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.
The 21st amino acid, selenocysteine (Sec), is incorporated translationally into proteins and is synthesized on its specific tRNA (tRNASec). In Bacteria, the selenocysteine synthase SelA converts Ser-tRNASec, formed by seryl-tRNA synthetase, to Sec-tRNASec. SelA, a member of the fold-type-I pyridoxal 5′-phosphate-dependent enzyme superfamily, has an exceptional homodecameric quaternary structure with a molecular mass of about 500 kDa. Our previously determined crystal structures of Aquifex aeolicus SelA complexed with tRNASec revealed that the ring-shaped decamer is composed of pentamerized SelA dimers, with two SelA dimers arranged to collaboratively interact with one Ser-tRNASec. The SelA catalytic site is close to the dimer–dimer interface, but the significance of the dimer pentamerization in the catalytic site formation remained elusive. In the present study, we examined the quaternary interactions and demonstrated their importance for SelA activity by systematic mutagenesis. Furthermore, we determined the crystal structures of “depentamerized” SelA variants with mutations at the dimer–dimer interface that prevent pentamerization. These dimeric SelA variants formed a distorted and inactivated catalytic site and confirmed that the pentamer interactions are essential for productive catalytic site formation. Intriguingly, the conformation of the non-functional active site of dimeric SelA shares structural features with other fold-type-I pyridoxal 5′-phosphate-dependent enzymes with native dimer or tetramer (dimer-of-dimers) quaternary structures.  相似文献   

2.
Half-minilamins, representing amino- and carboxy-terminal fragments of human lamins A, B1 and B2 with a truncated central rod domain, were investigated for their ability to form distinct head-to-tail-type dimer complexes. This mode of interaction represents an essential step in the longitudinal assembly reaction exhibited by full-length lamin dimers. As determined by analytical ultracentrifugation, the amino-terminal fragments were soluble under low ionic strength conditions sedimenting with distinct profiles and s-values (1.6-1.8 S) indicating the formation of coiled-coil dimers. The smaller carboxy-terminal fragments were, except for lamin B2, largely insoluble under these conditions. However, after equimolar amounts of homotypic amino- and carboxy-terminal lamin fragments had been mixed in 4 M urea, upon subsequent renaturation the carboxy-terminal fragments were completely rescued from precipitation and distinct soluble complexes with higher s-values (2.3-2.7 S) were obtained. From this behavior, we conclude that the amino- and carboxy-terminal coiled-coil dimers interact to form distinct oligomers (i.e. tetramers). Furthermore, a corresponding interaction occurred also between heterotypic pairs of A- and B-type lamin fragments. Hence, A-type lamin dimers may interact with B-type lamin dimers head-to-tail to yield linear polymers. These findings indicate that a lamin dimer principally has the freedom for a “combinatorial” head-to-tail association with all types of lamins, a property that might be of significant importance for the assembly of the nuclear lamina. Furthermore, we suggest that the head-to-tail interaction of the rod end domains represents a principal step in the assembly of cytoplasmic intermediate filament proteins too.  相似文献   

3.
Thioredoxin functions in nearly all organisms as the major thiol-disulfide oxidoreductase within the cytosol. Its prime purpose is to maintain cysteine-containing proteins in the reduced state by converting intramolecular disulfide bonds into dithiols in a disulfide exchange reaction. Thioredoxin has been reported to contribute to a wide variety of physiological functions by interacting with specific sets of substrates in different cell types. To investigate the function of the essential thioredoxin A (TrxA) in the low-GC Gram-positive bacterium Bacillus subtilis, we purified wild-type TrxA and three mutant TrxA proteins that lack either one or both of the two cysteine residues in the CxxC active site. The pure proteins were used for substrate-binding studies known as “mixed disulfide fishing” in which covalent disulfide-bonded reaction intermediates can be visualized. An unprecedented finding is that both active-site cysteine residues can form mixed disulfides with substrate proteins when the other active-site cysteine is absent, but only the N-terminal active-site cysteine forms stable interactions. A second novelty is that both single-cysteine mutant TrxA proteins form stable homodimers due to thiol oxidation of the remaining active-site cysteine residue. To investigate whether these dimers resemble mixed enzyme-substrate disulfides, the structure of the most abundant dimer, C32S, was characterized by X-ray crystallography. This yielded a high-resolution (1.5Å) X-ray crystallographic structure of a thioredoxin homodimer from a low-GC Gram-positive bacterium. The C32S TrxA dimer can be regarded as a mixed disulfide reaction intermediate of thioredoxin, which reveals the diversity of thioredoxin/substrate-binding modes.  相似文献   

