activity, except for that attributable to trace contamination by monomer, or partial dissociation to monomer during the incubation period. In guinea pig lethality tests, however, toxic activity varied depending upon the dose. At least 7-fold greater amounts of dimer than monomer (161 ng vs. 22 ng, respectively) were required to cause death at 18 h, whereas similar weights of the two toxin forms (22 ng) caused death at 120 h. This variability probably reflected slow dissociation of dimer to monomer in the animal. The dimer was unable to bind toxin receptors on the surface of susceptible cells, whereas it retained full activity in the ADP-ribosyltransferase, NAD-glycohydrolase, or ligand-binding assays. Thus, the lack of toxicity of the dimeric toxin may have resulted from distortion or occlusion of the receptor binding site on the B moiety. We propose that the dimer contains two monomeric units bound by hydrophobic interactions and that the points of contact involve regions of the B moieties that are normally buried in the native monomer.  相似文献   

4.
Antizyme inhibitor (AzI) regulates cellular polyamine homeostasis by binding to the polyamine-induced protein, Antizyme (Az), with greater affinity than ornithine decarboxylase (ODC). AzI is highly homologous to ODC but is not enzymatically active. In order to understand these specific characteristics of AzI and its differences from ODC, we determined the 3D structure of mouse AzI to 2.05 A resolution. Both AzI and ODC crystallize as a dimer. However, fewer interactions at the dimer interface, a smaller buried surface area, and lack of symmetry of the interactions between residues from the two monomers in the AzI structure suggest that this dimeric structure is nonphysiological. In addition, the absence of residues and interactions required for pyridoxal 5'-phosphate (PLP) binding suggests that AzI does not bind PLP. Biochemical studies confirmed the lack of PLP binding and revealed that AzI exists as a monomer in solution while ODC is dimeric. Our findings that AzI exists as a monomer and is unable to bind PLP provide two independent explanations for its lack of enzymatic activity and suggest the basis for its enhanced affinity toward Az.  相似文献   

5.
The serine hydroxymethyltransferase from Bacillus subtilis (bsSHMT) and B. stearothermophilus (bstSHMT) are both homodimers and share approximately 77% sequence identity; however, they show very different thermal stabilities and unfolding pathways. For investigating the role of N- and C-terminal domains in stability and unfolding of dimeric SHMTs, we have swapped the structural domains between bs- and bstSHMT and generated the two novel chimeric proteins bsbstc and bstbsc, respectively. The chimeras had secondary structure, tyrosine, and pyridoxal-5'-phosphate microenvironment similar to that of the wild-type proteins. The chimeras showed enzymatic activity slightly higher than that of the wild-type proteins. Interestingly, the guanidium chloride (GdmCl)-induced unfolding showed that unlike the wild-type bsSHMT, which undergoes dissociation of native dimer into monomers at low guanidium chloride (GdmCl) concentration, resulting in a non-cooperative unfolding of enzyme, its chimera bsbstc, having the C-terminal domain of bstSHMT was resistant to low GdmCl concentration and showed a GdmCl-induced cooperative unfolding from native dimer to unfolded monomer. In contrast, the wild-type dimeric bstSHMT was resistant to low GdmCl concentration and showed a GdmCl-induced cooperative unfolding, whereas its chimera bstbsc, having the C- terminal domain of bsSHMT, showed dissociation of native dimer into monomer at low GdmCl concentration and a GdmCl-induced non-cooperative unfolding. These results clearly demonstrate that the C-terminal domain of dimeric SHMT plays a vital role in stabilization of the oligomeric structure of the native enzyme hence modulating its unfolding pathway.  相似文献   

6.
The structure of the potent HIV-inactivating protein cyanovirin-N was previously found by NMR to be a monomer in solution and a domain-swapped dimer by X-ray crystallography. Here we demonstrate that, in solution, CV-N can exist both in monomeric and in domain-swapped dimeric form. The dimer is a metastable, kinetically trapped structure at neutral pH and room temperature. Based on orientational NMR constraints, we show that the domain-swapped solution dimer is similar to structures in two different crystal forms, exhibiting solely a small reorientation around the hinge region. Mutation of the single proline residue in the hinge to glycine significantly stabilizes the protein in both its monomeric and dimeric forms. By contrast, mutation of the neighboring serine to proline results in an exclusively dimeric protein, caused by a drastic destabilization of the monomer.  相似文献   

