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 15 毫秒
1.
Nitric oxide synthases (NOSs) share two invariant tryptophan residues within a conserved helical lariat that is part of the pterin-binding site and dimer interface. We mutated Staphylococcus aureus NOS Trp-314 (to alanine, phenylalanine, tyrosine and histidine) and Trp-316 (to alanine, phenylalanine and tyrosine) and characterized the effects of mutation on heme environment, quaternary structure, enzymatic activity, and substrate affinity. With arginine present, all saNOS variants bound heme with native thiolate ligation, formed high spin ferric complexes and were dimeric. All variants catalyze the peroxide-dependent oxidation of N-hydroxy-l-arginine, at rates from 10% to 55% of wild type activity. Arginine-free proteins are dimeric with the exception of W314A. Arginine affinity for all variants decreases with increasing temperature between 15 and 42 °C but is precipitous for position-314 variants. Previous structural and biophysical characterization of NOS oxygenase domains demonstrated that the protein can exist in either a tight or loose conformation, with the former corresponding to the active state of the protein. In the position-314 variants it is likely that the loose conformation is favoured, owing to the loss of a hydrogen bond between the indole side chain and the polypeptide backbone of the helical lariat.  相似文献   

2.
3.
Nitric oxide synthases (NOSs), which catalyze the formation of the ubiquitous biological messenger molecule nitric oxide, represent unique cytochrome P-450s, containing reductase and mono-oxygenase domains within one polypeptide and requiring tetrahydrobiopterin as cofactor. To investigate whether tetrahydrobiopterin functions as an allosteric effector of NOS, we have analyzed the effect of the pteridine on the conformation of neuronal NOS purified from porcine brain by means of circular dichroism, velocity sedimentation, dynamic light scattering and SDS-polyacrylamide gel electrophoresis. We report for the first time the secondary structure of NOS, showing that the neuronal isozyme contains 30% alpha-helix, 14% antiparallel beta-sheet, 7% parallel beta-sheet, 19% turns and 31% other structures. The secondary structure of neuronal NOS was neither modulated nor stabilized by tetrahydrobiopterin, and the pteridine did not affect the quaternary structure of the protein, which appears to be an elongated homodimer with an axial ratio of approximately 20/1 under native conditions. Low temperature SDS-polyacrylamide gel electrophoresis revealed that tetrahydrobiopterin and L-arginine synergistically convert neuronal NOS into an exceptionally stable, non-covalently linked homodimer surviving in 2% SDS and 5% 2-mercaptoethanol. Ligand-induced formation of an SDS-resistant dimer is unprecedented and suggests a novel role for tetrahydrobiopterin and L-arginine in the allosteric regulation of protein subunit interactions.  相似文献   

4.
Generation of RNA dimeric form of the human immunodeficiency virus type 1 (HIV-1) genome is crucial for viral replication. The dimerization initiation site (DIS) has been identified as a primary sequence that can form a stem-loop structure with a self-complementary sequence in the loop and a bulge in the stem. It has been reported that HIV-1 RNA fragments containing the DIS form two types of dimers, loose dimers and tight dimers. The loose dimers are spontaneously generated at the physiological temperature and converted into tight dimers by the addition of nucleocapsid protein NCp7. To know the biochemical process in this two-step dimerization reaction, we chemically synthesized a 39-mer RNA covering the entire DIS sequence and also a 23-mer RNA covering the self-complementary loop and its flanking stem within the DIS. Electrophoretic dimerization assays demonstrated that the 39-mer RNA reproduced the two-step dimerization process, whereas the 23-mer RNA immediately formed the tight dimer. Furthermore, deletion of the bulge from the 39-mer RNA prevented the NCp7-assisted tight-dimer formation. Therefore, the whole DIS sequence is necessary and sufficient for the two-step dimerization. Our data suggested that the bulge region regulates the stability of the stem and guides the DIS to the two-step dimerization process.  相似文献   