4.
Small heat shock proteins form large cytosolic assemblies from an “α-crystallin domain” (ACD) flanked by sequence extensions. Mutation of a conserved arginine in the ACD of several human small heat shock protein family members causes many common inherited diseases of the lens and neuromuscular system. The mutation R120G in αB-crystallin causes myopathy, cardiomyopathy and cataract. We have solved the X-ray structure of the excised ACD dimer of human αB R120G close to physiological pH and compared it with several recently determined wild-type vertebrate ACD dimer structures. Wild-type excised ACD dimers have a deep groove at the interface floored by a flat extended “bottom sheet.” Solid-state NMR studies of large assemblies of full-length αB-crystallin have shown that the groove is blocked in the ACD dimer by curvature of the bottom sheet. The crystal structure of R120G ACD dimer also reveals a closed groove, but here the bottom sheet is flat. Loss of Arg120 results in rearrangement of an extensive array of charged interactions across this interface. His83 and Asp80 on movable arches on either side of the interface close the groove by forming two new salt bridges. The residues involved in this extended set of ionic interactions are conserved in Hsp27, Hsp20, αA- and αB-crystallin sequences. They are not conserved in Hsp22, where mutation of the equivalent of Arg120 causes neuropathy. We speculate that the αB R120G mutation disturbs oligomer dynamics, causing the growth of large soluble oligomers that are toxic to cells by blocking essential processes.  相似文献   

5.
The hepatitis B virus core gene codes for two closely related antigens: a 21-kDa protein that forms dimers that assemble as multimegadalton capsids, and a 17-kDa protein that also forms dimers but that do not assemble. The proteins, respectively referred to as core antigen (HBcAg) and e-antigen (HBeAg), share a sequence of 149 residues but have different amino- and carboxy-termini. Their structural and serological relationship has long been unclear. With insights gained from recent structural studies on immune complexes of the capsids, the relationship was reassessed using recombinant forms of the antigens and a panel of monoclonal antibodies (mAbs) commonly believed to discriminate between core and e-antigen. Surface plasmon resonance (SPR) was used to measure the affinities, in contrast to previous studies that used more error-prone and less sensitive plate-type assays. Four of the six mAbs did not discriminate between core and e-antigen, nor did they discriminate between e-antigen and dimers of dissociated core antigen capsids. One mAb (3120) was specific for assembled capsids and one (e6) was specific for unassembled dimers. Epitope valency of the e-antigen was also studied, using a sandwich SPR assay where e-antigen was captured with one mAb and probed with a second. The e-antigen is often considered to be a monomeric protein on the basis of monovalent reactivity with antibody pairs specific for either an α or β epitope (in a prior nomenclature for e-antigen specificity). This model, however, is incorrect, because recombinant e-antigen is a stable dimer and its apparent monovalency is due to steric blockage. This was proven by the formation of a 2:1 Fab e6-e-antigen complex. These results suggest new approaches for the isolation of the authentic e-antigen, its biological assay, and its stabilization as an immune complex for structural studies.  相似文献   

