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.
N‐acetylglucosamine 6‐phosphate deacetylase (NagA) catalyzes the conversion of N‐acetylglucosamine‐6‐phosphate to glucosamine‐6‐phosphate in amino sugar catabolism. This conversion is an essential step in the catabolism of sialic acid in several pathogenic bacteria, including Pasteurella multocida, and thus NagA is identified as a potential drug target. Here, we report the unique structural features of NagA from P. multocida (PmNagA) resolved to 1.95 Å. PmNagA displays an altered quaternary architecture with unique interface interactions compared to its close homolog, the Escherichia coli NagA (EcNagA). We confirmed that the altered quaternary structure is not a crystallographic artifact using single particle electron cryo‐microscopy. Analysis of the determined crystal structure reveals a set of hot‐spot residues involved in novel interactions at the dimer‐dimer interface. PmNagA binds to one Zn2+ ion in the active site and demonstrates kinetic parameters comparable to other bacterial homologs. Kinetic studies reveal that at high substrate concentrations (~10‐fold the KM), the tetrameric PmNagA displays hysteresis similar to its distant neighbor, the dimeric Staphylococcus aureus NagA (SaNagA). Our findings provide key information on structural and functional properties of NagA in P. multocida that could be utilized to design novel antibacterials.  相似文献   

2.
The diuretic drug ethacrynic acid (EA), both an inhibitor and substrate of pi class glutathione S‐transferase (GST P1‐1), has been tested in clinical trials as an adjuvant in chemotherapy. We recently studied the role of the active site residue Tyr‐108 in binding EA to the enzyme and found that the analysis was complicated by covalent binding of this drug to the highly reactive Cys‐47. Previous attempts to eliminate this binding by chemical modification yielded ambiguous results and therefore we decided here to produce a double mutant C47S/Y108V by site directed mutagenesis and further expression in Escherichia coli and the interaction of EA and its GSH conjugate (EASG) examined by calorimetric studies and X‐ray diffraction. Surprisingly, in the absence of Cys‐47, Cys‐101 (located at the dimer interface) becomes a target for modification by EA, albeit at a lower conjugation rate than Cys‐47. The Cys‐47 → Ser mutation in the double mutant enzyme induces a positive cooperativity between the two subunits when ligands with affinity to G‐site bind to enzyme. However, this mutation does not seem to affect the thermodynamic properties of ligand binding to the electrophilic binding site (H‐site) and the thermal or chemical stability of this double mutant does not significantly affect the unfolding mechanism in either the absence or presence of ligand. Crystal structures of apo and an EASG complex are essentially identical with a few exceptions in the H‐site and in the water network at the dimer interface. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

3.
3,4‐Dihydroxy‐2‐butanone‐4‐phosphate synthase (DHBPS) encoded by ribB gene is one of the first enzymes in riboflavin biosynthesis pathway and catalyzes the conversion of ribulose‐5‐phosphate (Ru5P) to 3,4‐dihydroxy‐2‐butanone‐4‐phosphate and formate. DHBPS is an attractive target for developing anti‐bacterial drugs as this enzyme is essential for pathogens, but absent in humans. The recombinant DHBPS enzyme of Salmonella requires magnesium ion for its activity and catalyzes the formation of 3,4‐dihydroxy‐2‐butanone‐4‐phosphate from Ru5P at a rate of 199 nmol min?1 mg?1 with Km value of 116 μM at 37°C. Further, we have determined the crystal structures of Salmonella DHBPS in complex with sulfate, Ru5P and sulfate‐zinc ion at a resolution of 2.80, 2.52, and 1.86 Å, respectively. Analysis of these crystal structures reveals that the acidic loop (residues 34–39) responsible for the acid‐base catalysis is disordered in the absence of substrate or metal ion at the active site. Upon binding either substrate or sulfate and metal ions, the acidic loop becomes stabilized, adopts a closed conformation and interacts with the substrate. Our structure for the first time reveals that binding of substrate Ru5P alone is sufficient for the stabilization of the acidic active site loop into a closed conformation. In addition, the Glu38 residue from the acidic active site loop undergoes a conformational change upon Ru5P binding, which helps in positioning the second metal ion that stabilizes the Ru5P and the reaction intermediates. This is the first structural report of DHBPS in complex with either substrate or metal ion from any eubacteria. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
The crystal structure of bovine α-chymotrypsin (α-CHT) in complex with the bovine basic pancreatic trypsin inhibitor (BPTI) has been solved and refined at 2.8 Å resolution (R-factor=0.18). The proteinase:inhibitor complex forms a compact dimer (two α-CHT and two BPTI molecules), which may be stabilized by surface-bound sulphate ions, in the crystalline state. Each BPTI molecule, at opposite ends, is contacting both proteinase molecules in the dimer, through the reactive site loop and through residues next to the inhibitor's C-terminal region. Specific recognition between α-CHT and BPTI occurs at the (re)active site interface according to structural rules inferred from the analysis of homologous serine proteinase:inhibitor complexes. Lys15, the P1 residue of BPTI, however, does not occupy the α-CHT S1 specificity pocket, being hydrogen bonded to backbone atoms of the enzyme surface residues Gly216 and Ser217. © 1997 John Wiley & Sons, Ltd.  相似文献   