5.
Homodimer formation activates all nitric-oxide synthases (NOSs). It involves the interaction between two oxygenase domains (NOSoxy) that each bind heme and (6R)-tetrahydrobiopterin (H4B) and catalyze NO synthesis from L-Arg. Here we compared three NOSoxy isozymes regarding dimer strength, interface composition, and the ability of L-Arg and H4B to stabilize the dimer, promote its formation, and protect it from proteolysis. Urea dissociation studies indicated that the relative dimer strengths were NOSIIIoxy > NOSIoxy > NOSIIoxy (endothelial NOSoxy (eNOSoxy) > neuronal NOSOXY (nNOSoxy) > inducible NOSoxy (iNOSoxy)). Dimer strengths of the full-length NOSs had the same rank order as judged by their urea-induced loss of NO synthesis activity. NOSoxy dimers containing L-Arg plus H4B exhibited the greatest resistance to urea-induced dissociation followed by those containing either molecule and then by those containing neither. Analysis of crystallographic structures of eNOSoxy and iNOSoxy dimers showed more intersubunit contacts and buried surface area in the dimer interface of eNOSoxy than iNOSoxy, thus revealing a potential basis for their different stabilities. L-Arg plus H4B promoted dimerization of urea-generated iNOSoxy and nNOSoxy monomers, which otherwise was minimal in their absence, and also protected both dimers against trypsin proteolysis. In these respects, L-Arg alone was more effective than H4B alone for nNOSoxy, whereas for iNOSoxy the converse was true. The eNOSoxy dimer was insensitive to proteolysis under all conditions. Our results indicate that the three NOS isozymes, despite their general structural similarity, differ markedly in their strengths, interfaces, and in how L-Arg and H4B influence their formation and stability. These distinguishing features may provide a basis for selective control and likely help to regulate each NOS in its particular biologic milieu.  相似文献   

6.
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   

7.
Islet-activating protein (IAP), pertussis toxin, is a hexameric protein composed of an A protomer and a B oligomer, the residual pentamer having such a subunit assembly that two different dimers, dimer 1 and dimer 2, are connected with each other by means of the smallest C subunit. Incubation of IAP with formaldehyde and pyridine-borane produced the modified toxin in which most of the free amino groups were dimethylated. The methylated and nonmethylated (native) IAP were disintegrated into their respective constituent components, which were then cross combined to reconstitute hybrid toxins with the original hexameric structure. The binding of the B oligomer to the mammalian cell surface via dimer 2 was, but the binding via dimer 1 was not, seriously impaired by methylation of amino groups in the protein. The binding of the B oligomer allowed the A protomer to enter cells and to catalyze ADP-ribosylation of a membrane Mr 41 000 protein. The diverse biological activities of IAP occurring by this mechanism were mimicked by not only methylated IAP but also all hybrid toxins, indicating that the free amino groups in the protein were not essential for the enzyme activity of the A protomer and that the A protomer was able to enter cells if the B oligomer bound to cells "monovalently" via dimer 1. An additional effect of the B oligomer binding, i.e., the direct stimulation, without the transport of the A protomer, of cells leading to mitosis in lymphocytes in vitro or increases in circulating lymphocytes in vivo, was not mimicked by hybrid toxins containing methylated dimer 2.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Pant K  Bilwes AM  Adak S  Stuehr DJ  Crane BR 《Biochemistry》2002,41(37):11071-11079
Eukaryotic nitric oxide synthases (NOSs) produce nitric oxide to mediate intercellular signaling and protect against pathogens. Recently, proteins homologous to mammalian NOS oxygenase domains have been found in prokaryotes and one from Bacillus subtilis (bsNOS) has been demonstrated to produce nitric oxide [Adak, S., Aulak, K. S., and Stuehr, D. J. (2002) J. Biol. Chem. 277, 16167-16171]. We present structures of bsNOS complexed with the active cofactor tetrahydrofolate and the substrate L-arginine (L-Arg) or the intermediate N(omega)-hydroxy-L-arginine (NHA) to 1.9 or 2.2 A resolution, respectively. The bsNOS structure is similar to those of the mammalian NOS oxygenase domains (mNOS(ox)) except for the absence of an N-terminal beta-hairpin hook and zinc-binding region that interact with pterin and stabilize the mNOS(ox) dimer. Changes in patterns of residue conservation between bacterial and mammalian NOSs correlate to different binding modes for pterin side chains. Residue conservation on a surface patch surrounding an exposed heme edge indicates a likely interaction site for reductase proteins in all NOSs. The heme pockets of bsNOS and mNOS(ox) recognize L-Arg and NHA similarly, although a change from Val to Ile beside the substrate guanidinium may explain the 10-20-fold slower dissociation of product NO from the bacterial enzyme. Overall, these structures suggest that bsNOS functions naturally to produce nitrogen oxides from L-Arg and NHA in a pterin-dependent manner, but that the regulation and purpose of NO production by NOS may be quite different in B. subtilis than in mammals.  相似文献   