6.
The dissociation of apo- and metal-bound human copper-zinc superoxide dismutase (SOD1) dimers induced by the chaotrope guanidine hydrochloride (GdnHCl) or the reductant Tris(2-carboxyethyl)phosphine (TCEP) has been analyzed using analytical ultracentrifugation. Global fitting of sedimentation equilibrium data under native solution conditions (without GdnHCl or TCEP) demonstrate that both the apo- and metal-bound forms of SOD1 are stable dimers. Sedimentation velocity experiments show that apo-SOD1 dimers dissociate cooperatively over the range 0.5-1.0 M GdnHCl. In contrast, metal-bound SOD1 dimers possess a more compact shape and dissociate at significantly higher GdnHCl concentrations (2.0-3.0 M). Reduction of the intrasubunit disulfide bond within each SOD1 subunit by 5-10 mM TCEP promotes dissociation of apo-SOD1 dimers, whereas the metal-bound enzyme remains a stable dimer under these conditions. The Cys-57 --> Ser mutant of SOD1, a protein incapable of forming the intrasubunit disulfide bond, sediments as a monomer in the absence of metal ions and as a dimer when metals are bound. Taken together, these data indicate that the stability imparted to the human SOD1 dimer by metal binding and the formation of the intrasubunit disulfide bond are mediated by independent molecular mechanisms. By combining the sedimentation data with previous crystallographic results, a molecular explanation is provided for the existence of different SOD1 macromolecular shapes and multiple SOD1 dimeric species with different stabilities.  相似文献   

7.
The 70 S ribosomes of Escherichia coli were treated with 2-iminothiolane with the resultant addition of 110 sulfhydryl groups per ribosome. The modified ribosomes were oxidized to promote disulfide bond formation, some of which formed intermolecular crosslinks. About 50% of the crosslinked 70 S ribosomes did not dissociate when exposed to low concentrations of magnesium in the absence of reducting agent. Dissociation took place in the presence of reducing agents, which indicated that the subunits had become covalently linked by disulfide linkages. Proteins extracted from purified crosslinked 70 S ribosomes were first fractionated by polyacrylamide/urea gel electrophoresis. The proteins from sequential slices of these gels were analyzed by two-dimensional polyacrylamide/sodium dodecyl sulfate diagonal gel electrophoresis. Monomeric proteins derived from crosslinked dimers appeared below the diagonal containing non-crosslinked proteins, since the second electrophoresis, but not the first, is run under reducing conditions to cleave the crosslinked species. Final identification of the proteins in each dimer was made by radioiodination of the crosslinked proteins, followed by two-dimensional polyacrylamide/urea gel electrophoresis in the presence of non-radioactive total 70 S proteins as markers. This paper describes the identification of 23 protein dimers that contained one protein from each of the two different ribosomal subunits. The proteins implicated must have some part of their structure in proximity to the other ribosomal subunit and are therefore defined as “interface proteins”. The group of interface proteins thus defined includes 50 S proteins that are part of the 5 S RNA: protein complex and 30 S proteins at the initiation site. Correlations between the crosslinked interface proteins and other functional data are discussed.  相似文献   

8.
Glycosaminoglycan (GAG)-bound and soluble chemokine gradients in the vasculature and extracellular matrix mediate neutrophil recruitment to the site of microbial infection and sterile injury in the host tissue. However, the molecular principles by which chemokine-GAG interactions orchestrate these gradients are poorly understood. This, in part, can be directly attributed to the complex interrelationship between the chemokine monomer-dimer equilibrium and binding geometry and affinities that are also intimately linked to GAG length. To address some of this missing knowledge, we have characterized the structural basis of heparin binding to the murine CXCL1 dimer. CXCL1 is a neutrophil-activating chemokine and exists as both monomers and dimers (Kd = 36 μm). To avoid interference from monomer-GAG interactions, we designed a trapped dimer (dCXCL1) by introducing a disulfide bridge across the dimer interface. We characterized the binding of GAG heparin octasaccharide to dCXCL1 using solution NMR spectroscopy. Our studies show that octasaccharide binds orthogonally to the interhelical axis and spans the dimer interface and that heparin binding enhances the structural integrity of the C-terminal helical residues and stability of the dimer. We generated a quadruple mutant (H20A/K22A/K62A/K66A) on the basis of the binding data and observed that this mutant failed to bind heparin octasaccharide, validating our structural model. We propose that the stability enhancement of dimers upon GAG binding regulates in vivo neutrophil trafficking by increasing the lifetime of “active” chemokines, and that this structural knowledge could be exploited for designing inhibitors that disrupt chemokine-GAG interactions and neutrophil homing to the target tissue.  相似文献   

