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 546 毫秒
1.
Apple four in human blood coagulation factor XI mediates dimer formation.   总被引:5,自引:0,他引:5  
Human blood coagulation factor XI is a dimer composed of two identical subunits. Each subunit contains four apple domains as tandem repeats followed by a serine protease region. A disulfide bridge between Cys321 of each fourth apple domain links the subunits together. The role of Cys321 in the dimerization of factor XI was examined by mutagenesis followed by expression of its cDNA in baby hamster kidney cells. The recombinant proteins were then purified from the tissue culture medium and shown to have full biological activity. Normal recombinant factor XI was secreted as a dimer as determined by SDS-PAGE, while recombinant factor XI-Cys321 Ser migrated as a monomer under these conditions. Gel filtration studies, however, revealed that each protein existed as a dimer under native conditions, indicating that the disulfide bond between Cys321 of each factor XI monomer was not necessary for dimer formation. The fourth apple domain (apple4) of factor XI was then introduced into tissue plasminogen activator (tPA) to investigate its role in the dimerization of other polypeptide chains. The fusion protein, containing apple4 (apple4-tPA), formed dimers as detected by SDS-PAGE and gel filtration. Furthermore, dimerization was specific to apple4, while apple3 had no effect on dimerization. These data further indicated that the apple4 domain of factor XI mediates dimerization of the two subunits and the interchain disulfide bond involving Cys321 was not essential for dimer formation.  相似文献   

2.
Coagulation factor XI (FXI) is a covalent homodimer consisting of two identical subunits of 80 kDa linked by a disulfide bond formed by Cys-321 within the Apple 4 domain of each subunit. Because FXI(C321S) is a noncovalent dimer, residues within the interface between the two subunits must mediate its homodimeric structure. The crystal structure of FXI demonstrates formation of salt bridges between Lys-331 of one subunit and Glu-287 of the other subunit and hydrophobic interactions at the interface of the Apple 4 domains involving Ile-290, Leu-284, and Tyr-329. FXI(C321S), FXI(C321S,K331A), FXI(C321S,E287A), FXI(C321S,I290A), FXI(C321S,Y329A), FXI(C321S,L284A), FXI(C321S,K331R), and FXI(C321S,H343A) were expressed in HEK293 cells and characterized using size exclusion chromatography, analytical ultracentrifugation, electron microscopy, and functional assays. Whereas FXI(C321S) and FXI(C321S,H343A) existed in monomer/dimer equilibrium (K(d) approximately 40 nm), all other mutants were predominantly monomers with impaired dimer formation by analytical ultracentrifugation (K(d)=3-38 microm). When converted to the active enzyme, FXIa, all the monomeric mutants activated FIX similarly to wild-type dimeric FXIa. In contrast, these monomeric mutants could not be activated efficiently by FXIIa, thrombin, or autoactivation in the presence of dextran sulfate. We conclude that salt bridges formed between Lys-331 of one subunit and Glu-287 of the other together with hydrophobic interactions at the interface, involving residues Ile-290, Leu-284, and Tyr-329, are essential for homodimer formation. The dimeric structure of FXI is essential for normal proteolytic activation of FXI by FXIIa, thrombin, or FXIa either in solution or on an anionic surface but not for FIX activation by FXIa in solution.  相似文献   

3.
Factor XI is a plasma glycoprotein that participates in the blood coagulation cascade. Of the 19 disulfide bonds present in each of the subunits of the human protein, 16 were determined by amino acid sequence analysis of peptide fragments produced by chemical and enzymatic digestion. Four apple domains of 90 or 91 amino acids were identified in the tandem repeats present in the amino-terminal portion of each subunit of factor XI. The disulfide bonds in the carboxyl-terminal portion of the molecule were similar to those in the catalytic region of other serine proteases. The two identical subunits of factor XI were connected by a single disulfide bond at Cys321 linking each of the fourth apple domains while each of the Cys residues at position 11 in the first apple domains forms a disulfide bond with another Cys residue.  相似文献   