7.
The crystallographic dimer of the C-terminal fragment (CTF) of the L7/L12 ribosomal protein has been subjected to molecular dynamics (MD) simulations. A 90 picosecond (ps) trajectory for the protein dimer, 19 water molecules and two counter ions has been calculated at constant temperature. Effects of intermolecular interactions on the structure and dynamics have been studied. The exact crystallographic symmetry is lost and the atomic fluctuations differ from one monomer to the other. The average MD structure is more stable than the X-ray one, as judged by accessible surface area and energy calculations. Crystal (non-dimeric) interactions have been simulated in another 40 ps trajectory by using harmonic restraints to represent intermolecular hydrogen bonds. The conformational changes with respect ot the X-ray structure are then virtually suppressed.The unrestrained dimer trajectory has been scanned for cooperative motions involving secondary structure elements. The intrinsic collective motions of the monomer are transmitted via intermolecular contacts to the dimer structure.The existence of a stable dimeric form of CTF, resembling the crystallographic one, has been documented. At the cost of fairly small energy expenditure the dimer has considerable conformational flexibility. This flexibility may endow the dimer with some functional potential as an energy transducer.  相似文献   

8.
Lee JY  Yang ST  Lee SK  Jung HH  Shin SY  Hahm KS  Kim JI 《The FEBS journal》2008,275(15):3911-3920
The cathelicidin antimicrobial peptide bactenecin is a beta-hairpin molecule with a single disulfide bond and broad antimicrobial activity. The proform of bactenecin exists as a dimer, however, and it has been proposed that bactenecin is released as a dimer in vivo, although there has been little study of the dimeric form of bactenecin. To investigate the effect of bactenecin dimerization on its biological activity, we characterized the dimer's effect on phospholipid membranes, the kinetics of its bactericidal activity, and its salt sensitivity. We initially synthesized two bactenecin dimers (antiparallel and parallel) and two monomers (beta-hairpin and linear). Under oxidative folding conditions, reduced linear bactenecin preferentially folded into a dimer forming a ladder-like structure via intermolecular disulfide bonding. As compared to the monomer, the dimer had a greater ability to induce lysis of lipid bilayers and was more rapidly bactericidal. Interestingly, the dimer retained antimicrobial activity at physiological salt concentrations (150 mm NaCl), although the monomer was inactivated. This salt resistance was also seen with bactenecin dimer containing one intermolecular disulfide bond, and the bactenecin dimer appears to undergo multimeric oligomerization at high salt concentrations. Overall, dimeric bactenecin shows potent and rapid antimicrobial activity, and resists salt-induced inactivation under physiological conditions through condensation and oligomerization. These characteristics shed light on the features that a peptide would need to serve as an effective therapeutic agent.  相似文献   

9.
A mutant of the dimeric rabbit muscle creatine kinase (MM-CK) in which tryptophan 210 was replaced has been studied to assess the role of this residue in dimer cohesion and the importance of the dimeric state for the native enzyme stability. Wild-type protein equilibrium unfolding induced by guanidine hydrochloride occurs through intermediate states with formation of a molten globule and a premolten globule. Unlike the wild-type enzyme, the mutant inactivates at lower denaturant concentration and the loss of enzymatic activity is accompanied by the dissociation of the dimer into two apparently compact monomers. However, the Stokes radius of the monomer increases with denaturant concentration as determined by size exclusion chromatography, indicating that, upon monomerization, the protein structure is destabilized. Binding of 8-anilinonaphthalene-1-sulfonate shows that the dissociated monomer exposes hydrophobic patches at its surface, suggesting that it could be a molten globule. At higher denaturant concentrations, both wild-type and mutant follow similar denaturation pathways with formation of a premolten globule around 1.5-M guanidine, indicating that tryptophan 210 does not contribute to a large extent to the monomer conformational stability, which may be ensured in the dimeric state through quaternary interactions. Proteins 32:43–51, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
Extending previous work (Sung & Jordan, 1987 a, Biophys. J. 51, 661-672; 1988, Biophys. J.54, 519-526), we describe channel properties of five possible gramicidin dimers by studying dimerization energies and axial electrical potentials. Unlike the head-to-head dimer (the predominant channel former), both tail-to-tail and head-to-tail dimers with the same beta-helical monomer structure as the head-to-head dimer only form four intermonomer hydrogen bonds and are much less stable. Were channels formed from these dimers to be observed, their electrical potential profiles suggest that they should be cation selective, probably conduct less than the head-to-head dimer, have a central cation binding site, bind cations preferentially if crystallizable, and in the case of the head-to-tail dimer, rectify. Like the antiparallel double stranded helical dimer (a possible minor conducting pathway) the parallel double stranded helical dimer has 28 interstrand hydrogen bonds, but its hydrogen bond network is quite distorted and it is much less stable. If it formed, its electrical potential profile suggests that it would be cation selective, bind anions preferentially if crystallizable, rectify, and at high enough voltages, might exhibit a conductance greater than that of the antiparallel form.  相似文献   

