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.
Mitochondria from all plants, many fungi and some protozoa contain a cyanide-resistant, alternative oxidase that functions in parallel with cytochrome c oxidase as the terminal oxidase on the electron transfer chain. Characterization of the structural and potential regulatory features of the alternative oxidase has advanced considerably in recent years. The active site is proposed to contain a di-iron center belonging to the ribonucleotide reductase R2 family and modeling of a four-helix bundle to accommodate this active site within the C-terminal two-thirds of the protein has been carried out. The structural features of this active site are conserved among all known alternative oxidases. The post-translational regulatory features of the alternative oxidase are more variable among organisms. The plant oxidase is dimeric and can be stimulated by either alpha-keto acids or succinate, depending upon the presence or absence, respectively, of a critical cysteine residue found in a conserved block of amino acids in the N-terminal region of the plant protein. The fungal and protozoan alternative oxidases generally exist as monomers and are not subject to organic acid stimulation but can be stimulated by purine nucleotides. The origins of these diverse regulatory features remain unknown but are correlated with sequence differences in the N-terminal third of the protein.  相似文献   

2.
Two Cys residues, Cys(I) and Cys(II), are present in most plant alternative oxidases (AOXs). Cys(I) inactivates AOX by forming a disulfide bond with the corresponding Cys(I) residue on the adjacent subunit of the AOX homodimer. When reduced, Cys(I) associates with alpha-keto acids, such as pyruvate, to activate AOX, an effect mimicked by charged amino acid substitutions at the Cys(I) site. Cys(II) may also be a site of AOX activity regulation, through interaction with the small alpha-keto acid, glyoxylate. Comparison of Arabidopsis AOX1a (AtAOX1a) mutants with single or double substitutions at Cys(I) and Cys(II) confirmed that glyoxylate interacted with either Cys, while the effect of pyruvate (or succinate for AtAOX1a substituted with Ala at Cys(I)) was limited to Cys(I). A variety of Cys(II) substitutions constitutively activated AtAOX1a, indicating that neither the catalytic site nor, unlike at Cys(I), charge repulsion is involved. Independent effects at each Cys were suggested by lack of Cys(II) substitution interference with pyruvate stimulation at Cys(I), and close to additive activation at the two sites. However, results obtained using diamide treatment to covalently link the AtAOX1a subunits by the disulfide bond indicated that Cys(I) must be in the reduced state for activation at Cys(II) to occur.  相似文献   

3.
The homodimeric cyanide-resistant alternative oxidase of plant mitochondria reduces oxygen to water without forming ATP. Arabidopsis thaliana alternative oxidase AOX1a is stimulated by pyruvate or other alpha-keto acids associating with a regulatory cysteine at position 78, by succinate in a serine-78 mutant, and by site-directed mutation of position 78 to glutamate. The mechanism of activation was explored with additional amino acid substitutions made at Cys-78 in AOX1a, which was functionally expressed in Escherichia coli. Oxidases with positively charged substitutions (Lys and Arg) were insensitive to pyruvate or succinate but were more active than the wild type without pyruvate. Uncharged substitutions (Gln, Leu) produced an inactive enzyme. These results indicate that activation may be due to conformational changes caused by charge repulsion between the dimer subunits and not through a direct role of alpha-keto acids in catalysis. Oxygen isotope fractionation experiments suggest that the charge of the amino acid at position 78 also affects the entry of oxygen into the active site. Therefore, the N-terminal portion of the protein containing residue 78 can indirectly affect both catalysis at the diiron active site and the path of oxygen to that site. In addition, both positively and negatively substituted alternative oxidases were stimulated by glyoxylate, suggesting the presence of a second activation site, possibly Cys-128.  相似文献   

4.
The Erv flavoenzymes contain a compact module that catalyzes the pairing of cysteine thiols into disulfide bonds. High-resolution structures of plant, animal, and fungal Erv enzymes that function in different contexts and intracellular compartments have been determined. Structural features can be correlated with biochemical properties, revealing how core sulfhydryl oxidase activity has been tailored to various functional niches. The introduction of disulfides into cysteine-containing substrates by Erv sulfhydryl oxidases is compared with the mechanisms used by NADPH-driven disulfide reductases and thioredoxin-like oxidoreductases to reduce and transfer disulfides, respectively.  相似文献   