5.
Hua Li  Gerwald Jogl 《Proteins》2013,81(3):538-543
Decaprenylphosphoryl‐β‐D ‐ribose 2'‐epimerase (DprE1) is an essential enzyme in the biosynthesis of cell wall components and a target for development of anti‐tuberculosis drugs. We determined the crystal structure of a truncated form of DprE1 from Mycobacterium smegmatis in two crystal forms to up to 2.35 Å resolution. The structure extends from residue 75 to the C‐terminus and shares homology with FAD‐dependent oxidoreductases of the vanillyl‐alcohol oxidase family including the DprE1 homologue from M. tuberculosis. The M. smegmatis DprE1 structure reported here provides further insights into the active site geometry of this tuberculosis drug target. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

6.
Aldehyde dehydrogenases (ALDHs) couple the oxidation of aldehydes to the reduction of NAD(P)+. These enzymes have gained importance as they have been related to the detoxification of aldehydes generated in several diseases involving oxidative stress. It has been determined that tetrameric ALDHs work only with two of their four active sites (half‐of‐the‐sites reactivity), but the mechanistic reason for this feature remains unknown. In this study, tetrameric human aldehyde dehydrogenase class 1A1 (ALDH1A1) was dimerized to study the correlation of the oligomeric structure with the presence of half‐of‐the‐sites reactivity. Stable dimers from ALDH1A1 were generated by combining the mutation of two residues of the dimer–dimer interface in the tetramer (previously shown to render a low‐active and unstable enzyme) and the fusion of green fluorescent protein (GFP) in the C‐terminus of the mutant. Some kinetic properties of the GFP‐fusion mutant resembled those of human aldehyde dehydrogenase class 3A1, a native dimer, in that the fusion dimer did not show burst in the generation of nicotinamide adenine dinucleotide (NADH) and was less sensitive to the action of specific modulators. The presence of primary isotope effect indicated that the rate‐limiting step changed from NADH release to hydride transfer. The mutant showed higher activity with malondialdehyde and acrolein and was more resistant to inactivation by acrolein compared with the wild type. The mutant kinetic profile showed two hyperbolic components when the substrates were varied, suggesting the presence of two active sites with different affinities and catalytic capacities. In conclusion, the ALDH1A1–GFP dimeric mutant exhibits full site reactivity, suggesting that only the tetrameric structure induces the half‐of‐the‐sites reactivity. Proteins 2013; 81:1330–1339. © 2013 Wiley Periodicals, Inc.  相似文献   

7.
Purification and in vitro protein‐folding schemes were developed to produce monodisperse samples of the mature wild‐type HIV‐2 protease (PR2), enabling a comprehensive set of biochemical and biophysical studies to assess the dissociation of the dimeric protease. An E37K substitution in PR2 significantly retards autoproteolytic cleavage during expression. Furthermore, it permits convenient measurement of the dimer dissociation of PR2E37K (elevated Kd ~20 nM) by enzyme kinetics. Differential scanning calorimetry reveals a Tm of 60.5 for PR2 as compared with 65.7°C for HIV‐1 protease (PR1). Consistent with weaker binding of the clinical inhibitor darunavir (DRV) to PR2, the Tm of PR2 increases by 14.8°C in the presence of DRV as compared with 22.4°C for PR1. Dimer interface mutations, such as a T26A substitution in the active site (PR2T26A) or a deletion of the C‐terminal residues 96–99 (PR21–95), drastically increase the Kd (>105‐fold). PR2T26A and PR21–95 consist predominantly of folded monomers, as determined by nuclear magnetic resonance (NMR) and size‐exclusion chromatography coupled with multiangle light scattering and refractive index measurements (SMR), whereas wild‐type PR2 and its active‐site mutant PR2D25N are folded dimers. Addition of twofold excess active‐site inhibitor promotes dimerization of PR2T26A but not of PR21–95, indicating that subunit interactions involving the C‐terminal residues are crucial for dimer formation. Use of SMR and NMR with PR2 facilitates probing for potential inhibitors that restrict protein folding and/or dimerization and, thus, may provide insights for the future design of inhibitors to circumvent drug resistance.  相似文献   

