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.
2.
The molecular mechanism whereby the small heat-shock protein (sHsp) chaperones interact with and prevent aggregation of other proteins is not fully understood. We have characterized the sHsp-substrate protein interaction at normal and increased temperatures utilizing a model substrate protein, citrate synthase (CS), widely used in chaperone assays, and a dodecameric plant sHsp, Hsp21, by chemical cross-linking with 3,3'-Dithiobis[sulfosuccinimidylpropionate] (DTSSP) and mass spectrometric peptide mapping. In the absence of CS, the cross-linker captured Hsp21 in dodecameric form, even at increased temperature (47 degrees C). In the presence of equimolar amounts of CS, no Hsp21 dodecamer was captured, indicating a substrate-induced Hsp21 dodecamer dissociation by equimolar amounts of CS. Cross-linked Hsp21-Hsp21 dipeptides indicated an exposure of the Hsp21 C-terminal tails and substrate-binding sites normally covered by the C terminus. Cross-linked Hsp21-CS dipeptides mapped to several sites on the surface of the CS dimer, indicating that there are numerous weak and short-lived interactions between Hsp21 and CS, even at normal temperatures. The N-terminal arms especially interacted with a motif in the CS dimer, which is absent in thermostable forms of CS. The cross-linking data suggest that the presence of substrate rather than temperature influences the conformation of Hsp21.  相似文献   

3.
In this study, we report the purification, initial structural characterization, and functional analysis of the molecular chaperone ClpB from the gram-positive, halophilic lactic acid bacterium Tetragenococcus halophilus. A recombinant T. halophilus ClpB (ClpB(Tha)) was overexpressed in Escherichia coli and purified by affinity chromatography, hydroxyapatite chromatography, and gel filtration chromatography. As demonstrated by gel filtration chromatography, chemical cross-linking with glutaraldehyde, and electron microscopy, ClpB(Tha) forms a homohexameric single-ring structure in the presence of ATP under nonstress conditions. However, under stress conditions, such as high-temperature (>45 degrees C) and high-salt concentrations (>1 M KCl), it dissociated into dimers and monomers, regardless of the presence of ATP. The hexameric ClpB(Tha) reactivated heat-aggregated proteins dependent upon the DnaK system from T. halophilus (KJE(Tha)) and ATP. Interestingly, the mixture of dimer and monomer ClpB(Tha), which was formed under stress conditions, protected substrate proteins from thermal inactivation and aggregation in a manner similar to those of general molecular chaperones. From these results, we hypothesize that ClpB(Tha) forms dimers and monomers to function as a holding chaperone under stress conditions, whereas it forms a hexamer ring to function as a disaggregating chaperone in cooperation with KJE(Tha) and ATP under poststress conditions.  相似文献   

4.
Three forms (termed I, II, and III) of ribonuclease H (RNase H) [EC 3.1.4.34] activity are present in rat liver cytosol. These enzymes degrade RNA specifically in RNA-DNA hybrid structures. They were eluted at 0 M, 0.25 M, and 0.5 M KCl in phosphocellulose chromatography, and were further purified by using blue Sepharose. They are further distinguished from one another by their ionic requirements, optimal pH, molecular weights, sedimentation coefficients, and sensitivity to the -SH reagent, p-chloromercuribenzoate, although I and III have similar characteristics. They liberate a mixture of oligonucleotides with 5'-phosphate and 3'-hydroxyl termini.  相似文献   