5.
The Moco (molybdenum cofactor) sulfurase ABA3 from Arabidopsis thaliana catalyses the sulfuration of the Moco of aldehyde oxidase and xanthine oxidoreductase, which represents the final activation step of these enzymes. ABA3 consists of an N-terminal NifS-like domain that exhibits L-cysteine desulfurase activity and a C-terminal domain that binds sulfurated Moco. The strictly conserved Cys430 in the NifS-like domain binds a persulfide intermediate, which is abstracted from the substrate L-cysteine and finally needs to be transferred to the Moco of aldehyde oxidase and xanthine oxidoreductase. In addition to Cys?3?, another eight cysteine residues are located in the NifS-like domain, with two of them being highly conserved among Moco sulfurase proteins and, at the same time, being in close proximity to Cys?3?. By determination of the number of surface-exposed cysteine residues and the number of persulfide-binding cysteine residues in combination with the sequential substitution of each of the nine cysteine residues, a second persulfide-binding cysteine residue, Cys2??, was identified. Furthermore, the active-site Cys?3? was found to be located on top of a loop structure, formed by the two flanking residues Cys?2? and Cys?3?, which are likely to form an intramolecular disulfide bridge. These findings are confirmed by a structural model of the NifS-like domain, which indicates that Cys?2? and Cys?3? are within disulfide bond distance and that a persulfide transfer from Cys?3? to Cys2?? is indeed possible.  相似文献   

6.
Regulation of alternative oxidase activity in higher plants   总被引:10,自引:0,他引:10  
Plant mitochondria contain two terminal oxidases: cytochrome oxidase and the cyanideinsensitive alternative oxidase. Electron partioning between the two pathways is regulated by the redox poise of the ubiquinone pool and the activation state of the alternative oxidase. The alternative oxidase appears to exist as a dimer which is active in the reduced, noncovalently linked form and inactive when in the oxidized, covalently linked form. Reduction of the oxidase in isolated tobacco mitochondria occurs upon oxidation of isocitrate or malate and may be mediated by matrix NAD(P)H. The activity of the reduced oxidase is governed by certain other organic acids, notably pyruvate, which appear to interact directly with the enzyme. Pyruvate alters the interaction between the alternative oxidase and ubiquinol so that the oxidase becomes active at much lower levels of ubiquinol and competes with the cytochrome pathway for electrons. These requirements for activation of the alternative oxidase constitute a sophisticated feed-forward control mechanism which determines the extent to which electrons are directed away from the energy-conserving cytochrome pathway to the non-energy conserving alternative oxidase. Such a mechanism fits well with the proposed role of the alternative oxidase as a protective enzyme which prevents over-reduction of the cytochrome chain and fermentation of accumulated pyruvate.  相似文献   

7.
The human type 1 (placenta, breast tumors) and type 2 (gonads, adrenals) isoforms of 3beta-hydroxysteroid dehydrogenase/isomerase (3beta-HSD) are key enzymes in steroidogenic pathways leading to the production of all active steroid hormones. Kinetic analyses of purified 3beta-HSD1 show that the Michaelis-Menten constants (Km) for substrates and cofactor are decreased dramatically (three- to eight-fold) by the addition of beta-mercaptoethanol (BME), which suggest that a disulfide bond may be critical to ligand utilization. Western immunoblots and SDS-PAGE of purified 3beta-HSD1 in the presence or absence of BME showed a lack of intersubunit disulfide bonds in the dimeric enzyme. The Rossmann-fold domain of 3beta-HSD1 contains two Cys residues, Cys72 and Cys111, which are capable of forming an intrasubunit disulfide bond based on their proximity in our structural model. Our structural model also predicts that Cys83 may affect the orientation of substrate and cofactor. To test these predictions, the C72S, C72F, C111S, C111A, C83S and C83A mutants of 3beta-HSD1 were produced, expressed, and purified. BME failed to diminish the Km values of substrate and cofactor for C72S, C72F, C111S and C111A but produced a 2.5 decrease in Km values for C83A ligands similar to wild-type 3beta-HSD. Thus, our results support the presence of an intrasubunit disulfide bond between Cys72 and Cys111 that participates in the tertiary structure of the Rossmann-fold domain. Although C83S had no enzyme activity, the C83A mutant enzyme exhibited two- to five-fold higher Km values for substrate and cofactor but had similar K(cat) values compared to wild-type 3beta-HSD. These data characterize the roles of Cys residues in 3beta-HSD and validate the predictions of our structural model.  相似文献   