9.
We analyzed subunit interfaces in 315 homodimers with an X-ray structure in the Protein Data Bank, validated by checking the literature for data that indicate that the proteins are dimeric in solution and that, in the case of the “weak” dimers, the homodimer is in equilibrium with the monomer. The interfaces of the 42 weak dimers, which are smaller by a factor of 2.4 on average than in the remainder of the set, are comparable in size with antibody-antigen or protease-inhibitor interfaces. Nevertheless, they are more hydrophobic than in the average transient protein-protein complex and similar in amino acid composition to the other homodimer interfaces. The mean numbers of interface hydrogen bonds and hydration water molecules per unit area are also similar in homodimers and transient complexes. Parameters related to the atomic packing suggest that many of the weak dimer interfaces are loosely packed, and we suggest that this contributes to their low stability. To evaluate the evolutionary selection pressure on interface residues, we calculated the Shannon entropy of homologous amino acid sequences at 60% sequence identity. In 93% of the homodimers, the interface residues are better conserved than the residues on the protein surface. The weak dimers display the same high degree of interface conservation as other homodimers, but their homologs may be heterodimers as well as homodimers. Their interfaces may be good models in terms of their size, composition, and evolutionary conservation for the labile subunit contacts that allow protein assemblies to share and exchange components, allosteric proteins to undergo quaternary structure transitions, and molecular machines to operate in the cell.  相似文献   

10.
Most proteins adopt a well defined three-dimensional structure; however, it is increasingly recognized that some proteins can exist with at least two stable conformations. Recently, a class of intracellular chloride ion channel proteins (CLICs) has been shown to exist in both soluble and integral membrane forms. The structure of the soluble form of CLIC1 is typical of a soluble glutathione S-transferase superfamily protein but contains a glutaredoxin-like active site. In this study we show that on oxidation CLIC1 undergoes a reversible transition from a monomeric to a non-covalent dimeric state due to the formation of an intramolecular disulfide bond (Cys-24-Cys-59). We have determined the crystal structure of this oxidized state and show that a major structural transition has occurred, exposing a large hydrophobic surface, which forms the dimer interface. The oxidized CLIC1 dimer maintains its ability to form chloride ion channels in artificial bilayers and vesicles, whereas a reducing environment prevents the formation of ion channels by CLIC1. Mutational studies show that both Cys-24 and Cys-59 are required for channel activity.  相似文献   

11.
The assembly of von Willebrand factor multimers in the Golgi apparatus requires D1D2 domains of the von Willebrand factor propeptide, which may act as an oxidoreductase to promote disulfide bond formation or rearrangement between two D3 domains in the mature subunit. This mechanism predicts that the propeptide should form a transient intrachain disulfide bond with the D3 domain before multimerization. Such an intermediate was detected using truncated subunits that simplify the analysis of the multimerization process. When only the D1D2D'D3 region of von Willebrand factor was expressed in baby hamster kidney cells, the propeptide and D'D3 formed an intrachain disulfide-linked species in the endoplasmic reticulum that could be identified by two-dimensional gel electrophoresis after cleavage with thrombin or furin. This intermediate rearranged in the Golgi to form free propeptide and D'D3 dimers that were secreted. A similar intracellular disulfide-linked species was identified in cells expressing the propeptide and D'D3 as separate proteins and in cells expressing full-length von Willebrand factor. These results support a model in which the propeptide acts as an oxidoreductase to promote von Willebrand factor multimerization in the Golgi apparatus.  相似文献   