5.
6.
Using phosphocellulose followed by single-stranded DNA-cellulose chromatography for purification of UvrC proteins from overproducing cells, we found that UvrC elutes at two peaks: 0.4 m KCl (UvrCI) and 0.6 m KCl (UvrCII). Both forms of UvrC have a major peptide band (>95%) of the same molecular weight and identical N-terminal amino acid sequences, which are consistent with the initiation codon being at the unusual GTG site. Both forms of UvrC are active in incising UV-irradiated, supercoiled phiX-174 replicative form I DNA in the presence of UvrA and UvrB proteins; however, the specific activity of UvrCII is one-fourth that of UvrCI. The molecular weight of UvrCII is four times that of UvrCI on the basis of results of size exclusion chromatography and glutaraldehyde cross-linking reactions, indicating that UvrCII is a tetramer of UvrCI. Functionally, these two forms of UvrC proteins can be distinguished under reaction conditions in which the protein/nucleotide molar ratio is >0.06 by using UV-irradiated, (32)P-labeled DNA fragments as substrates; under these conditions UvrCII is inactive in incision, but UvrCI remains active. The activity of UvrCII in incising UV-irradiated, (32)P- labeled DNA fragments can be restored by adding unirradiated competitive DNA, and the increased level of incision corresponds to a decreased level of UvrCII binding to the substrate DNA. The sites of incision at the 5' and 3' sides of a UV-induced pyrimidine dimer are the same for UvrCI and UvrCII. Nitrocellulose filter binding and gel retardation assays show that UvrCII binds to both UV-irradiated and unirradiated double-stranded DNA with the same affinity (K(a), 9 x 10(8)/m) and in a concentration-dependent manner, whereas UvrCI does not. These two forms of UvrC were also produced by the endogenous uvrC operon. We propose that UvrCII-DNA binding may interfere with Uvr(A)(2)B-DNA damage complex formation. However, because of its low copy number and low binding affinity to DNA, UvrCII may not interfere with Uvr(A)(2)B-DNA damage complex formation in vivo, but instead through double-stranded DNA binding UvrCII may become concentrated at genomic areas and therefore may facilitate nucleotide excision repair.  相似文献   

7.
8.
Purified intact Sindbis virus nucleocapsids were treated at different pH values or with various concentrations of divalent cations, cation chelators, salt, or formamide. The resulting structures were examined by velocity sedimentation, electron microscopy, and protein-protein cross-linking. Changes in each of the test conditions led to alterations in the sedimentation profile of treated nucleocapsids. Appropriate concentrations of formamide or divalent cations generated beaded strandlike structures similar in morphology to those generated from adenovirus cores and nucleosomes. The capsid protein and RNA remained associated with each other at NaCl concentrations less than or equal to 1 M or after treatment of the structures with alkaline pH up to and including pH 10.7. Protein and RNA were dissociated by salt concentrations of greater than 1 M, suggesting that the arginine-rich, amino-terminal portion of the capsid protein is responsible for binding the RNA. Protein-protein cross-linking also indicated that the capsid proteins remained associated in small aggregates under some of the conditions that caused dissociation of the nucleocapsid and suggested the presence of more than one type of protein-protein interaction in the nucleocapsids. Collectively, these data suggest that, like histones and adenovirus core proteins, the Sindbis virus capsid protein serves to package segments of the genome into nucleoprotein beads which are capable of interacting with each other to form the nucleocapsid structure.  相似文献   

9.
The recombinant bovine factor B, obtained by a newly developed bacterial expression system, was found to exhibit features characteristic of a reversible self-associating system. Using size-sieving chromatography, distribution of the factor B species ranged from a monomer to a trimer, but not oligomers of higher molecular weights. At high protein concentrations, factor B migrated as a single band in a native gel. Cross-linking with the amino-reactive cross-linking reagent bis (sulfosuccinimidyl) suberate (BS), at a low cross-linker to protein ratio yielded cross-linked products identified as factor B dimer and trimer. The cross-linking pattern was shown to be a function of the protein and cross-linker concentrations. The range of sedimentation coefficients in a sedimentation velocity experiment suggested that the largest particle present in the distribution was more than twice as large as the smallest. The data obtained under multiple conditions in the sedimentation equilibrium experiments are best fit to a model describing a reversible self-association of a monomer-trimer of factor B species, with a dissociation constant Kd(1,3)=2.48x10(-10) M(2).  相似文献   