8.
Respiratory rates involving the alternative oxidase (AO) were studied in mitochondria from Tapesia acuformis. There was no evidence for regulation by pyruvate, in contrast with plant AO. The site of interaction of pyruvate with the plant AO is a conserved cysteine. The primary sequence was obtained for AO from Magnaporthe grisea and compared with four published sequences for fungal AO. In all cases this cysteine was absent. Sequence data were obtained for the C-terminal domain of a further five fungal AOs. In this region the fungal sequences were all consistent with a four-helix, di-iron binding structure as in the ferritin-fold family. A molecular model of this domain was deduced from the structure of Delta-9 desaturase. This is in general agreement with that developed for plant AOs, despite very low sequence identity between the two kingdoms. Further modelling indicated an appropriate active site for binding of ubiquinol, required in the AO redox reaction.  相似文献   

9.
The N-terminal cysteine-rich somatomedin B (SMB) domain (residues 1-44) of the human glycoprotein vitronectin contains the high-affinity binding sites for plasminogen activator inhibitor-1 (PAI-1) and the urokinase receptor (uPAR). We previously showed that the eight cysteine residues of recombinant SMB (rSMB) are organized into four disulfide bonds in a linear uncrossed pattern (Cys(5)-Cys(9), Cys(19)-Cys(21), Cys(25)-Cys(31), and Cys(32)-Cys(39)). In the present study, we use an alternative method to show that this disulfide bond arrangement remains a major preferred one in solution, and we determine the solution structure of the domain using NMR analysis. The solution structure shows that the four disulfide bonds are tightly packed in the center of the domain, replacing the traditional hydrophobic core expected for a globular protein. The few noncysteine hydrophobic side chains form a cluster on the outside of the domain, providing a distinctive binding surface for the physiological partners PAI-1 and uPAR. The hydrophobic surface consists mainly of side chains from the loop formed by the Cys(25)-Cys(31) disulfide bond, and is surrounded by conserved acidic and basic side chains, which are likely to contribute to the specificity of the intermolecular interactions of this domain. Interestingly, the overall fold of the molecule is compatible with several arrangements of the disulfide bonds. A number of different disulfide bond arrangements were able to satisfy the NMR restraints, and an extensive series of conformational energy calculations performed in explicit solvent confirmed that several disulfide bond arrangements have comparable stabilization energies. An experimental demonstration of the presence of alternative disulfide conformations in active rSMB is provided by the behavior of a mutant in which Asn(14) is replaced by Met. This mutant has the same PAI-1 binding activity as rVN1-51, but its fragmentation pattern following cyanogen bromide treatment is incompatible with the linear uncrossed disulfide arrangement. These results suggest that active forms of the SMB domain may have a number of allowed disulfide bond arrangements as long as the Cys(25)-Cys(31) disulfide bond is preserved.  相似文献   

10.
Penaeidins are a family of antimicrobial peptides of 47-63 residues isolated from several species of shrimp. These peptides display a proline-rich domain (N-terminal part) and a cysteine-rich domain (C-terminal part) stabilized by three conserved disulfide bonds whose arrangement has not yet been characterized. The recombinant penaeidin-3a of Litopenaeus vannamei (63 residues) and its [T8A]-Pen-3a analogue were produced in Saccharomyces cerevisiae and showed similar antimicrobial activity. The solution structure of the [T8A]-Pen-3a analogue was determined by using two-dimensional 1H NMR and simulated annealing calculations. The proline-rich domain, spanning residues 1-28 was found to be unconstrained. In contrast, the cysteine-rich domain, spanning residues 29-58, displays a well defined structure, which consists of an amphipathic helix (41-50) linked to the upstream and the downstream coils by two disulfide bonds (Cys32-Cys47 and Cys48-Cys55). These two coils are in turn linked together by the third disulfide bond (Cys36-Cys54). Such a disulfide bond packing, which is in agreement with the analysis of trypsin digests by ESI-MS, contributes to the highly hydrophobic core. Side chains of Arg45 and Arg50, which belong to the helix, and side chains of Arg37 and Arg53, which belong to the upstream and the downstream coils, are located in two opposite parts of this globular and compact structure. The environment of these positively charged residues, either by hydrophobic clusters at the surface of the cysteine-rich domain or by sequential hydrophobic residues in the unconstrained proline-rich domain, gives rise to the amphipathic character required for antimicrobial peptides. We hypothesize that the antimicrobial activity of penaeidins can be explained by a cooperative effect between the proline-rich and cysteine-rich features simultaneously present in their sequences.  相似文献   