4.
Factor XI binds to activated platelets where it is efficiently activated by thrombin. The factor XI receptor is the platelet membrane glycoprotein (GP) Ib-IX-V complex (Baglia, F. A., Badellino, K. O., Li, C. Q., Lopez, J. A., and Walsh, P. N. (2002) J. Biol. Chem. 277, 1662-1668), a significant fraction of which exists within lipid rafts on stimulated platelets (Shrimpton, C. N., Borthakur, G., Larrucea, S., Cruz, M. A., Dong, J. F., and Lopez, J. A. (2002) J. Exp. Med. 196, 1057-1066). Lipid rafts are membrane microdomains enriched in cholesterol and sphingolipids implicated in localizing membrane ligands and in cellular signaling. We now show that factor XI was localized to lipid rafts in activated platelets ( approximately 8% of total bound) but not in resting platelets. Optimal binding of factor XI to membrane rafts required prothrombin (and Ca2+) or high molecular weight kininogen (and Zn2+), which are required for factor XI binding to platelets. An antibody to GPIb (SZ-2) that disrupts factor XI binding to the GPIb-IX-V complex also disrupted factor XI-raft association. The isolated recombinant Apple 3 domain of factor XI, which mediates factor XI binding to platelets, also completely displaces factor XI from membrane rafts. To investigate the physiological relevance of the factor XI-raft association, the structural integrity of lipid rafts was disrupted by cholesterol depletion utilizing methyl-beta-cyclodextrin. Cholesterol depletion completely prevented FXI binding to lipid rafts, and initial rates of factor XI activation by thrombin on activated platelets were inhibited >85%. We conclude that factor XI is localized to GPIb in membrane rafts and that this association is important for promoting the activation of factor XI by thrombin on the platelet surface.  相似文献   

5.
The Apple 4 (A4) domain of human plasma factor XI (FXI) was used to investigate the process of FXI noncovalent dimer formation. Recombinant 6-histidine-tagged A4 domain proteins were prepared utilizing a bacterial expression system. Purification was accomplished under denaturing conditions, followed by a refolding protocol to facilitate correct disulfide bond formation. Analysis of the A4 domain (C321S mutant) by size exclusion chromatography indicated the presence of a slowly equilibrating reversible monomer-dimer equilibrium. The elution profiles reveal highly symmetrical peaks for both dimeric and monomeric species with elution times that were highly reproducible for varying amounts of both the dimeric and monomeric species. The monomer-dimer equilibrium was found to be dependent upon changes in both pH and salt concentration. Under conditions approximating physiologic salt concentration and pH (20 mm HEPES, 100 mm NaCl, and 1 mm EDTA, pH 7.4), it was determined that the monomer-dimer equilibrium was characterized by a dissociation constant (K(D)) value of 229 +/- 26 nm with a calculated Delta G value of 9.1 kcal/mol. This report identifies electrostatic contributions and the presence of a hydrophobic component that mediate interactions at the A4 domain interface. The rate of dissociation for the recombinant A4 domain C321S mutant was examined by monitoring the increase in 4,4'-dianilino-1,1'-binaphthyl-5,5'-disulfonic acid dipotassium salt fluorescence under dissociating conditions, giving a value for a dissociation rate constant (k(off)) of 4.3 x 10(-3) s(-1).  相似文献   