8.
The structure of trypanosomal triosephosphate isomerase (TIM)has been solved at a resolution of 2.1Å in a new crystal form grown at pH 8.8 from PEG6000. In this new crystal form (space group C2, cell dimensions 94.8 Å, 48.3 Å, 131.0 Å, 90.0°, 100.3°, 90.0°), TIM is present in a ligand-free state. The asymmetric unit consists of two TIM subunits. Each of these subunits is part of a dimer which is sitting on a crystallographic twofold axis, such that the crystal packing is formed from two TIM dimers in two distinct environments. The two constituent monomers of a given dimer are, therefore, crystallographically equivalent. In the ligand-free state of TIM in this crystal form, the two types of dimer are very similar in structure, with the flexible loops in the “Open” conformation. For one dimer (termed molecule-1), the flexible loop (loop-6) is involved in crystal contacts. Crystals of this type have been used in soaking experiments with 0.4 M ammonium sulphate (studied at 2.4 Å resolution), and with 40 μM phosphoglycolohydroxamate (studied at 2.5 Å resolution). It is found that transfer to 0.4 M ammonuum sulphate (equal to 80 times the Ki of sulphate for TIM), gives rise to significant sulphate binding at the active site of one dimer (termed molecule-2), and less significant binding at the active site of the other. In neither dimer does sulphate induce a “closed” conformation. In a mother liquor containing 40 μM phosphoglycolohydroxamate (equal to 10 times the Ki of phosphoglycolohydroxamate for TIM), an inhibitor molecule binds at the active site of only that dimer of which the flexible loop is free from crystal contacts (molecule-2). In this dimer, it induces a closed conformation. These three structures are compared and discussed with respect to the mode of binding of ligand in the active site as well as with respect to the conformational changes resulting from ligand binding. © 1993 Wiley-Liss, Inc.  相似文献   