9.
The oligomeric "sliding clamp" processivity factors, such as PCNA, are thought to rely on a loose, topological association with DNA to slide freely along dsDNA. Unlike PCNA, the processivity subunit of the herpes simplex virus DNA polymerase, UL42, is a monomer and has an intrinsic affinity for dsDNA that is remarkably high for a sequence-independent DNA binding protein. Using a DNase footprinting assay, we demonstrate that UL42 translocates with the catalytic subunit of the polymerase during chain elongation. In addition, footprinting and electrophoretic mobility shift assays show that, despite its tight DNA binding, UL42 is capable of linear diffusion on DNA at a rate of between 17 and 47 bp/s. Our results thus suggest that, despite profound biochemical differences with the sliding clamps, UL42 can freely slide downstream with the catalytic subunit during DNA replication.  相似文献   

10.
The complete sequence has been determined for the C1 subunit of crustacyanin, an astaxanthin-binding protein from the carapace of the lobster Homarus gammarus (L.). The polypeptide, 181 residues long, is similar (38% identity) to the other main subunit, A2 and to plasma retinol-binding protein. The tertiary structure of the C1 subunit has been modelled on that derived for the A2 subunit from the coordinates of retinol-binding protein. Residues lining the putative binding cavities and at the putative carotenoid binding sites of the two subunits are highly conserved. The carotenoid environments are characterized by a preponderance of aromatic and polar residues and the absence of charged side-chains. A tentative model for the dimer, beta-crustacyanin, formed between the two subunits with their associated carotenoid ligands, is discussed. The model is based on the crystal structure of the dimer of bilin-binding protein, a member of the same superfamily. This structure has enabled us to examine mechanisms for the bathochromic spectral shift of the protein-bound carotenoid and to identify likely contact regions between dimers in octameric alpha-crustacyanin.  相似文献   

11.
Pressure-induced dissociation of tight couple ribosomes   总被引:3,自引:0,他引:3  
M Gross  R Jaenicke 《FEBS letters》1990,267(2):239-241
Ribosomes from Escherichia coli have been shown to undergo subunit dissociation at elevated hydrostatic pressure. This holds for both crude and highly purified ribosomes. No inhibitory effect could be detected by addition of either the S100 supernatant, or tRNA, polyuridylic acid, and spermine. Light scattering experiments at pressures up to 1000 bar reveal different susceptibility of tight couple and loose couple ribosomes toward pressure dissociation. Tight couples are subjected to EF-Tu-catalyzed binding of aminoacyl-tRNA, thus yielding a model system of the elongating ribosome before the peptidyl transfer step. High pressure dissociation of this compound suggests that enzymatic binding converts tight couples into loose couples. A hypothesis referring to conformational changes during the elongation cycle is presented.  相似文献   

12.
Prokaryotic genes related to the oxygenase domain of mammalian nitric oxide synthases (NOSs) have recently been identified. Although they catalyze the same reaction as the eukaryotic NOS oxygenase domain, their biological function(s) are unknown. In order to explore rationally the biochemistry and evolution of the prokaryotic NOS family, we have determined the crystal structure of SANOS, from methicillin-resistant Staphylococcus aureus (MRSA), to 2.4 A. Haem and S-ethylisothiourea (SEITU) are bound at the SANOS active site, while the intersubunit site, occupied by the redox cofactor tetrahydrobiopterin (H(4)B) in mammalian NOSs, has NAD(+) bound in SANOS. In common with all bacterial NOSs, SANOS lacks the N-terminal extension responsible for stable dimerization in mammalian isoforms, but has alternative interactions to promote dimer formation.  相似文献   