6.
The blood coagulation protein factor XI (FXI) consists of a pair of disulfide-linked chains each containing four apple domains and a catalytic domain. The apple 4 domain (A4; F272-E362) mediates non-covalent homodimer formation even when the cysteine involved in an intersubunit disulfide is mutated to serine (C321S). To understand the role of non-covalent interactions stabilizing the FXI dimer, equilibrium unfolding of wild-type A4 and its C321S variant was monitored by circular dichroism, intrinsic tyrosine fluorescence and dynamic light scattering measurements as a function of guanidine hydrochloride concentration. Global analysis of the unimolecular unfolding transition of wild-type A4 revealed a partially unfolded equilibrium intermediate at low to moderate denaturant concentrations. The optically detected equilibrium of C321S A4 also fits best to a three-state model in which the native dimer unfolds via a monomeric intermediate state. Dimer dissociation is characterized by a dissociation constant, K(d), of approximately 90 nM (in terms of monomer), which is in agreement with the dissociation constant measured independently using fluorescence anisotropy. The results imply that FXI folding occurs via a monomeric equilibrium intermediate. This observation sheds light on the effect of certain naturally occurring mutations, such as F283L, which lead to intracellular accumulation of non-native forms of FXI. To investigate the structural and energetic consequences of the F283L mutation, which perturbs a cluster of aromatic side-chains within the core of the A4 monomer, it was introduced into the dissociable dimer, C321S A4. NMR chemical shift analysis confirmed that the mutant can assume a native-like dimeric structure. However, equilibrium unfolding measurements show that the mutation causes a fourfold increase in the K(d) value for dissociation of the native dimer and a 1 kcal/mol stabilization of the monomer, resulting in a highly populated intermediate. Since the F283 side-chain does not directly participate in the dimer interface, we propose that the F283L mutation leads to increased dimer dissociation by stabilizing a monomeric state with altered side-chain packing that is unfavorable for homodimer formation.  相似文献   

7.
Isocitrate dehydrogenase kinase/phosphatase (IDHK/P) is a homodimeric enzyme which controls the oxidative metabolism of Escherichia coli, and exibits a high intrinsic ATPase activity. When subjected to electrophoresis under nonreducing conditions, the purified enzyme migrates partially as a dimer. The proportion of the dimer over the monomer is greatly increased by treatment with cupric 1,10 phenanthrolinate or 5,5'-dithio-bis(2-nitrobenzoic acid), and fully reversed by dithiothreitol, indicating that covalent dimerization is produced by a disulfide bond. To identify the residue(s) involved in this intermolecular disulfide-bond, each of the eight cysteines of the enzyme was individually mutated into a serine. It was found that, under nonreducing conditions, the electrophoretic patterns of all corresponding mutants are identical to that of the wild-type, except for the Cys67-->Ser which migrates exclusively as a monomer and for the Cys108-->Ser which migrates preferentially as a dimer. Furthermore, in contrast to the wild-type enzyme and all the other mutants, the Cys67-->Ser mutant still migrates as a monomer after treatment with cupric 1,10 phenanthrolinate. This result indicates that the intermolecular disulfide bond involves only Cys67 in each IDHK/P wild-type monomer. This was further supported by mass spectrum analysis of the tryptic peptides derived from either the cupric 1,10 phenanthrolinate-treated wild-type enzyme or the native Cys108-->Ser mutant, which show that they both contain a Cys67-Cys67 disulfide bond. Moreover, both the cupric 1,10 phenanthrolinate-treated wild-type enzyme and the native Cys108-->Ser mutant contain another disulfide bond between Cys356 and Cys480. Previous results have shown that this additional Cys356-Cys480 disulfide bond is intramolecular [Oudot, C., Jault, J.-M., Jaquinod, M., Negre, D., Prost, J.-F., Cozzone, A.J. & Cortay, J.-C. (1998) Eur. J. Biochem. 258, 579-585].  相似文献   