12.
The first gluconolactonase crystal structure from bacteria has been determined to a resolution of 1.61 Å using X-ray crystallography. It belongs to the senescence marker protein 30/gluconolaconase superfamily but exhibits substrate specificity mainly toward d-glucono-δ-lactone. It forms a novel disulfide-bonded clamshell dimer comprising two doughnut-shaped six-bladed β-propeller domains, yet with an exceptionally long N-terminal subdomain forming an extra helix and four additional β-strands to enclose half of the outermost β-strands of each propeller. Extensive interactions, including H-bonds, salt bridges, disulfide bonds, and coordination bonds, along with numerous bridging water molecules, are present in the interface to institute the “top-to-top” clamshell-type dimer. Three calcium ions per subunit were observed. Two are present in the central water-filled channel, with the top one coordinated to four highly conserved amino acids and is possibly involved in substrate hydrolysis, while the bottom one is coordinated to the backbone oxygen atoms, which is possibly for stabilizing the propeller domain. One calcium ion is situated in the interface also to stabilize the dimer form. Since gluconolactonase is essential in the glucose secondary metabolic pathways leading to the synthesis of pentose, vitamin C, or “antiaging” factors, determination of its tertiary structure should help understand these important biochemical processes.  相似文献   

13.
In biological systems, proteins rarely act as isolated monomers. Association to dimers or higher oligomers is a commonly observed phenomenon. As an example, small heat shock proteins form spherical homo-oligomers of mostly 24 subunits, with the dimeric α-crystallin domain as the basic structural unit. The structural hierarchy of this complex is key to its function as a molecular chaperone. In this article, we analyze the folding and association of the basic building block, the α-crystallin domain dimer, from the hyperthermophilic archaeon Methanocaldococcus jannaschii Hsp16.5 in detail. Equilibrium denaturation experiments reveal that the α-crystallin domain dimer is highly stable against chemical denaturation. In these experiments, protein dissociation and unfolding appear to follow an “all-or-none” mechanism with no intermediate monomeric species populated. When the mechanical stability was determined by single-molecule force spectroscopy, we found that the α-crystallin domain dimer resists high forces when pulled at its termini. In contrast to bulk denaturation, stable monomeric unfolding intermediates could be directly observed in the mechanical unfolding traces after the α-crystallin domain dimer had been dissociated by force. Our results imply that for this hyperthermophilic member of the small heat shock protein family, assembly of the spherical 24mer starts from folded monomers, which readily associate to the dimeric structure required for assembly of the higher oligomer.  相似文献   

14.
The transmembrane enzymes disulfide bond forming enzyme B (DsbB) and vitamin K epoxide reductase (VKOR) are central to oxidative protein folding in the periplasm of prokaryotes. Catalyzed formation of structural disulfide bonds in proteins also occurs in the cytoplasm of some hyperthermophilic prokaryotes through currently, poorly defined mechanisms. We aimed to determine whether DsbB and VKOR can be inverted in the membrane with retention of activity. By rational design of inversion of membrane topology, we engineered DsbB mutants that catalyze disulfide bond formation in the cytoplasm of Escherichia coli. This represents the first engineered inversion of a transmembrane protein with demonstrated conservation of activity and substrate specificity. This successful designed engineering led us to identify two naturally occurring and oppositely oriented VKOR homologues from the hyperthermophile Aeropyrum pernix that promote oxidative protein folding in the periplasm or cytoplasm, respectively, and hence defines the probable route for disulfide bond formation in the cytoplasm of hyperthermophiles. Our findings demonstrate how knowledge on the determinants of membrane protein topology can be used to de novo engineer a metabolic pathway and to unravel an intriguingly simple evolutionary scenario where a new “adaptive” cellular process is constructed by means of membrane protein topology inversion.  相似文献   

