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.
Bovine neurophysin-II is shown to be very susceptible to partial reduction in the absence of urea. Reduction of an average of one disulfide leads to major changes in conformation and disulfide optical activity, manifest in part by pronounced far-uv ellipticity changes, complete loss of the 248-nm ellipticity band, and a shift of the 278-nm ellipticity band to shorter wavelengths with loss of half its intensity; the reduction process generates a mixture of products and appears to be accompanied by disulfide interchange. The circular dichroism data indicate that the disulfide(s) most susceptible to reduction or interchange are either the principal contributors to the 248- and 278-nm ellipticity bands or that the optical activity of other disulfides is dependent on their integrity. Peptides that bind to the hormone-binding site of neurophysin-II protect against reduction. On reoxidation of partially reduced neurophysin-II there is only a partial return of the native circular dichroism spectrum and electrophoretic behavior. The percentage of native protein in samples reoxidized following different degrees of reduction was estimated by comparison of the circular dichroism spectra of these samples with those of the fractionated native and denatured components of monoreduced-reoxidized neurophysin. Under our reoxidation conditions, less than 50% native protein was found in monoreduced-reoxidized neurophysin and less than 10% native protein was found in completely reduced-reoxidized neurophysin. The results are interpreted with qualified reference to a model in which one or more disulfides are "strained" in the native state and in which the native protein is unstable relative to species in which the disulfides are differently paired.  相似文献   

2.
The chloroplastic enzyme NADP-malate dehydrogenase is activated by a reversible thiol/disulfide interchange with reduced thioredoxin. Its target disulfide bridge is considered to be located at the amino terminus. To further substantiate the regulatory role of this disulfide, site-directed mutagenesis has been used to replace each or both of the amino-terminal cysteines of the sorghum leaf NADP-malate dehydrogenase, expressed in Escherichia coli, by serines. A truncation mutant lacking the amino terminus has also been produced. Surprisingly, the mutant proteins still required activation by reduced thioredoxin. However, their activation was almost instantaneous, whereas the native enzyme reached full activity after a 10-20 min preincubation. The 8 1/2 for reduced thioredoxin was decreased 2-fold in the mutants, but their Km values for NADPH and oxaloacetate did not change significantly. The inhibition of activation by NADP and inhibition of activity by thiol-derivatizing agents were also retained. These results are interpreted as an indication that two thioredoxin-dependent reduction steps are involved in NADP-dependent malate dehydrogenase light activation, hence that two disulfides per monomer participate in the process. The overall activation rate would depend on a conformational change following the reduction of the amino-terminal disulfide bridge. The amino terminus also plays a role in the dimerization of the protein.  相似文献   

3.
The sulfhydryl oxidase Ero1 oxidizes protein disulfide isomerase (PDI), which in turn catalyzes disulfide formation in proteins folding in the endoplasmic reticulum (ER). The extent to which other members of the PDI family are oxidized by Ero1 and thus contribute to net disulfide formation in the ER has been an open question. The yeast ER contains four PDI family proteins with at least one potential redox-active cysteine pair. We monitored the direct oxidation of each redox-active site in these proteins by yeast Ero1p in vitro. In this study, we found that the Pdi1p amino-terminal domain was oxidized most rapidly compared with the other oxidoreductase active sites tested, including the Pdi1p carboxyl-terminal domain. This observation is consistent with experiments conducted in yeast cells. In particular, the amino-terminal domain of Pdi1p preferentially formed mixed disulfides with Ero1p in vivo, and we observed synthetic lethality between a temperature-sensitive Ero1p variant and mutant Pdi1p lacking the amino-terminal active-site disulfide. Thus, the amino-terminal domain of yeast Pdi1p is on a preferred pathway for oxidizing the ER thiol pool. Overall, our results provide a rank order for the tendency of yeast ER oxidoreductases to acquire disulfides from Ero1p.  相似文献   

4.
Reduced and denatured hirudin (65 amino acids and 3 disulfides) refolds in vitro to become an active molecule. The folding process adopts a mechanism of "trial and error" without predominant pathways. Throughout the entire folding process, the 6 cysteines were about equally involved in the disulfide shuffling. Among the first 20% of 3-disulfide species accumulated during the early phase of refolding, two-thirds were inactive and were reshuffled in the presence of thiol catalyst to regain correct disulfide pairing. When refolding was performed in the presence of strong denaturant (guanidinium chloride) without thiol catalyst, 8% of the active hirudin was obtained. This figure is close to the probability (6.7%) that would be expected from the random disulfide pairing of a molecule containing 6 sulfhydryl groups.  相似文献   