10.
Escherichia coli ribonuclease III, purified to homogeneity from an overexpressing bacterial strain, exhibits a high catalytic efficiency and thermostable processing activity in vitro. The RNase III-catalyzed cleavage of a 47 nucleotide substrate (R1.1 RNA), based on the bacteriophage T7 R1.1 processing signal, follows substrate saturation kinetics, with a Km of 0.26 microM, and kcat of 7.7 min.-1 (37 degrees C, in buffer containing 250 mM potassium glutamate and 10 mM MgCl2). Mn2+ and Co2+ can support the enzymatic cleavage of the R1.1 RNA canonical site, and both metal ions exhibit concentration dependences similar to that of Mg2+. Mn2+ and Co2+ in addition promote enzymatic cleavage of a secondary site in R1.1 RNA, which is proposed to result from the altered hydrolytic activity of the metalloenzyme (RNase III 'star' activity), exhibiting a broadened cleavage specificity. Neither Ca2+ nor Zn2+ support RNase III processing, and Zn2+ moreover inhibits the Mg(2+)-dependent enzymatic reaction without blocking substrate binding. RNase III does not require monovalent salt for processing activity; however, the in vitro reactivity pattern is influenced by the monovalent salt concentration, as well as type of anion. First, R1.1 RNA secondary site cleavage increases as the salt concentration is lowered, perhaps reflecting enhanced enzyme binding to substrate. Second, the substitution of glutamate anion for chloride anion extends the salt concentration range within which efficient processing occurs. Third, fluoride anion inhibits RNase III-catalyzed cleavage, by a mechanism which does not involve inhibition of substrate binding.  相似文献   

11.
The native form of phospholamban in cardiac sarcoplasmic reticulum membranes was investigated using photosensitive heterobifunctional cross-linkers, both cleavable and noncleavable, and common protein modifiers. The photosensitive heterobifunctional cleavable cross-linker ethyl 4-azidophenyl-1, 4-dithiobutyrimidate was used in native SR vesicles and it cross-linked phospholamban into an apparent phospholamban-phospholamban dimer and into an approximately 110,000-Da species. The phospholamban dimer migrated at approximately 12,000 Da on sodium dodecyl sulfate-polyacrylamide gels, and upon cleavage of the cross-linker before electrophoresis the dimer disappeared. The approximately 110,000-Da cross-linked species was not affected by boiling in sodium dodecyl sulfate prior to electrophoresis. This cross-linked form of phospholamban migrated approximately 5500 Da above the Ca2(+)-ATPase, which was visualized using fluorescein 5'-isothiocynate, a fluorescent marker that binds specifically to the Ca2(+)-ATPase. p-Azidophenacyl bromide, iodoacetic acid, and N-ethylmaleimide, all of which react with sulfhydryl groups, were also employed to further characterize phospholamban in native sarcoplasmic reticulum membranes. Cross-linking with p-azidophenacyl bromide resulted in only monomeric and dimeric forms of phospholamban as observed on sodium dodecyl sulfate-polyacrylamide gels. Iodoacetic acid and N-ethylmalemide were found to be effective in disrupting the pentameric form of phospholamban only when reacted with sodium dodecyl sulfate solubilized sarcoplasmic reticulum. In view of these findings, the amino acid sequence of phospholamban was examined for possible protein-protein interaction sites. Analysis by hydropathic profiling and secondary structure prediction suggests that the region of amino acids 1-14 may form an amphipathic alpha helix and the hydrophobic surface on one of its sites could interact with the reciprocal hydrophobic surface of another protein, such as the Ca2(+)-ATPase.  相似文献   