8.
Activated coagulation factor XI (factor XIa) proteolytically cleaves its substrate, factor IX, in an interaction requiring the factor XI A3 domain (Sun, Y., and Gailani, D. (1996) J. Biol. Chem. 271, 29023-29028). To identify key amino acids involved in factor IX activation, recombinant factor XIa proteins containing alanine substitutions for wild-type sequence were expressed in 293 fibroblasts and tested in a plasma clotting assay. Substitutions for Ile(183)-Val(191) and Ser(195)-Ile(197) at the N terminus and for Ser(258)-Ser(264) at the C terminus of the A3 domain markedly decreased factor XI coagulant activity. The plasma protease prekallikrein is structurally homologous to factor XI, but activated factor IX poorly. A chimeric factor XIa molecule with the A3 domain replaced with A3 from prekallikrein (FXI/PKA3) activated factor IX with a K(m) 35-fold greater than that of wild-type factor XI. FXI/PKA3 was used as a template for a series of proteins in which prekallikrein A3 sequence was replaced with factor XI sequence to restore factor IX activation. Clotting and kinetics studies using these chimeras confirmed the results obtained with alanine mutants. Amino acids between Ile(183) and Val(191) are necessary for proper factor IX activation, but additional sequence between Ser(195) and Ile(197) or between Phe(260) and Ser(265) is required for complete restoration of activation.  相似文献   

9.
Miller TN  Sinha D  Baird TR  Walsh PN 《Biochemistry》2007,46(50):14450-14460
The zymogen, factor XI, and the enzyme, factor XIa, interact specifically with functional receptors on the surface of activated platelets. These studies were initiated to identify the molecular subdomain within factor XIa that binds to activated platelets. Both factor XIa (Ki approximately 1.4 nM) and a chimeric factor XIa containing the Apple 3 domain of prekallikrein (Ki approximately 2.7 nM) competed with [125I]factor XIa for binding sites on activated platelets, suggesting that the factor XIa binding site for platelets is not located in the Apple 3 domain which mediates factor XI binding to platelets. The recombinant catalytic domain (Ile370-Val607) inhibited the binding of [125I]factor XIa to the platelets (Ki approximately 3.5 nM), whereas the recombinant factor XI heavy chain did not, demonstrating that the platelet binding site is located in the light chain of factor XIa. A conformationally constrained cyclic peptide (Cys527-Cys542) containing a high-affinity (KD approximately 86 nM) heparin-binding site within the catalytic domain of factor XIa also displaced [125I]factor XIa from the surface of activated platelets (Ki approximately 5.8 nM), whereas a scrambled peptide of identical composition was without effect, suggesting that the binding site in factor XIa that interacts with the platelet surface resides in the catalytic domain near the heparin binding site of factor XIa. These data support the conclusion that a conformational transition accompanies conversion of factor XI to factor XIa that conceals the Apple 3 domain factor XI (zymogen) platelet binding site and exposes the factor XIa (enzyme) platelet binding site within the catalytic domain possibly comprising residues Cys527-Cys542.  相似文献   

10.
The melanocortin 1 receptor (MC1R) is a dimeric G protein-coupled receptor expressed in melanocytes, where it regulates the amount and type of melanins produced and determines the tanning response to ultraviolet radiation. We have studied the mechanisms of MC1R dimerization. Normal dimerization of a deleted mutant lacking the seventh transmembrane fragment and the C-terminal cytosolic extension excluded coiled-coil interactions as the basis of dimerization. Conversely, the electrophoretic pattern of wild type receptor and several Cys → Ala mutants showed that four disulfide bonds are established between the monomers. Disruption of any of these bonds abolished MC1R function, but only the one involving Cys35 was essential for traffic to the plasma membrane. A quadruple Cys35-267-273-275Ala mutant migrating as a monomer in SDS-PAGE in the absence of reducing agents was able to dimerize with WT, suggesting that in addition to disulfide bond formation, dimerization involves non-covalent interactions, likely of domain swap type.  相似文献   