11.
E Alvarez  N Gironès    R J Davis 《The EMBO journal》1989,8(8):2231-2240
The human transferrin receptor is expressed as a disulfide-linked dimer at the cell surface. The sites of intermolecular disulfide bonds are Cys-89 and Cys-98. We have examined the functional significance of the covalent dimeric structure of the transferrin receptor by substitution of Cys-89 and Cys-98 with serine residues. Wild-type and mutated transferrin receptors were expressed in Chinese hamster ovary cells (clone TF-) that lack detectable endogenous transferrin receptors. The rates of receptor endocytosis and recycling were measured and the accumulation of iron by cells incubated with [59Fe]diferric transferrin was investigated. No significant differences between these rates were observed when cells expressing wild-type and mutated receptors were compared. The structure of the mutant receptor lacking intermolecular disulfide bonds was investigated. The presence of a population of mutant receptors with a non-covalent dimeric structure was indicated by cross-linking studies using diferric [125I]transferrin and the bifunctional reagent disuccinimidyl suberimidate. However, sucrose density gradient sedimentation analysis of Triton X-100 solubilized transferrin receptors demonstrated that the mutant receptor existed as a monomer in the absence of diferric transferrin and as an apparent dimer in the presence of this receptor ligand. We conclude that the covalent dimeric structure of the transferrin receptor is not required for the expression of the dimeric state and functional activity of the receptor.  相似文献   

12.
Curcuma longa rhizome lectin, of non-seed origin having antifungal, antibacterial and α-glucosidase inhibitory activities, forms a homodimer with high thermal stability as well as acid tolerance. Size exclusion chromatography and dynamic light scattering show it to be a dimer at pH 7, but it converts to a monomer near pH 2. Circular dichroism spectra and fluorescence emission maxima are virtually indistinguishable from pH 7 to 2, indicating secondary and tertiary structures remain the same in dimer and monomer within experimental error. The tryptophan environment as probed by acrylamide quenching data yielded very similar data at pH 2 and pH 7, implying very similar folding for monomer and dimer. Differential scanning calorimetry shows a transition at 350.3 K for dimer and at 327.0 K for monomer. Thermal unfolding and chemical unfolding induced by guanidinium chloride for dimer are both reversible and can be described by two-state models. The temperatures and the denaturant concentrations at which one-half of the protein molecules are unfolded, are protein concentration-dependent for dimer but protein concentration-independent for monomer. The free energy of unfolding at 298 K was found to be 5.23 Kcal mol−1 and 14.90 Kcal mol−1 for the monomer and dimer respectively. The value of change in excess heat capacity upon protein denaturation (ΔCp) is 3.42 Kcal mol−1 K−1 for dimer. The small ΔCp for unfolding of CLA reflects a buried hydrophobic core in the folded dimeric protein. These unfolding experiments, temperature dependent circular dichroism and dynamic light scattering for the dimer at pH 7 indicate its higher stability than for the monomer at pH 2. This difference in stability of dimeric and monomeric forms highlights the contribution of inter-subunit interactions in the former.  相似文献   