12.
BACKGROUND: Aquifex aeolicus Ribonuclease III (Aa-RNase III) belongs to the family of Mg(2+)-dependent endonucleases that show specificity for double-stranded RNA (dsRNA). RNase III is conserved in all known bacteria and eukaryotes and has 1-2 copies of a 9-residue consensus sequence, known as the RNase III signature motif. The bacterial RNase III proteins are the simplest, consisting of two domains: an N-terminal endonuclease domain, followed by a double-stranded RNA binding domain (dsRBD). The three-dimensional structure of the dsRBD in Escherichia coli RNase III has been elucidated; no structural information is available for the endonuclease domain of any RNase III. RESULTS: We present the crystal structures of the Aa-RNase III endonuclease domain in its ligand-free form and in complex with Mn(2+). The structures reveal a novel protein fold and suggest a mechanism for dsRNA cleavage. On the basis of structural, genetic, and biological data, we have constructed a hypothetical model of Aa-RNase III in complex with dsRNA and Mg(2+) ion, which provides the first glimpse of RNase III in action. CONCLUSIONS: The functional Aa-RNase III dimer is formed via mainly hydrophobic interactions, including a "ball-and-socket" junction that ensures accurate alignment of the two monomers. The fold of the polypeptide chain and its dimerization create a valley with two compound active centers at each end of the valley. The valley can accommodate a dsRNA substrate. Mn(2+) binding has significant impact on crystal packing, intermolecular interactions, thermal stability, and the formation of two RNA-cutting sites within each compound active center.  相似文献   

13.
Lyophilization of bovine ribonuclease A (RNase A; Sigma, type XII-A) from 40% acetic acid solutions leads to the formation of approximately 14 aggregated species that can be separated by ion-exchange chromatography. Several aggregates were identified, including two variously deamidated dimeric subspecies, two distinct trimeric and two distinct tetrameric RNase A conformers, besides the two forms of dimer characterized previously [Gotte, G. & Libonati, M. (1998) Two different forms of aggregated dimers of ribonuclease A. Biochim. Biophys. Acta 1386, 106-112]. We also have possible evidence for the existence of two forms of pentameric RNase A. The two forms of trimers and tetramers are characterized by: (a) slightly different gel filtration patterns; (b) different retention times in ion-exchange chromatography; and (c) different mobilities in cathodic gel electrophoresis under nondenaturing conditions. Therefore, they appear to have distinct structural organizations responsible for a different availability of their positively charged amino acid residues. All RNase A oligomers, in particular the two distinct trimeric and tetrameric conformers, degrade poly(A).poly(U), viral double-stranded RNA and polyadenylate with a catalytic efficiency that is in general higher for the more basic species. On the contrary, the activity of the RNase A oligomers, from dimer to pentamer, on yeast RNA and poly(C) (Kunitz assay) is lower than that of monomeric RNase A.  相似文献   

14.
A monoclonal antibody (BF4) has been used to characterize and purify the heat-shock protein of Mr approximately 90,000 (hsp 90) present in the chick oviduct. In low salt cytosol, the sedimentation coefficient of hsp 90 is approximately 6.8 S, the Stokes radius approximately 7.1 nm, and the calculated Mr approximately 204,000, thus suggesting a dimeric structure. In 0.4 M KCl cytosol, only slightly smaller values were determined (approximately 6.5 S, approximately 6.8 nm, and approximately 187,000). Following purification by ion exchange and immunoaffinity chromatography, hsp 90 migrated as a single silver-stained band at Mr approximately 90,000 in sodium dodecyl sulfate-polyacrylamide gel electrophoresis, while the sedimentation coefficient 6.2 S, the Stokes radius approximately 6.8 nm, and the Mr approximately 178,000 confirmed the dimeric structure. However, in both antigen or antibody excess conditions, only one molecule of monoclonal antibody could be bound to the hsp 90 dimer. Whether steric hindrance in a homodimer or the presence of two different 90-kDa proteins in a heterodimer explains this result cannot yet be decided. The dimer is not dissociated by high salt (1 M KCl) or the chaotropic agent (0.5 M NaSCN), but is disrupted by 4 M urea, suggesting a stabilization of the structure by hydrogen bonds. The molybdate-stabilized progesterone receptor hetero-oligomer form of approximately 8 S sedimentation coefficient was purified, and its hsp 90 component was then released by salt treatment. It was found to sediment at approximately 5.8 S and have a Stokes radius approximately 7.1 nm, giving Mr approximately 174,000. This observation is consistent with a previous report suggesting from specific activity determination, scanning of polyacrylamide gels, and cross-linking experiments that each purified nontransformed progesterone receptor molecule includes one progesterone binding unit per two 90-kDa protein molecules (Renoir, J. M., Buchou, T., Mester, J., Radanyi, C., and Baulieu, E. E. (1984) Biochemistry 23, 6016-6023). This work brings direct evidence that both free hsp 90 and the non-hormone binding hsp 90 component released from the nontransformed steroid receptor in the cytosol are in a dimeric form.  相似文献   