11.
Chicken cystatin (cC) mutant I66Q is located in the hydrophobic core of the protein and increases the propensity for amyloid formation. Here, we demonstrate that under physiological conditions, the replacement of Ile with the Gln in the I66Q mutant increases the susceptibility for the disulfide bond Cys71–Cys81 to be reduced when compared to the wild type (WT) cC. Molecular dynamics (MD) simulations under conditions favoring cC amyloid fibril formation are in agreement with the experimental results. MD simulations were also performed to investigate the impact of disrupting the Cys71–Cys81 disulfide bond on the conformational stability of cC at the atomic level, and highlighted major disruption to the cC appendant structure. Domain swapping and extensive unfolding has been proposed as one of the possible mechanisms initiating amyloid fibril formation by cystatin. Our in silico studies suggest that disulfide bond formation between residues Cys95 and Cys115 is necessary to maintain conformational stability of the I66Q mutant following breakage of the Cys71–Cys81 disulfide bridge. Subsequent breakage of disulfide bond Cys95–Cys115 resulted in large structural destabilization of the I66Q mutant, which increased the α–β interface distance and expanded the hydrophobic core. These experimental and computational studies provide molecular-level insight into the relationship between disulfide bond formation and progressive unfolding of amyloidogenic cC mutant I66Q.

An animated Interactive 3D Complement (I3DC) is available in Proteopedia at http://proteopedia.org/w/Journal:JBSD:23  相似文献   

12.
In vivo formation and stability of engineered disulfide bonds in subtilisin   总被引:9,自引:0,他引:9  
Computer modeling suggested that a disulfide bond could be built into Bacillus amyloliquefaciens subtilisin between positions 22 (wild-type, Thr) and 87 (Ser) or between positions 24 (Ser) and 87 (Ser). Single cysteines were introduced into this cysteine-free protease at positions 22, 24, or 87 by site-directed mutagenesis of the cloned subtilisin gene. The corresponding double-cysteine mutants were constructed, and recombinant plasmids were expressed in Bacillus subtilis. Double-cysteine mutant enzymes were secreted as efficiently as wild-type, and disulfide bonds were formed quantitatively in vivo. These disulfide bonds were introduced approximately 24 A away from the catalytic site and had no detectable effect on either the specific activities or the pH optima of the mutant enzymes. The equilibrium constants for the reduction of the mutant disulfide bonds by dithiothreitol were determined to be 82 +/- 22 and 20 +/- 5 for Cys22/Cys87 and Cys24/Cys87, respectively. Studies of autoproteolytic inactivation of wild-type subtilisin support a relationship between autolytic stability and conformational stability of the protein. The stabilities of Cys24/Cys87 and wild-type enzymes to autolysis were essentially the same; however, Cys22/Cys87 was actually less stable to autolysis. Reduction of the disulfide cross-bridge lowered the autolytic stability of both double-cysteine mutants relative to their disulfide forms. This correlates with a lowered autolytic stability for the Cys22 and Cys87 single-cysteine mutants, and the fact that an intramolecular hydrogen bond between the hydroxyl groups of Thr22 and Ser87 is likely to be disrupted in the Cys22 and Cys87 single-cysteine mutant proteins.  相似文献   

13.
A set of wild-type and mutant human, woodchuck, and duck hepatitis viral core proteins have been prepared and used to study the free thiol groups and the disulfide bonding pattern present within the core particle. Human (HBcAg) and woodchuck (WHcAg) core proteins contain 4 cysteine residues, whereas duck (DHcAg) core protein contains a single cysteine residue. Each of the cysteines of HBcAg has been eliminated, either singly or in combinations, by a two-step mutagenesis procedure. All of the proteins were shown to have very similar physical and immunochemical properties. All assemble into essentially identical core particle structures. Therefore disulfide bonds are not essential for core particle formation. No intra-chain disulfide bonds occur. Cys107 is a free thiol buried within the particle structure, whereas Cys48 is present partly as a free sulfhydryl which is exposed at the surface of the particle. Cys61 is always and Cys48 is partly involved in interchain disulfide bonds with the identical residues of another monomer, whereas Cys183 is always involved in a disulfide bond with the Cys183 of a different monomer. WHcAg has the same pattern of bonding, whereas DHcAg lacks any disulfide bonds, and the single free sulfhydryl, Cys153 which is equivalent to Cys107 of HBcAg, is buried.  相似文献   