13.
The number of artificial protein supramolecules has been increasing; however, control of protein oligomer formation remains challenging. Cytochrome c′ from Allochromatium vinosum (AVCP) is a homodimeric protein in its native form, where its protomer exhibits a four‐helix bundle structure containing a covalently bound five‐coordinate heme as a gas binding site. AVCP exhibits a unique reversible dimer–monomer transition according to the absence and presence of CO. Herein, domain‐swapped dimeric AVCP was constructed and utilized to form a tetramer and high‐order oligomers. The X‐ray crystal structure of oxidized tetrameric AVCP consisted of two monomer subunits and one domain‐swapped dimer subunit, which exchanged the region containing helices αA and αB between protomers. The active site structures of the domain‐swapped dimer subunit and monomer subunits in the tetramer were similar to those of the monomer subunits in the native dimer. The subunit–subunit interactions at the interfaces of the domain‐swapped dimer and monomer subunits in the tetramer were also similar to the subunit–subunit interaction in the native dimer. Reduced tetrameric AVCP dissociated to a domain‐swapped dimer and two monomers upon CO binding. Without monomers, the domain‐swapped dimers formed tetramers, hexamers, and higher‐order oligomers in the absence of CO, whereas the oligomers dissociated to domain‐swapped dimers in the presence of CO, demonstrating that the domain‐swapped dimer maintains the CO‐induced subunit dissociation behavior of native ACVP. These results suggest that protein oligomer formation may be controlled by utilizing domain swapping for a dimer–monomer transition protein.  相似文献   

14.
Muskett FW  May FE  Westley BR  Feeney J 《Biochemistry》2003,42(51):15139-15147
The trefoil protein TFF3 forms a homodimer (via a disulfide linkage) that is thought to have increased biological activity over the monomer. The solution structure of the TFF3 dimer has been determined by NMR and compared with the structure of the TFF3 monomer and with other trefoil dimer structures (TFF1 and TFF2). The most significant structural differences between the trefoil domain in the monomer and dimer TFF3 are in the orientations of the N-terminal 3(10)-helix (residues 10-12) and in the presence in the dimer of an additional 3(10)-helix (residues 53-55) outside of the core region. The TFF3 dimer forms a more compact structure as compared with the TFF1 dimer where the two trefoil domains are connected by a flexible region with the monomer units being at variable distances from each other and in many different orientations. Although TFF2 is also a compact structure, the dispositions of its monomer units are very different from those of TFF3. The structural differences between the dimers result in the two putative receptor/ligand binding sites that remain solvent exposed in the dimeric structures having very different dispositions in the different dimers. Such differences have significant implications for the mechanism of action and functional specificity for the TFF class of proteins.  相似文献   

15.
suc1 has two native states, a monomer and a domain-swapped dimer, in which one molecule exchanges a beta strand with an identical partner. Thus, monomer and dimer have the same structures but are topologically distinct. Importantly, residues that exchange are part of the folding nucleus of the monomer and therefore forming these interactions in the dimer would be expected to incur a large entropic cost. Here we present the transition state for folding/unfolding of domain-swapped dimeric suc1 and compare it with its monomeric counterpart. The same overall structure is observed in the two transition states but the phi values are consistently higher for the domain-swapped dimer. Thus, a greater entropic penalty for bringing together the key interactions in the dimer is overcome by mobilizing more contacts in the transition state, thereby achieving a greater enthalpic gain.  相似文献   

16.
The Serratia endonuclease is an extracellularly secreted enzyme capable of cleaving both single- and double-stranded forms of DNA and RNA. It is the first member of a large class of related and usually dimeric endonucleases for which a structure is known. Using X-ray crystallography, the structure of monomer of this enzyme was reported by us previously (Miller MD et al., 1994, Nature Struct Biol 1:461-468). We now confirm the dimeric nature of this enzyme through light-scattering experiments and identify the physiologic dimer interface through crystal packing analysis. This dimerization occurs through an isologous twofold interaction localized to the carboxy-terminal subdomain of the enzyme. The dimer is a prolate ellipsoid with dimensions 30 A x 35 A x 90 A. The dimer interface is flat and contains four salt links, several hydrogen bonds, and nonpolar interactions. Buried water is prominent in this interface and it includes an unusual "cubic" water cluster. The position of the two active sites in the dimer suggests that they can act independently in their cleavage of DNA, but have a geometrical advantage in attacking substrate relative to the monomer.  相似文献   