11.
Oxidation is an important biochemical defense mechanism, but it also elicits toxicity; therefore, oxidation must be under strict control. In phagocytotic events in neutrophils, the voltage-gated H+ (Hv) channel is a key regulator of the production of reactive oxygen species against invading bacteria. The cytoplasmic domain of the Hv channel forms a dimeric coiled coil underpinning a dimerized functional unit. Importantly, in the alignment of the coiled-coil core, a conserved cysteine residue forms a potential intersubunit disulfide bond. In this study, we solved the crystal structures of the coiled-coil domain in reduced, oxidized, and mutated (Cys → Ser) states. The crystal structures indicate that a pair of Cys residues forms an intersubunit disulfide bond dependent on the redox conditions. CD spectroscopy revealed that the disulfide bond increases the thermal stability of the coiled-coil protein. We also reveal that two thiol modifier molecules are able to bind to Cys in a redox-dependent manner without disruption of the dimeric coiled-coil assembly. Thus, the biochemical properties of the cytoplasmic coiled-coil domain in the Hv channel depend on the redox condition, which may play a role in redox sensing in the phagosome.  相似文献   

12.
Part of the dimer and B/C domain interface of the Escherichia coli mannitol permease (EII(mtl)) has been identified by the generation of disulfide bridges in a single-cysteine EII(mtl), with only the activity linked Cys(384) in the B domain, and in a double-cysteine EII(mtl) with cysteines at positions 384 and 124 in the first cytoplasmic loop of the C domain. The disulfide bridges were formed in the enzyme in inside-out membrane vesicles and in the purified enzyme by oxidation with Cu(II)-(1,10-phenanthroline)(3), and they were visualized by SDS-polyacrylamide gel electrophoresis. Discrimination between possible disulfide bridges in the dimeric double-cysteine EII(mtl) was done by partial digestion of the protein and the formation of heterodimers, in which the cysteines were located either on different subunits or on one subunit. The disulfide bridges that were identified are an intersubunit Cys(384)-Cys(384), an intersubunit Cys(124)-Cys(124), an intersubunit Cys(384)-Cys(124), and an intrasubunit Cys(384)-Cys(124). The disulfide bridges between the B and C domain were observed with purified enzyme and confirmed by matrix-assisted laser desorption ionization-time of flight mass spectrometry. Mannitol did not influence the formation of the disulfide between Cys(384) and Cys(124). The close proximity of the two cysteines 124 was further confirmed with a separate C domain by oxidation with Cu(II)-(1,10-phenanthroline)(3) or by reactions with dimaleimides of different length. The data in combination with other work show that the first cytoplasmic loop around residue 124 is located at the dimer interface and involved in the interaction between the B and C domain.  相似文献   

13.
Pyruvate, alpha-ketoglutarate, and branched-chain alpha-keto acids which were transaminated products of valine, leucine, and isoleucine inhibited glycine decarboxylation by rat liver mitochondria. However, glycine synthesis (the reverse reaction of glycine decarboxylation) was stimulated by those alpha-keto acids with the concomitant decarboxylation of alpha-keto acid added in the absence of NADH. Both the decarboxylation and the synthesis of glycine by mitochondrial extract were affected similarly by alpha-ketoglutarate and branched-chain alpha-keto acids in the absence of pyridine nucleotide, but not by pyruvate. This failure of pyruvate to have an effect was due to the lack of pyruvate oxidation activity in the mitochondrial extract employed. It indicated that those alpha-keto acids exerted their effects by providing reducing equivalents to the glycine cleavage system, possibly through lipoamide dehydrogenase, a component shared by the glycine cleavage system and alpha-keto acid dehydrogenase complexes. On the decarboxylation of pyruvate, alpha-ketoglutarate, and branched-chain alpha-keto acids in intact mitochondria, those alpha-keto acids inhibited one another. In similar experiments with mitochondrial extract, decarboxylations of alpha-ketoglutarate and branched-chain alpha-keto acid were inhibited by branched-chain alpha-keto acid and alpha-ketoglutarate, respectively, but not by pyruvate. NADH was unlikely to account for the inhibition. We suggest that the lipoamide dehydrogenase component is an indistinguishable constituent among alpha-keto acid dehydrogenase complexes and the glycine cleavage system in mitochondria in nature, and that lipoamide dehydrogenase-mediated transfer of reducing equivalents might regulate alpha-keto acid oxidation as well as glycine oxidation.  相似文献   