14.
Factor XI (FXI) binds specifically and reversibly to high affinity sites on the surface of stimulated platelets (Kd app of approximately 10 nm; Bmax of approximately 1,500 sites/platelet) utilizing residues exposed on the Apple 3 domain in the presence of high molecular weight kininogen and Zn2+ or prothrombin and Ca2+. Because the FXI receptor in the platelet membrane is contained within the glycoprotein Ibalpha subunit of the glycoprotein Ib-IX-V complex (Baglia, F. A., Badellino, K. O., Li, C. Q., Lopez, J. A., and Walsh, P. N. (2002) J. Biol. Chem. 277, 1662-1668), we utilized mocarhagin, a cobra venom metalloproteinase, to generate a fragment (His1-Glu282) of glycoprotein Ibalpha that contains the leucine-rich repeats of the NH2-terminal globular domain and excludes the macroglycopeptide portion of glycocalicin, the soluble extracytoplasmic portion of glycoprotein Ibalpha. This fragment was able to compete with FXI for binding to activated platelets (Ki of 3.125 +/- 0.25 nm) with a potency similar to that of intact glycocalicin (Ki of 3.72 +/- 0.30 nm). However, a synthetic glycoprotein Ibalpha peptide, Asp269-Asp287, containing a thrombin binding site had no effect on the binding of FXI to activated platelets. Moreover, the binding of 125I-labeled thrombin to glycocalicin was unaffected by the presence of FXI at concentrations up to 10(-5) m. The von Willebrand factor A1 domain, which binds the leucine-rich repeats, inhibited the binding of FXI to activated platelets. Thus, we examined the effect of synthetic peptides of each of the seven leucine-rich repeats on the binding of 125I-FXI to activated platelets. All leucine-rich repeat (LRR) peptides derived from glycoprotein Ibalpha were able to inhibit FXI binding to activated platelets in the following order of decreasing potency: LRR7, LRR1, LRR4, LRR5, LRR6, LRR3, and LRR2. However, the leucine-rich repeat synthetic peptides derived from glycoprotein Ibbeta and Toll protein had no effect. We conclude that FXI binds to glycoprotein Ibalpha at sites comprising the leucine-rich repeat sequences within the NH2-terminal globular domain that are separate and distinct from the thrombin-binding site.  相似文献   

15.
R Procyk  B Blomb?ck 《Biochemistry》1990,29(6):1501-1507
Fibrinogen contains 29 disulfide bonds. Limited reduction in buffers containing calcium led to cleavage of three of them: the two A alpha 442Cys-A alpha 472Cys intrapeptide disulfide bonds and the symmetrical A alpha 28Cys-A alpha 28Cys bond. The limited reduction did not affect clotting by thrombin. However, a prolongation of the thrombin clotting time occurred when the limited reduction took place in the absence of calcium. The bonds reduced under this condition included the three already mentioned and also the two gamma 326Cys-gamma 339Cys intrapeptide disulfide bonds located in the C-terminal ends of the gamma-chain. N-Terminal analysis of thrombin-treated samples showed that thrombin cleavage occurred at the normal A alpha 16-A alpha 17 site in fibrinogen that was partially reduced in the presence of calcium. By contrast, thrombin cleaved at the A alpha 19-A alpha 20 site in fibrinogen that was partially reduced in the absence of calcium, rendering the protein unclottable by removing the A alpha 17Gly-18Pro-19Arg peptide. The loss of thrombin clottability may have also come from gamma 326Cys-gamma 339Cys disulfide bond reduction since the structure supported by this bond may be important for the function of the C-terminal polymerization site. In samples of the partially reduced fibrinogen lacking the A alpha 17-19 residues, gel formation occurred through an oligomerization mechanism catalyzed by factor XIII.  相似文献   