9.
Human α‐amino‐β‐carboxymuconate‐ε‐semialdehyde decarboxylase determines the fate of tryptophan metabolites in the kynurenine pathway by controlling the quinolinate levels for de novo nicotinamide adenine dinucleotide biosynthesis. The unstable nature of its substrate has made gaining insight into its reaction mechanism difficult. Our electron paramagnetic resonance (EPR) spectroscopic study on the Cu‐substituted human enzyme suggests that the native substrate does not directly ligate to the metal ion. Substrate binding did not result in a change of either the hyperfine structure or the super‐hyperfine structure of the EPR spectrum. We also determined the crystal structure of the human enzyme in its native catalytically active state (at 1.99 Å resolution), a substrate analogue‐bound form (2.50 Å resolution), and a selected active site mutant form with one of the putative substrate binding residues altered (2.32 Å resolution). These structures illustrate that each asymmetric unit contains three pairs of dimers. Consistent with the EPR findings, the ligand‐bound complex structure shows that the substrate analogue does not directly coordinate to the metal ion but is bound to the active site by two arginine residues through noncovalent interactions. Proteins 2015; 83:178–187. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
The structural and functional role of conserved residue G86 in HIV‐1 protease (PR) was investigated by NMR and crystallographic analyses of substitution mutations of glycine to alanine and serine (PRG86A and PRG86S). While PRG86S had undetectable catalytic activity, PRG86A exhibited ~6000‐fold lower catalytic activity than PR. 1H‐15N NMR correlation spectra revealed that PRG86A and PRG86S are dimeric, exhibiting dimer dissociation constants (Kd) of ~0.5 and ~3.2 μM, respectively, which are significantly lower than that seen for PR with R87K mutation (Kd > 1 mM). Thus, the G86 mutants, despite being partially dimeric under the assay conditions, are defective in catalyzing substrate hydrolysis. NMR spectra revealed no changes in the chemical shifts even in the presence of excess substrate, indicating very poor binding of the substrate. Both NMR chemical shift data and crystal structures of PRG86A and PRG86S in the presence of active‐site inhibitors indicated high structural similarity to previously described PR/inhibitor complexes, except for specific perturbations within the active site loop and around the mutation site. The crystal structures in the presence of the inhibitor showed that the region around residue 86 was connected to the active site by a conserved network of hydrogen bonds, and the two regions moved further apart in the mutants. Overall, in contrast to the role of R87 in contributing significantly to the dimer stability of PR, G86 is likely to play an important role in maintaining the correct geometry of the active site loop in the PR dimer for substrate binding and hydrolysis. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
Bioinformatics studies have shown that the genomes of trypanosomatid species each encode one SCP2‐thiolase‐like protein (SLP), which is characterized by having the YDCF thiolase sequence fingerprint of the Cβ2‐Cα2 loop. SLPs are only encoded by the genomes of these parasitic protists and not by those of mammals, including human. Deletion of the Trypanosoma brucei SLP gene (TbSLP) increases the doubling time of procyclic T. brucei and causes a 5‐fold reduction of de novo sterol biosynthesis from glucose‐ and acetate‐derived acetyl‐CoA. Fluorescence analyses of EGFP‐tagged TbSLP expressed in the parasite located the TbSLP in the mitochondrion. The crystal structure of TbSLP (refined at 1.75 Å resolution) confirms that TbSLP has the canonical dimeric thiolase fold. In addition, the structures of the TbSLP‐acetoacetyl‐CoA (1.90 Å) and TbSLP‐malonyl‐CoA (2.30 Å) complexes reveal that the two oxyanion holes of the thiolase active site are preserved. TbSLP binds malonyl‐CoA tightly (Kd 90 µM), acetoacetyl‐CoA moderately (Kd 0.9 mM) and acetyl‐CoA and CoA very weakly. TbSLP possesses low malonyl‐CoA decarboxylase activity. Altogether, the data show that TbSLP is a mitochondrial enzyme involved in lipid metabolism. Proteins 2016; 84:1075–1096. © 2016 Wiley Periodicals, Inc.  相似文献   

12.
Human cytosolic aspartyl‐tRNA synthetase (DRS) catalyzes the attachment of the amino acid aspartic acid to its cognate tRNA and it is a component of the multi‐tRNA synthetase complex (MSC) which has been known to be involved in unexpected signaling pathways. Here, we report the crystal structure of DRS at a resolution of 2.25 Å. DRS is a homodimer with a dimer interface of 3750.5 Å2 which comprises 16.6% of the monomeric surface area. Our structure reveals the C‐terminal end of the N‐helix which is considered as a unique addition in DRS, and its conformation further supports the switching model of the N‐helix for the transfer of tRNAAsp to elongation factor 1α. From our analyses of the crystal structure and post‐translational modification of DRS, we suggest that the phosphorylation of Ser146 provokes the separation of DRS from the MSC and provides the binding site for an interaction partner with unforeseen functions.Proteins 2013; 81:1840–1846. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Snake venom serine proteinases (SVSPs) are hemostatically active toxins that perturb the maintenance and regulation of both the blood coagulation cascade and fibrinolytic feedback system at specific points, and hence, are widely used as tools in pharmacological and clinical diagnosis. The crystal structure of a thrombin‐like enzyme (TLE) from Bothrops jararacussu venom (Jararacussin‐I) was determined at 2.48 Å resolution. This is the first crystal structure of a TLE and allows structural comparisons with both the Agkistrodon contortrix contortrix Protein C Activator and the Trimeresurus stejnegeri plasminogen activator. Despite the highly conserved overall fold, significant differences in the amino acid compositions and three‐dimensional conformations of the loops surrounding the active site significantly alter the molecular topography and charge distribution profile of the catalytic interface. In contrast to other SVSPs, the catalytic interface of Jararacussin‐I is highly negatively charged, which contributes to its unique macromolecular selectivity.  相似文献   