17.
Among proteins utilized as sweeteners, neoculin and miraculin are taste-modifying proteins that exhibit pH-dependent sweetness. Several experiments on neoculin have shown that His11 of neoculin is responsible for pH dependence. We investigated the molecular mechanism of the pH dependence of neoculin by molecular dynamics (MD) calculations. The MD calculations for the dimeric structures of neoculin and His11 mutants showed no significant structural changes for each monomer at neutral and acidic pH levels. The dimeric structure of neoculin dissociated to form isolated monomers under acidic conditions but was maintained at neutral pH. The dimeric structure of the His11Ala mutant, which is sweet at both neutral and acidic pH, showed dissociation at both pH 3 and 7. The His11 residue is located at the interface of the dimer in close proximity to the Asp91 residue of the other monomer. The MD calculations for His11Phe and His11Tyr mutants demonstrated the stability of the dimeric structures at neutral pH and the dissociation of the dimers to isolated monomers. The dissociation of the dimer caused a flexible backbone at the surface that was different from the dimeric interface at the point where the other monomer interacts to form an oligomeric structure. Further MD calculations on the tetrameric structure of neoculin suggested that the flexible backbone contributed to further dissociation of other monomers under acidic conditions. These results suggest that His11 plays a role in the formation of oligomeric structures at pH 7 and that the isolated monomer of neoculin at acidic pH is responsible for sweetness.  相似文献   

18.
Experimental nuclear magnetic resonance results for the Arc Repressor have shown that this dimeric protein dissociates into a molten globule at high pressure. This structural change is accompanied by a modification of the hydrogen-bonding pattern of the intermolecular beta-sheet: it changes its character from intermolecular to intramolecular with respect to the two monomers. Molecular dynamics simulations of the Arc Repressor, as a monomer and a dimer, at elevated pressure have been performed with the aim to study this hypothesis and to identify the major structural and dynamical changes of the protein under such conditions. The monomer appears less stable than the dimer. However, the complete dissociation has not been seen because of the long timescale needed to observe this phenomenon. In fact, the protein structure altered very little when increasing the pressure. It became slightly compressed and the dynamics of the side-chains and the unfolding process slowed down. Increasing both, temperature and pressure, a tendency of conversion of intermolecular into intramolecular hydrogen bonds in the beta-sheet region has been detected, supporting the mentioned hypothesis. Also, the onset of denaturation of the separated chains was observed.  相似文献   

19.
The coding region of copper/zinc-superoxide dismutase (Cu/Zn-SOD) cDNA from sweet potato, Ipomoea batatas (L.) Lam. cv. Tainong 57, was introduced into an expression vector, pET-20b(+). The Cu/Zn-SOD purified by His-tagged technique showed two active forms (dimer and monomer). The amount of proteins of dimer and monomer appeared to be equal, but the activity of dimeric form was seven times higher than that of monomeric form. The enzyme was dissociated into monomer by imidazole buffer above 1.0 M, acidic pH (below 3.0), or SDS (above 1%). The enzyme is quite stable. The enzyme activity is not affected at 85 °C for 20 min, in alkali pH 11.2, or in 0.1 M EDTA and also quite resistant to proteolytic attack. Dimer is more stable than monomer. The thermal inactivation rate constant k dcalculated for the monomer at 85 °C was 0.029 min-1 and the half-life for inactivation was about 28 min. In contrast, there is no significant change of dimer activity after 40 min at 85 °C. The enzyme dimer and monomer retained 83% and 58% of original activity, respectively, after 3 h incubation with trypsin at 37 °C, while those retained 100% and 31% of original activity with chymotrypsin under the same condition. These results suggest subunit interaction might change the enzyme conformation and greatly improve the catalytic activity and stability of the enzyme. It is also possible that the intersubunit contacts stabilize a particular optimal conformation of the protein or the dimeric structure enhances catalytic activity by increasing the electrostatic steering of substrate into the active site.  相似文献   

20.
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   

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