16.
To gain insight into the molecular architecture of the cytoplasmic surface of G protein-coupled receptors, we have developed a disulfide cross-linking strategy using the m3 muscarinic receptor as a model system. To facilitate the interpretation of disulfide cross-linking data, we initially generated a mutant m3 muscarinic receptor (referred to as m3'(3C)-Xa) in which most native Cys residues had been deleted or substituted with Ala or Ser (remaining Cys residues Cys-140, Cys-220, and Cys-532) and in which the central portion of the third intracellular loop had been replaced with a factor Xa cleavage site. Radioligand binding and second messenger assays showed that the m3'(3C)-Xa mutant receptor was fully functional. In the next step, pairs of Cys residues were reintroduced into the m3'(3C)-Xa construct, thus generating 10 double Cys mutant receptors. All 10 mutant receptors contained a Cys residue at position 169 at the beginning of the second intracellular loop and a second Cys within the C-terminal portion of the third intracellular loop, at positions 484-493. Radioligand binding studies and phosphatidylinositol assays indicated that all double Cys mutant receptors were properly folded. Membrane lysates prepared from COS-7 cells transfected with the different mutant receptor constructs were incubated with factor Xa protease and the oxidizing agent Cu(II)-(1,10-phenanthroline)3, and the formation of intramolecular disulfide bonds between juxtaposed Cys residues was monitored by using a combined immunoprecipitation/immunoblotting strategy. To our surprise, efficient disulfide cross-linking was observed with 8 of the 10 double Cys mutant receptors studied (Cys-169/Cys-484 to Cys-491), suggesting that the intracellular m3 receptor surface is characterized by pronounced backbone fluctuations. Moreover, [35S]guanosine 5'-3-O-(thio)triphosphate binding assays indicated that the formation of intramolecular disulfide cross-links prevented or strongly inhibited receptor-mediated G protein activation, suggesting that the highly dynamic character of the cytoplasmic receptor surface is a prerequisite for efficient receptor-G protein interactions. This is the first study using a disulfide mapping strategy to examine the three-dimensional structure of a hormone-activated G protein-coupled receptor.  相似文献   

17.
Six rhodopsin mutants containing disulfide cross-links between different cytoplasmic regions were prepared: disulfide bond 1, between Cys65 (interhelical loop I-II) and Cys316 (end of helix VII); disulfide bond 2, between Cys246 (end of helix VI) and Cys312 (end of helix VII); disulfide bond 3, between Cys139 (end of helix III) and Cys248 (end of helix VI); disulfide bond 4, between Cys139 (end of helix III) and Cys250 (end of helix VI); disulfide bond 5, between Cys135 (end of helix III) and Cys250 (end of helix VI); and disulfide bond 6, between Cys245 (end of helix VI) and Cys338 (C-terminus). The effects of local restrictions caused by the cross-links on transducin (G(T)) activation and phosphorylation by rhodopsin kinase (RK) following illumination were studied. Disulfide bond 1 showed little effect on either G(T) activation or phosphorylation by RK, suggesting that the relative motion between interhelical loop I-II and helix VII is not crucial for recognition by G(T) or by RK. In contrast, disulfide bonds 2-5 abolished both G(T) activation and phosphorylation by RK. Disulfide bond 6 resulted in enhanced G(T) activation but abolished phosphorylation by RK, suggesting the structure recognized by G(T) was stabilized in this mutant by cross-linking of the C-terminus to the cytoplasmic end of helix VI. Thus, the consequences of the disulfide cross-links depended on the location of the restriction. In particular, relative motions of helix VI, with respect to both helices III and VII upon light activation, are required for recognition of rhodopsin by both G(T) and RK. Further, the conformational changes in the cytoplasmic face that are necessary for protein-protein interactions need not be cooperative, and may be segmental.  相似文献   