15.
When concentrated in mildly acidic solutions, bovine pancreatic ribonuclease (RNase A) forms long-lived oligomers including two types of dimer, two types of trimer, and higher oligomers. In previous crystallographic work, we found that the major dimeric component forms by a swapping of the C-terminal beta-strands between the monomers, and that the minor dimeric component forms by swapping the N-terminal alpha-helices of the monomers. On the basis of these structures, we proposed that a linear RNase A trimer can form from a central molecule that simultaneously swaps its N-terminal helix with a second RNase A molecule and its C-terminal strand with a third molecule. Studies by dissociation are consistent with this model for the major trimeric component: the major trimer dissociates into both the major and the minor dimers, as well as monomers. In contrast, the minor trimer component dissociates into the monomer and the major dimer. This suggests that the minor trimer is cyclic, formed from three monomers that swap their C-terminal beta-strands into identical molecules. These conclusions are supported by cross-linking of lysyl residues, showing that the major trimer swaps its N-terminal helix, and the minor trimer does not. We verified by X-ray crystallography the proposed cyclic structure for the minor trimer, with swapping of the C-terminal beta-strands. This study thus expands the variety of domain-swapped oligomers by revealing the first example of a protein that can form both a linear and a cyclic domain-swapped oligomer. These structures permit interpretation of the enzymatic activities of the RNase A oligomers on double-stranded RNA.  相似文献   

16.
Park C  Raines RT 《Biochemistry》2003,42(12):3509-3518
The value of k(cat)/K(M) for catalysis of RNA cleavage by ribonuclease (RNase) A can exceed 10(9) M(-1) s(-1) in a solution of low salt concentration. This value approaches that expected for the diffusional encounter of the enzyme and its substrate. To reveal the physicochemical constraints upon catalysis by RNase A, the effects of salt concentration, pH, solvent isotope, and solvent viscosity on catalysis were determined with synthetic substrates that bind to all of the enzymic subsites and thereby enable a meaningful analysis. The pK(a) values determined from pH-k(cat)/K(M) profiles at 0.010, 0.20, and 1.0 M NaCl are inconsistent with the known macroscopic pK(a) values of RNase A. This incongruity indicates that catalysis of RNA cleavage by RNase A is limited by the rate of substrate association, even at 1.0 M NaCl. The effect of solvent isotope and solvent viscosity on catalysis support this conclusion. The data are consistent with a mechanism in which RNase A associates with RNA in an intermediate complex, which is stabilized by Coulombic interactions, prior to the formation of a Michaelis complex. Thus, RNase A has evolved to become an enzyme limited by physics rather than chemistry, a requisite attribute of a perfect catalyst.  相似文献   

17.
Tang J  Luo M  Niu S  Zhou H  Cai X  Zhang W  Hu Y  Yin Y  Huang A  Wang D 《The protein journal》2010,29(8):583-590
RNase E functions as the rate-limiting enzyme in the global mRNA metabolism as well as in the maturation of functional RNAs. The endoribonuclease, binding to the PNPase trimer, the RhlB monomer, and the enolase dimer, assembles into an RNA degradosome necessary for effective RNA metabolism. The RNase E processing is found to be negatively regulated by the protein modulator RraA which appears to work by interacting with the non-catalytic region of the endoribonuclease and significantly reduce the interaction between RNase E and PNPase, RhlB and enolase of the RNA degradosome. Here we report the crystal structure of RraA from P. aeruginosa to a resolution of 2.0 ?. The overall architecture of RraA is very similar to other known RraAs, which are highly structurally conserved. Gel filtration and dynamic light scattering experiments suggest that the protein regulator is arranged as a hexamer, consistent with the crystal packing of "a dimer of trimer" arrangement. Structure and sequence conservation analysis suggests that the hexamer RraA contains six putative charged protein-protein interaction sites which may serve as binding sites for RNase E.  相似文献   