14.
Aminoglycosides were one of the first classes of broad‐spectrum antibacterial drugs clinically used to effectively combat infections. The rise of resistance to these drugs, mediated by enzymatic modification, has since compromised their utility as a treatment option, prompting intensive research into the molecular function of resistance enzymes. Here, we report the crystal structure of aminoglycoside nucleotidyltransferase ANT(4′)‐IIb in apo and tobramycin‐bound forms at a resolution of 1.6 and 2.15 Å, respectively. ANT(4′)‐IIb was discovered in the opportunistic pathogen Pseudomonas aeruginosa and conferred resistance to amikacin and tobramycin. Analysis of the ANT(4′)‐IIb structures revealed a two‐domain organization featuring a mixed β‐sheet and an α‐helical bundle. ANT(4′)‐IIb monomers form a dimer required for its enzymatic activity, as coordination of the aminoglycoside substrate relies on residues contributed by both monomers. Despite harbouring appreciable primary sequence diversity compared to previously characterized homologues, the ANT(4′)‐IIb structure demonstrates a surprising level of structural conservation highlighting the high plasticity of this general protein fold. Site‐directed mutagenesis of active site residues and kinetic analysis provides support for a catalytic mechanism similar to those of other nucleotidyltransferases. Using the molecular insights provided into this ANT(4′)‐IIb‐represented enzymatic group, we provide a hypothesis for the potential evolutionary origin of these aminoglycoside resistance determinants.  相似文献   

15.
Kijanimicin is an antitumor antibiotic isolated from Actinomadura kijaniata. It is composed of three distinct moieties: a pentacyclic core, a monosaccharide referred to as d ‐kijanose, and a tetrasaccharide chain composed of l ‐digitoxose units. d ‐Kijanose is a highly unusual nitro‐containing tetradeoxysugar, which requires at least ten enzymes for its production. Here we describe a structural analysis of one of these enzymes, namely KijD1, which functions as a C‐3′‐methyltransferase using S‐adenosylmethionine as its cofactor. For this investigation, two ternary complexes of KijD1, determined in the presence of S‐adenosylhomocysteine (SAH) and dTDP or SAH and dTDP‐3‐amino‐2,3,6‐trideoxy‐4‐keto‐3‐methyl‐d ‐glucose, were solved to 1.7 or 1.6 Å resolution, respectively. Unexpectedly, these structures, as well as additional biochemical analyses, demonstrated that the quaternary structure of KijD1 is a dimer. Indeed, this is in sharp contrast to that previously observed for the sugar C‐3′‐methyltransferase isolated from Micromonospora chalcea. By the judicious use of site‐directed mutagenesis, it was possible to convert the dimeric form of KijD1 into a monomeric version. The quaternary structure of KijD1 could not have been deduced based solely on bioinformatics approaches, and thus this investigation highlights the continuing need for experimental validation.  相似文献   

16.
Members of the Drosophila behavior/human splicing (DBHS) protein family have been characterized in the vertebrates Homo sapiens and Mus musculus, and the invertebrates Drosophila melanogaster and Chironomus tentans. Collectively, both vertebrate and invertebrate DBHS proteins function throughout gene regulation, largely but not always, within the nucleus. In this study, we report a structural and bioinformatic analysis of the DBHS protein family to guide future studies into DBHS protein function. To explore the structural plasticity of the family, we describe the 2.4 Å crystal structure of Caenorhabditis elegans non‐POU domain‐containing octamer‐binding protein 1 (NONO‐1). The structure is dimeric, with a domain arrangement consistent with mammalian DBHS proteins. Comparison with the DBHS structures available from H. sapiens reveals that there is inherent domain flexibility within the homologous DBHS region. Mapping amino acid similarity within the family to the NONO‐1 dimer highlights the dimer interface, coiled‐coil oligomerization motif, and putative RNA binding surfaces. Surprisingly, the interior surface of RNA recognition motif 2 (RRM2) that faces a large internal void is highly variable, but the external β2–β3 loops of RRM2 show remarkable preservation. Overall, the DBHS region is under strong purifying selection, whereas the sequences N‐ and C‐terminal to the DBHS region are less constrained. The findings described in this study provide a molecular basis for further investigation into the mechanistic function of the DBHS protein family in biology.  相似文献   