15.
Thymidylate synthase (TS), a dimeric enzyme, forms large soluble aggregates at concentrations of urea (3.3-5M), well below that required for complete denaturation, as established by fluorescence and size-exclusion chromatography. In contrast to the wild-type enzyme, an engineered mutant of TS (T155C/E188C/C244T), TSMox, in which two subunits are crosslinked by disulfide bridges between residues 155-188' and 188-155' does not show this behavior. Aggregation behavior is restored upon disulfide bond reduction in the mutant protein, indicating the involvement of interface segments in forming soluble associated species. Intermolecular disulfide crosslinking has been used as a probe to investigate the formation of larger non-native aggregates. The studies argue for the formation of large multimeric species via a sticky patch of polypeptide from the dimer interface region that becomes exposed on partial unfolding. Covalent reinforcement of relatively fragile protein-protein interfaces may be a useful strategy in minimizing aggregation of non-native structures in multimeric proteins.  相似文献   

16.
Cleavable Crosslinking reagents were used to study interactions among proteins of the surface coat of Trypanosoma brucei. The proteins were resolved by two-dimensional polyacrylamide gel electrophoresis in sodium dodecyl sulfate. When intact cells were treated with dithiobis(succinimidylpropionate), we obtained extensive intermolecular Crosslinking of major variable surface coat glycoprotein (VSCG) molecules. This reagent generated no apparent crosslinks between VSCG and other membrane-associated proteins. Complete conversion to oligomers equal to or greater than octamers occurred within 20 min. When purified VSCG in solution was treated with dithiobis(succimidylpropionate), dimers were found. A complex of Cu2+ and 1,10-phenanthroline was used to catalyze air oxidation of adjacent sulfhydryls to disulfide bonds; however, no crosslinking among VSCG molecules nor between VSCG and other proteins was observed. The results presented indicate that VSCG in solution exists predominately in the form of dimers. Whether VSCG in situ also occurred as dimers could not be determined; however, since we observed trimeric and tetrameric forms of VSCG when untreated cells were analyzed, it is likely that weak interactions occur among the protein molecules. These interactions are less stable than the dimer association observed with purified VSCG. Finally, the analysis indicated that VSCGs of this stock of T. brucei, derived from UGANDA/ 60/TREU/164[ETat3], contained at least one intramolecular disulfide bond. We examined T. brucei stocks 427 and EATRO 110 and obtained similar results. Thus, it appears that intramolecular disulfide bonding is a general feature of T. brucei VSCGs.  相似文献   

17.
Fur (ferric uptake regulator) proteins are involved in the control of a variety of processes in most prokaryotes. Although it is assumed that this regulator binds its DNA targets as a dimer, the way in which this interaction occurs remains unknown. We have focused on FurA from the cyanobacterium Anabaena sp. PCC 7120. To assess the molecular mechanism by which FurA specifically binds to “iron boxes” in PfurA, we examined the topology arrangement of FurA–DNA complexes by atomic force microscopy. Interestingly, FurA–PfurA complexes exhibit several populations, in which one is the predominant and depends clearly on the regulator/promoter ratio on the environment. Those results together with EMSA and other techniques suggest that FurA binds PfurA using a sequential mechanism: (i) a monomer specifically binds to an “iron box” and bends PfurA; (ii) two situations may occur, that a second FurA monomer covers the free “iron box" or that joins to the previously used forming a dimer which would maintain the DNA kinked; (iii) trimerization in which the DNA is unbent; and (iv) finally undergoes a tetramerization; the next coming molecules cover the DNA strands unspecifically. In summary, the bending appears when an “iron box” is bound to one or two molecules and decreases when both “iron boxes” are covered. These results suggest that DNA bending contributes at the first steps of FurA repression promoting the recruitment of new molecules resulting in a fine regulation in the Fur-dependent cluster associated genes.  相似文献   