18.
Rat 3-mercaptopyruvate sulfurtransferase (MST) contains three exposed cysteines as follows: a catalytic site cysteine, Cys(247), in the active site and Cys(154) and Cys(263) on the surface of MST. The corresponding cysteine to Cys(263) is conserved in mammalian MSTs, and Cys(154) is a unique cysteine. MST has monomer-dimer equilibrium with the assistance of oxidants and reductants. The monomer to dimer ratio is maintained at approximately 92:8 in 0.2 m potassium phosphate buffer containing no reductants under air-saturated conditions; the dimer might be symmetrical via an intersubunit disulfide bond between Cys(154) and Cys(154) and between Cys(263) and Cys(263), or asymmetrical via an intersubunit disulfide bond between Cys(154) and Cys(263). Escherichia coli reduced thioredoxin (Trx) cleaved the intersubunit disulfide bond to activate MST to 2.3- and 4.9-fold the levels of activation of dithiothreitol (DTT)-treated and DTT-untreated MST, respectively. Rat Trx also activated MST. On the other hand, reduced glutathione did not affect MST activity. E. coli C35S Trx, in which Cys(35) was replaced with Ser, formed some adducts with MST and activated MST after treatment with DTT. Thus, Cys(32) of E. coli Trx reacted with the redox-active cysteines, Cys(154) and Cys(263), by forming an intersubunit disulfide bond and a sulfenyl Cys(247). A consecutively formed disulfide bond between Trx and MST must be cleaved for the activation. E. coli C32S Trx, however, did not activate MST. Reduced Trx turns on a redox switch for the enzymatic activation of MST, which contributes to the maintenance of cellular redox homeostasis.  相似文献   

19.
The production of class A beta-lactamases is a major cause of clinical resistance to beta-lactam antibiotics. Some of class A beta-lactamases are known to have a disulfide bridge. Both narrow spectrum and extended spectrum beta-lactamases of TEM and the SHV enzymes possess a disulfide bond between Cys77 and Cys123, and the enzymes with carbapenem-hydrolyzing activity have a well-conserved disulfide bridge between Cys69 and Cys238. We produced A77C/G123C mutant of the extended-spectrum beta-lactamase Toho-1 in order to introduce a disulfide bond between the cysteine residues at positions 77 and 123. The result of 5,5'-dithiobis-2-nitrobenzoic acid (DTNB) titrations confirmed formation of a new disulfide bridge in the mutant. The results of irreversible heat inactivation and circular dichroism (CD) melting experiments indicated that the disulfide bridge stabilized the enzyme significantly. Though kinetic analysis indicated that the catalytic properties of the mutant were quite similar to those of the wild-type enzyme, E. coli producing this mutant showed drug resistance significantly higher than E. coli producing the wild-type enzyme. We speculate that the stability of the enzymes provided by the disulfide bond may explain the wide distribution of TEM and SHV derivatives and explain how various mutations can cause broadened substrate specificity without loss of stability.  相似文献   

20.
Human lysozyme is made up of 130 amino acid residues and has four disulfide bonds at Cys6-Cys128, Cys30-Cys116, Cys65-Cys81, and Cys77-Cys95. Our previous results using the Saccharomyces cerevisiae secretion system indicate that the individual disulfide bonds of human lysozyme have different functions in the correct in vivo folding and enzymatic activity of the protein (Taniyama, Y., Yamamoto, Y., Nakao, M., Kikuchi, M., and Ikehara, M. (1988) Biochem. Biophys. Res. Commun. 152, 962-967). In this paper, we report the results of experiments that were focused on the roles of Cys65 and Cys81 in the folding of human lysozyme protein in yeast. A mutant protein (C81A), in which Cys81 was replaced with Ala, had almost the same enzymatic activity and conformation as those of the native enzyme. On the other hand, another mutant (C65A), in which Cys65 was replaced with Ala, was not found to fold correctly. These results indicate that Cys81 is not a requisite for both correct folding and activity, whereas Cys65 is indispensable. The mutant protein C81A is seen to contain a new, non-native disulfide bond at Cys65-Cys77. The possible occurrence of disulfide bond interchange during our mapping experiments cannot be ruled out by the experimental techniques presently available, but characterization of other mutant proteins and computer analysis suggest that the intramolecular exchange of disulfide bonds is present in the folding pathway of human lysozyme in vivo.  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