17.
Nitric oxide reductase (NOR) catalyzes the generation of nitrous oxide (N2O) via the reductive coupling of two nitric oxide (NO) molecules at a heme/non‐heme Fe center. We report herein on the structures of the reduced and ligand‐bound forms of cytochrome c‐dependent NOR (cNOR) from Pseudomonas aeruginosa at a resolution of 2.3–2.7 Å, to elucidate structure‐function relationships in NOR, and compare them to those of cytochrome c oxidase (CCO) that is evolutionarily related to NOR. Comprehensive crystallographic refinement of the CO‐bound form of cNOR suggested that a total of four atoms can be accommodated at the binuclear center. Consistent with this, binding of bulky acetaldoxime (CH3‐CH=N‐OH) to the binuclear center of cNOR was confirmed by the structural analysis. Active site reduction and ligand binding in cNOR induced only ~0.5 Å increase in the heme/non‐heme Fe distance, but no significant structural change in the protein. The highly localized structural change is consistent with the lack of proton‐pumping activity in cNOR, because redox‐coupled conformational changes are thought to be crucial for proton pumping in CCO. It also permits the rapid decomposition of cytotoxic NO in denitrification. In addition, the shorter heme/non‐heme Fe distance even in the bulky ligand‐bound form of cNOR (~4.5 Å) than the heme/Cu distance in CCO (~5 Å) suggests the ability of NOR to maintain two NO molecules within a short distance in the confined space of the active site, thereby facilitating N‐N coupling to produce a hyponitrite intermediate for the generation of N2O. Proteins 2014; 82:1258–1271. © 2013 Wiley Periodicals, Inc.  相似文献   

18.
19.
Lactobacillus casei L ‐lactate dehydrogenase (LCLDH) is activated through the homotropic and heterotropic activation effects of pyruvate and fructose 1,6‐bisphosphate (FBP), respectively, and exhibits unusually high pH‐dependence in the allosteric effects of these ligands. The active (R) and inactive (T) state structures of unliganded LCLDH were determined at 2.5 and 2.6 Å resolution, respectively. In the catalytic site, the structural rearrangements are concerned mostly in switching of the orientation of Arg171 through the flexible intersubunit contact at the Q‐axis subunit interface. The distorted orientation of Arg171 in the T state is stabilized by a unique intra‐helix salt bridge between Arg171 and Glu178, which is in striking contrast to the multiple intersubunit salt bridges in Lactobacillus pentosus nonallosteric L ‐lactate dehydrogenase. In the backbone structure, major structural rearrangements of LCLDH are focused in two mobile regions of the catalytic domain. The two regions form an intersubunit linkage through contact at the P‐axis subunit interface involving Arg185, replacement of which with Gln severely decreases the homotropic and hetertropic activation effects on the enzyme. These two regions form another intersubunit linkage in the Q‐axis related dimer through the rigid NAD‐binding domain, and thus constitute a pivotal frame of the intersubunit linkage for the allosteric motion, which is coupled with the concerted structural change of the four subunits in a tetramer, and of the binding sites for pyruvate and FBP. The unique intersubunit salt bridges, which are observed only in the R state structure, are likely involved in the pH‐dependent allosteric equilibrium. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
Size exclusion chromatographic analyses showed that Ca2+‐free VILIP‐1 contained both monomeric and dimeric forms, while no appreciable dimerization was noted with Ca2+‐free VILIP‐3. Swapping of EF‐hands 3 and 4 of VILIP‐1 with those of VILIP‐3 caused the inability of the resulting chimeric protein to form dimeric protein. Nonreducing SDS‐PAGE analyses revealed that most of the dimeric VILIP‐1 was noncovalently bound together. Reduced glutathione (GSH)/oxidized glutathione (GSSG) treatment notably enhanced the formation of disulfide‐linked VILIP‐1 dimer, while Ca2+ and Mg2+ enhanced disulfide dimerization of VILIP‐1 marginally in the presence of thiol compounds. Cys‐187 at the C‐terminus of VILIP‐1 contributed greatly to form S‐S‐crosslinked dimer as revealed by mutagenesis studies. The ability of GSH/GSSG‐treated VILIP‐1 to activate guanylyl cyclase B was reduced by substituting Cys‐187 with Ala. Together with disulfide dimer of VILIP‐1 detected in rat brain extracts, our data may imply the functional contribution of disulfide dimer to the interaction of VILIP‐1 with its physiological target(s). Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  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