14.
Abstract

Kinetic properties of novel amine oxidases isolated from a mold Aspergillus niger AKU 3302 were compared to those of typical plant amine oxidase from pea seedling (EC 1.4.3.6). Pea amine oxidase showed highest affinity with diamines, such as putrescine and cadaverine, while fungal enzymes oxidized preferably n-hexylamine and tyramine. All enzymes were inhibited by carbonyl reagents, copper chelating agents, some substrate analogs and alkaloids, but there were quite significant differences in the sensitivity and inhibition modes. Aminoguanidine, which strongly inhibited pea amine oxidases showed only little effect on fungal enzymes. Substrate analogs such as 1,5-diamino-3-pentanone and l-amino-3-phenyl-3-propanone, which were potent competitive inhibitors of pea amine oxidases, inhibited fungal enzymes much more weakly and non competitively. Also various alkaloids behaving as competitive inhibitors of pea amine oxidases inhibited the fungal enzymes non competitively. Very surprising was the potent inhibition of fungal enzymes by artificial substrates of pea amine oxidases, E- and Z-1,4-diamino-2-butene. The relationships between the different inhibition modes and possible binding at the active site are discussed.  相似文献   

15.
The alpha 2 beta 2 structure of the insulin receptor has previously been shown to involve one disulfide bridge between the alpha-subunits in the region containing Cys435, Cys468 and Cys524. We have digested the soluble extracellular domain of the insulin receptor with succinylated trypsin, partially separated the resulting peptides, and sequenced a number of fractions. The peptides containing Cys435 and Cys468 appeared in the same fraction, indicating that these two form a disulfide bond, and in another fraction we found the sequence of the peptide containing Cys524. Since it has been shown that the extracellular domain of the insulin receptor has no free thiols and since no other sequences containing cysteine were found in these fractions, we conclude that Cys524 forms a disulfide bond to the Cys524 in the other alpha-subunit.  相似文献   

16.
The Escherichia coli disulfide isomerase, DsbC is a V-shaped homodimer with each monomer comprising a dimerization region that forms part of a putative peptide-binding pocket and a thioredoxin catalytic domain. Disulfide isomerases from prokaryotes and eukaryotes exhibit little sequence homology but display very similar structural organization with two thioredoxin domains facing each other on top of the dimerization/peptide-binding region. To aid the understanding of the mechanistic significance of thioredoxin domain dimerization and of the peptide-binding cleft of DsbC, we constructed a series of protein chimeras comprising unrelated protein dimerization domains fused to thioredoxin superfamily enzymes. Chimeras consisting of the dimerization domain and the alpha-helical linker of the bacterial proline cis/trans isomerase FkpA and the periplasmic oxidase DsbA gave rise to enzymes that catalyzed the folding of multidisulfide substrate proteins in vivo with comparable efficiency to E. coli DsbC. In addition, expression of FkpA-DsbAs conferred modest resistance to CuCl2, a phenotype that depends on disulfide bond isomerization. Selection for resistance to elevated CuCl2 concentrations led to the isolation of FkpA-DsbA mutants containing a single amino acid substitution that changed the active site of the DsbA domain from CPHC into CPYC, increasing the similarity to the DsbC active site (CGYC). Unlike DsbC, which is resistant to oxidation by DsbB-DsbA and does not normally catalyze disulfide bond formation under physiological conditions, the FkpA-DsbA chimeras functioned both as oxidases and isomerases. The engineering of these efficient artificial isomerases delineates the key features of catalysis of disulfide bond isomerization and enhances our understanding of its evolution.  相似文献   

17.
Erv2p is a small, dimeric FAD-dependent sulfhydryl oxidase that generates disulfide bonds in the lumen of the endoplasmic reticulum. Mutagenic and structural studies suggest that Erv2p uses an internal thiol-transfer relay between the FAD-proximal active site cysteine pair (Cys121-Cys124) and a second cysteine pair (Cys176-Cys178) located in a flexible, substrate-accessible C-terminal tail of the adjacent dimer subunit. Here, we demonstrate that Cys176 and Cys178 are the only amino acids in the tail region required for disulfide transfer and that their relative positioning within the tail peptide is important for activity. However, intragenic suppressor mutations could be isolated that bypass the requirement for Cys176 and Cys178. These mutants were found to disrupt Erv2p dimerization and to increase the activity of Erv2p for thiol substrates such as glutathione. We propose that the two Erv2p subunits act together to direct the disulfide transfer to specific substrates. One subunit provides the catalytic domain composed of the active site cysteine residues and the FAD cofactor, while the second subunit appears to have two functions: it facilitates disulfide transfer to substrates via the tail cysteine residues, while simultaneously shielding the active site cysteine residues from non-specific reactions.  相似文献   