18.
Yeast Rnt1 is a member of the double-stranded RNA (dsRNA)-specific RNase III family identified by conserved dsRNA binding (dsRBD) and nuclease domains. Comparative sequence analyses have revealed an additional N-terminal domain unique to the eukaryotic homologues of RNase III. The deletion of this domain from Rnt1 slowed growth and led to mild accumulation of unprocessed 25S pre-rRNA. In vitro, deletion of the N-terminal domain reduced the rate of RNA cleavage under physiological salt concentration. Size exclusion chromatography and cross-linking assays indicated that the N-terminal domain and the dsRBD self-interact to stabilize the Rnt1 homodimer. In addition, an interaction between the N-terminal domain and the dsRBD was identified by a two-hybrid assay. The results suggest that the eukaryotic N-terminal domain of Rnt1 ensures efficient dsRNA cleavage by mediating the assembly of optimum Rnt1-RNA ribonucleoprotein complex.  相似文献   

19.
A facile method for the formation of zero-length covalent cross-links between protein molecules in the lyophilized state without the use of chemical reagents has been developed. The cross-linking process is performed by simply sealing lyophilized protein under vacuum in a glass vessel and heating at 85 degrees C for 24 h. Under these conditions, approximately one-third of the total protein present becomes cross-linked, and dimer is the major product. Chemical and mass spectroscopic evidence obtained shows that zero-length cross-links are formed as a result of the condensation of interacting ammonium and carboxylate groups to form amide bonds between adjacent molecules. For the protein examined in the most detail, RNase A, the cross-linked dimer has only one amide cross-link and retains the enzymatic activity of the monomer. The in vacuo cross-linking procedure appears to be general in its applicability because five different proteins tested gave substantial cross-linking, and co-lyophilization of lysozyme and RNase A also gave a heterogeneous covalently cross-linked dimer.  相似文献   

20.
Analysis of native disulfide-bonded protein oligomers in paramyxoviruses showed that some viral proteins are consistently present as covalent complexes. In isolated Sendai virus the hemagglutinating protein HN is present in homodimeric and homotetrameric forms, and the minor nucleocapsid protein P exists partly as a monomer and partly as a disulfide-linked homotrimer. Similar disulfide-linked complexes were observed in Newcastle disease virus (strain HP-16), in which HN exists as a homodimer and some of the major nucleocapsid protein NP exists as a homotrimer. Noncovalent intermolecular interactions between proteins were studied with the reversible chemical cross-linkers dimethyl-3,3'-dithiobispropionimidate and methyl 3-[(p-azidophenyl)dithio]propionimidate, which contain disulfide bridges and a 1.1-nm separation between their functional groups. The same results were achieved with both reagents. The conditions of preparation, isolation, and storage of the viruses affected the protein-protein interactions observed upon cross-linking. Homooligomers of the glycoprotein F, the matrix protein M, and the major nucleocapsid protein NP were produced in both Sendai and Newcastle disease viruses after mild cross-linking of all viral preparations examined, but NP-M heterodimer formation in both viruses was most prevalent in early harvest preparations that were cross-linked soon after isolation. The ability of NP and M to form a heterodimer upon cross-linking indicates that the matrix protein layer lies in close proximity (within 1.1 nm) to the nucleocapsid in the newly formed virion. Some noncovalent intermolecular protein interactions in Sendai and Newcastle disease viruses, i.e., those leading to the formation of F, NP, and M homooliogmers upon cross-linking, are more stable to virus storage than others, i.e., those leading to the formation of an NP-M heterodimer upon cross-linking. The storage-induced loss of the ability of NP and M to form a heterodimer is not accompanied by any apparent loss of infectivity. This indicates that some spacial relationships which form during virus assembly can alter after particle formation and are not essential for the ensuing stages of the infectious process.  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