13.
Previously, an RNA stem-loop (TR) encompassing 19 nt of the genome of bacteriophage MS2 was shown to act as an allosteric effector of conformational switching in the coat protein during in vitro capsid assembly. TR RNA binding to symmetric coat protein dimers results in conformational changes, principally at the FG-loop connecting the F and G β-strands in each subunit, yielding an asymmetric structure. The FG-loops define the quasi-equivalent conformers of the coat protein subunit (A, B, and C) in the T = 3 capsid. Efficient assembly of this capsid in vitro requires that both symmetrical and asymmetrical forms of the coat protein dimer be present in solution, implying that they closely resemble the quasi-equivalent dimers (A/B and C/C) seen in the final capsid. Experiments show that assembly can be triggered by a number of RNA stem-loops unrelated to TR in sequence and detailed secondary structure, suggesting that there is little sequence specificity to the allosteric effect. Since the stem-loop binding site on the coat protein dimer is distal to the FG-loops the mechanism of this switching effect needs to be investigated. We have analyzed the vibrational modes of both TR-bound and RNA-free coat protein dimers using an all-atom normal-mode analysis. The results suggest that asymmetric contacts between the A-duplex RNA phosphodiester backbone and the EF-loop in one coat protein subunit result in the FG-loop of that subunit becoming more dynamic, whilst the equivalent loop on the other monomer decreases its mobility. The increased dynamic behaviour occurs in the FG-loop of the subunit required to undergo the largest conformational change when adopting the quasi-equivalent B conformation. The free energy barrier on the pathway to form this new structure would consequently be reduced compared to the unbound subunit. Our results also imply that the allosteric effect should be independent of the base sequence of the bound stem-loop, as observed experimentally. As a test of this model, we also examined the vibrational modes of a known assembly mutant, W82R, which cannot assemble beyond dimer. This mutation leads to an increased mobility of the DE-loop rather than the FG-loop after TR binding, consistent with the non-assembling phenotype of this mutant protein.  相似文献   

14.
Band 3 (Mr = 95,000), the anion transport protein of human erythrocyte membranes exists primarily as a dimer in solutions of nonionic detergents such as octaethylene glycol mono-n-dodecyl ether (C12E8). The role of the oligomeric structure of Band 3 in the binding of [14C]4-benzamido-4'-aminostilbene-2,2'-disulfonate (BADS), an inhibitor of anion transport (Ki = 1-2 microM), was studied by characterizing the interaction of BADS with dimers and monomers of Band 3 covalently attached to p-mercuribenzoate-Sepharose 4B. BADS bound to matrix-bound Band 3 dimers with an affinity of approximately 3 microM at a stoichiometry of 1 BADS molecule/Band 3 monomer, in agreement with the BADS binding characteristic of Band 3 in the membrane and in solutions of C12E8. Band 3 dimers could be attached to the matrix via one subunit by limiting the amount of p-chloromercuribenzoate on the Sepharose bead. Matrix-bound monomers were formed by dissociation of the dimers with dodecyl sulfate or guanidine hydrochloride. Complete removal of the denaturants allowed formation of refolded Band 3 monomers since the matrix-bound subunits could not reassociate. These refolded Band 3 monomers were unable to bind BADS. Release of the monomers from the matrix with 2-mercaptoethanol allowed reformation of dimers with recovery of the BADS binding sites. These results suggest that the dimeric structure of Band 3 is required for BADS binding and that the BADS binding sites may be at the interface between the two halves of the Band 3 dimer.  相似文献   

15.
Structure of a novel Bence-Jones protein (Rhe) fragment at 1.6 A resolution   总被引:10,自引:0,他引:10  
The crystal structure of Rhe, a lambda-type Bence-Jones protein fragment, has been solved and refined to a resolution of 1.6 A. A model fragment consisting of the complete variable domain and the first three residues of the constant domain yields a crystallographic residual RF value of 0.149. The protein exists as a dimer both in solution and in the crystals. Although the "immunoglobulin fold" is generally preserved in the structure, there are significant differences in both the monomer conformation and in the mode of association of monomers into dimers, when compared to other known Bence-Jones proteins or Fab fragments. The variations in conformation within monomers are particularly significant as they involve non-hypervariable residues, which previously were believed to be part of a "structurally invariant" framework common to all immunoglobulin variable domains. The novel mode of dimerization is equally important, as it can result in combining site shapes and sizes unobtainable with the conventional mode of dimerization. A comparison of the structure with other variable domain dimers reveals further that the variations within monomers and between domains in the dimer are coupled. Some possible functional implications revealed by this coupling are greater variability, induced fitting of the combining site to better accommodate antigenic determinants, and a mechanism for relaying binding information from one end of the variable domain dimer to the other. In addition to providing the most accurate atomic parameters for an immunoglobulin domain yet obtained, the high resolution and extensive refinement resulted in identification of several tightly bound water molecules in key structural positions. These water molecules may be regarded as integral components of the protein. Other water molecules appear to be required to stabilize the novel conformation.  相似文献   