18.
Sanejouand YH 《Proteins》2004,57(1):205-212
It has recently been shown that disulfide bond Cys130-Cys159 in domain 2 of monomeric CD4 is involved in the formation of CD4 disulfide-bonded dimers on cell surfaces and that it can influence the permissiveness of cells to HIV infection. Because this disulfide bond is buried in the monomer, a large conformational change must take place in order to allow for such disulfide exchange. Using standard optimization techniques, whose efficiency was first checked in the well-documented CD2 case, we have shown that 3D domain swapping is a likely candidate for the conformational change, the hinge loop, or linker, being loop E-F. Indeed, as a consequence of domain swapping, because Cys130 and Cys159 belong to beta-strands C and F, respectively, two disulfide bonds become established between Cys130 in one monomer and Cys159 in the other one. Such a disulfide exchange has already been observed when the nuclear magnetic resonance (NMR) structure of the prion protein was compared to the crystallographic, dimeric one. In both cases, domain swapping implies disulfide exchange because the linker is located in the sequence between two disulfide-bonded cysteines. As in the CD2 case, the proposed configuration of the CD4 dimer is found as a pair of neighboring monomers in the crystallographic unit cell. Moreover, because in this configuration the epitope of monoclonal antibody MT151, which does not compete with Gp120 for CD4 binding, is in the cleft between the pair of CD4 monomers, it is suggested that MT151 achieves its HIV-blocking activity by interfering with the formation of CD4 domain-swapped dimers on cell surface.  相似文献   

19.
The Gel/Gas/Phr family of fungal beta(1,3)-glucanosyltransferases plays an important role in cell wall biogenesis by processing the main component beta(1,3)-glucan. Two subfamilies are distinguished depending on the presence or absence of a C-terminal cysteine-rich domain, denoted "Cys-box." The N-terminal domain (NtD) contains the catalytic residues for transglycosidase activity and is separated from the Cys-box by a linker region. To obtain a better understanding of the structure and function of the Cys-box-containing subfamily, we identified the disulfide bonds in Gas2p from Saccharomyces cerevisiae by an improved mass spectrometric methodology. We mapped two separate intra-domain clusters of three and four disulfide bridges. One of the bonds in the first cluster connects a central Cys residue of the NtD with a single conserved Cys residue in the linker. Site-directed mutagenesis of the Cys residue in the linker resulted in an endoplasmic reticulum precursor that was not matured and underwent a gradual degradation. The relevant disulfide bond has a crucial role in folding as it may stabilize the NtD and facilitate its interaction with the C-terminal portion of a Gas protein. The four disulfide bonds in the Cys-box are arranged in a manner consistent with a partial structural resemblance with the plant X8 domain, an independent carbohydrate-binding module that possesses only three disulfide bonds. Deletion of the Cys-box in Gas2 or Gas1 proteins led to the formation of an NtD devoid of any enzymatic activity. The results suggest that the Cys-box is required for proper folding of the NtD and/or substrate binding.  相似文献   

20.
A functional disulfide bond in both the HIV envelope glycoprotein, gp120, and its immune cell receptor, CD4, is involved in viral entry, and compounds that block cleavage of the disulfide bond in these proteins inhibit HIV entry and infection. The disulfide bonds in both proteins are cleaved at the cell surface by the small redox protein, thioredoxin. The target gp120 disulfide and its mechanism of cleavage were determined using a thioredoxin kinetic trapping mutant and mass spectrometry. A single disulfide bond was cleaved in isolated and cell surface gp120, but not the gp160 precursor, and the extent of the reaction was enhanced when gp120 was bound to CD4. The Cys(32) sulfur ion of thioredoxin attacks the Cys(296) sulfur ion of the gp120 V3 domain Cys(296)-Cys(331) disulfide bond, cleaving the bond. Considering that V3 sequences largely determine the chemokine receptor preference of HIV, we propose that cleavage of the V3 domain disulfide, which is facilitated by CD4 binding, regulates chemokine receptor binding. There are 20 possible disulfide bond configurations, and, notably, the V3 domain disulfide has the same unusual -RHStaple configuration as the functional disulfide bond cleaved in CD4.  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