18.
Su HP  Lin DY  Garboczi DN 《Journal of virology》2006,80(15):7706-7713
The possibility of the release of smallpox virus into a predominantly nonimmunized population highlights the importance of understanding poxvirus biology. Poxviruses encode a conserved pathway that is required to oxidize disulfide bonds in nascent viral proteins that fold in the reducing environment of the eukaryotic host cytoplasm. We present the structure of the last enzyme of the vaccinia virus pathway, G4, which is almost identical in smallpox virus. G4 catalyzes the formation of disulfide bonds in proteins that are critical for virus maturation and host cell infection. G4 contains a thioredoxin fold and a Cys-X-X-Cys active site. In solution, G4 monomers and dimers are observed. In the crystal, G4 is found as a dimer that buries 4,500 A(2) in the interface and occludes the active site, which could protect the reactive disulfide from reduction in the cytoplasm. The structure serves as a model for drug design targeting viral disulfide bond formation.  相似文献   

19.
The emergence of resistance to vancomycin and related glycopeptide antibiotics is spurring efforts to develop new antimicrobial therapeutics. High-resolution structural information about antibiotic-ligand recognition should prove valuable in the rational design of improved drugs. We have determined the X-ray crystal structure of the complex of vancomycin with N-acetyl-d-Ala-d-Ala, a mimic of the natural muramyl peptide target, and refined this structure at a resolution of 1.3 Å to R and Rfree values of 0.172 and 0.195, respectively. The crystal asymmetric unit contains three back-back vancomycin dimers; two of these dimers participate in ligand-mediated face-face interactions that produce an infinite chain of molecules running throughout the crystal. The third dimer packs against the side of a face-face interface in a tight “side-side” interaction that involves both polar contacts and burial of hydrophobic surface. The trimer of dimers found in the asymmetric unit is essentially identical to complexes seen in three other crystal structures of glycopeptide antibiotics complexed with peptide ligands. These four structures are derived from crystals belonging to different space groups, suggesting that the trimer of dimers may not be simply a crystal packing artifact and prompting us to ask if ligand-mediated oligomerization could be observed in solution. Using size-exclusion chromatography, dynamic light scattering, and small-angle X-ray scattering, we demonstrate that vancomycin forms discrete supramolecular complexes in the presence of tripeptide ligands. Size estimates for these complexes are consistent with assemblies containing four to six vancomycin monomers.  相似文献   

20.
Surface proteins Shr, Shp, and the ATP-binding cassette (ABC) transporter HtsABC are believed to make up the machinery for heme uptake in Streptococcus pyogenes. Shp transfers its heme to HtsA, the lipoprotein component of HtsABC, providing the only experimentally demonstrated example of direct heme transfer from a surface protein to an ABC transporter in Gram-positive bacteria. To understand the structural basis of heme transfer in this system, the heme-binding domain of Shp (Shp180) was crystallized, and its structure determined to a resolution of 2.1 Å. Shp180 exhibits an immunoglobulin-like β-sandwich fold that has been recently found in other pathogenic bacterial cell surface heme-binding proteins, suggesting that the mechanisms of heme acquisition are conserved. Shp shows minimal amino acid sequence identity to these heme-binding proteins and the structure of Shp180 reveals a unique heme-iron coordination with the axial ligands being two methionine residues from the same Shp molecule. A negative electrostatic surface of protein structure surrounding the heme pocket may serve as a docking interface for heme transfer from the more basic outer cell wall heme receptor protein Shr. The crystal structure of Shp180 reveals two exogenous, weakly bound hemins, which form a large interface between the two Shp180 molecules in the asymmetric unit. These “extra” hemins form a stacked pair with a structure similar to that observed previously for free hemin dimers in aqueous solution. The propionates of the protein-bound heme coordinate to the iron atoms of the exogenous hemin dimer, contributing to the stability of the protein interface. Gel filtration and analytical ultracentrifugation studies indicate that both full-length Shp and Shp180 are monomeric in dilute aqueous solution.  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