16.
17.
Mouse salivary androgen-binding protein (ABP) isa family of dimeric proteins that may play a pheromonalrole in Mus musculus. The protein dimer consists of acommon alpha subunit disulfide-bonded to avariable (beta or gamma) subunit. Here wereport N-terminal sequences of the beta and gammasubunits, showing that they are very similar to eachother while being quite different from the alphasubunit.We demonstrate differential androgen binding bythe two dimers. Both bind dihydrotestosterone to aboutthe same extent but the alpha:beta dimer bindssignificantly more testosterone than the alpha:gammadimer. We discuss the possible significance ofthis diversity of androgen binding with respect to thepossibility that androgen binding is related to aputative pheromonal role for the protein.  相似文献   

18.
Fibronectin exists in a compact or extended conformation, depending upon environmental pH and salt concentration. Using recombinant fragments expressed in bacteria and baculovirus, we determined the domains responsible for producing fibronectin's compact conformation. Our velocity and equilibrium sedimentation data show that FN2-14 (a protein containing FN-III domains 2 through 14) forms dimers in low salt. Experiments with smaller fragments indicates that the compact conformation is produced by binding of FN12-14 of one subunit to FN2-3 of the other subunit in the dimer. The binding is weakened at higher salt concentrations, implying an electrostatic interaction. Furthermore, segment FN7-14+A, which contains the alternatively spliced A domain between FN11 and 12, forms dimers, whereas FN7-14 without A does not. Segment FN12-14+A also forms dimers, but the isolated A domain does not. These data imply an association of domain A with FN12-14, and the presence of A may favor an open conformation by competing with FN2-3 for binding to FN12-14.  相似文献   

19.
Griaznova O  Traut RR 《Biochemistry》2000,39(14):4075-4081
Escherichia coli ribosomal protein L10 binds the two L7/L12 dimers and thereby anchors them to the large ribosomal subunit. C-Terminal deletion variants (Delta10, Delta20, and Delta33 amino acids) of ribosomal protein L10 were constructed in order to define the binding sites for the two L7/L12 dimers and then to make and test ribosomal particles that contain only one of the two dimers. None of the deletions interfered with binding of L10 variants to ribosomal core particles. Deletion of 20 or 33 amino acids led to the inability of the proteins to bind both dimers of protein L7/L12. The L10 variant with deletion of 10 amino acids bound one L7/L12 dimer in solution and when reconstituted into ribosomes promoted the binding of only one L7/L12 dimer to the ribosome. The ribosomes that contained a single L7/L12 dimer were homogeneous by gel electrophoresis where they had a mobility between wild-type 50S subunits and cores completely lacking L7/L12. The single-dimer ribosomal particles supported elongation factor G dependent GTP hydrolysis and protein synthesis in vitro with the same activity as that of two-dimer particles. The results suggest that amino acids 145-154 in protein L10 determine the binding site ("internal-site") for one L7/L12 dimer (the one reported here), and residues 155-164 ("C-terminal-site") are involved in the interaction with the second L7/L12 dimer. Homogeneous ribosomal particles containing a single L7/L12 dimer in each of the distinct sites present an ideal system for studying the location, conformation, dynamics, and function of each of the dimers individually.  相似文献   

20.
With the use of special DNA binding sites, but not the natural aral binding site, the dimeric AraC protein can be forced to make sandwich structures in which two DNA molecules are joined by two AraC protein dimers. Apparently one subunit from each dimer contacts each DNA molecule in an extended structure. These sandwich structures form only in the absence of arabinose. This behavior is consistent with the protein's ability to form DNA loops by binding to separated half sites in the absence of arabinose and its preference for binding to adjacent half-sites in the presence of arabinose.  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