5.
T Marti  S J R?sselet  K Titani  K A Walsh 《Biochemistry》1987,26(25):8099-8109
In the course of identifying substructural domains within the homooligomeric protein von Willebrand factor [270 kilodaltons (kDa) per polypeptide chain], seven large fragments of 8-90 kDa have been generated by limited proteolysis. A monomeric fragment that binds coagulation factor VIIIc is identified as residues 1-272. A fragment that binds platelet glycoprotein Ib is identified as a homodimer containing two pairs of identical chains, i.e., residues 273-511 and 674-728. Disulfide bonds have been identified by several methods, including direct observation of the phenylthiohydantoin of cystine during Edman degradation of isolated peptides. Among half-cystine residues in the amino-terminal 1365-residue region, 52 have been paired. They place structural constraints on folding possibilities within three structural domains. Additional clusters of disulfide bonds are evident. It has been shown that at least 35 disulfides must form intrachain bridges, specifically the cystines among residues 1-272 and 906-1492. Intersubunit disulfide bonds are partially localized in an interior region (residues 283-695) and a carboxyl-terminal region (residues 1908-2050). Each of these regions appears to be linked to a corresponding region of a neighboring subunit in the network of interconnected chains. The difficulties of pairing all 169 half-cystines (per chain) and of distinguishing intrachain from interchain disulfides are evaluated.  相似文献   

6.
An oxytocin/bovine neurophysin I biosynthetic precursor, [N epsilon-diacetimidyl-30,71, des-His106]pro-OT/BNPI, was synthesized from a synthetic oxytocinyl peptide, 1/2Cys-Tyr-Ile-Gln-Asn-1/2Cys-Pro-Leu-Gly-Gly-Lys-Arg, and native neurophysin by chemical semisynthesis. The semisynthetic precursor contains the entire sequence of the biosynthetic precursor deduced from the complementary DNA structure except for omission of the carboxyl-terminal histidine residue. The covalent structure of the semisynthetic product was verified by amino acid analysis and amino-terminal analysis. Analytical affinity chromatography was employed to evaluate noncovalent binding properties of the precursor. The precursor does not bind significantly to immobilized Met-Tyr-Phe, a hormone binding site ligand. In contrast, the acetimidated precursor binds to immobilized bovine neurophysin II, with a 13-fold higher affinity than does acetimidated neurophysin itself. When a hormonal ligand, [Lys8]vasopressin, was added to the elution buffer at the concentration of 0.1 mM so that a major portion of the immobilized BNPII was liganded, the affinity between the immobilized liganded BNPII and the precursor was enhanced 8-fold and approached the affinity for the liganded (bovine neurophysin I-immobilized BNPII) interaction. The data imply that the precursor can self-associate and that this self-association is closely related to that of liganded neurophysin. The tripeptide affinity matrix data argue that, in the precursor, the ligand binding site of the neurophysin domain is occupied intramolecularly by the hormone domain. The data verify the view that both the self-association surface and hormone binding site are established upon precursor folding. A disulfide stability analysis showed the resistance, to disulfide interchange by dithiothreitol, of semisynthetic precursor but not of neurophysin, as judged by protein association and peptide ligand binding activities, respectively. The results argue that the molecular structure of the precursor is established upon precursor folding and before enzymatic processing that produces mature hormone and neurophysin.  相似文献   

7.
Guo ZY  Shen L  Feng YM 《Biochemistry》2002,41(34):10585-10592
Insulin and insulin-like growth factor 1 (IGF-1) share homologous sequence, similar three-dimensional structure, and weakly overlapping biological activity, but different folding information is stored in their homologous sequences: the sequence of insulin encodes one unique thermodynamically stable three-dimensional structure while that of IGF-1 encodes two disulfide isomers with different three-dimensional structure but similar thermodynamic stability. Their different folding behavior probably resulted from the different energetic state of the intra A-chain/domain disulfide: the intra A-chain disulfide of insulin is a stable bond while that of IGF-1 is a strained bond with high energy. To find out the sequence determinant of the different energetic state of their intra A-chain/domain disulfide, the following experiments were carried out. First, a local chimeric single-chain insulin (PIP) with the A8-A10 residues replaced by the corresponding residues of IGF-1 was prepared. Second, the disulfide stability of two global hybrids of insulin and IGF-1, Ins(A)/IGF-1(B) and Ins(B)/IGF-1(A), was investigated. The local segment swap had no effect on the fidelity of disulfide pairing and the disulfide stability of PIP molecule although the swapped segment is close to the intra A-chain/domain disulfide. In redox buffer which favors the disulfide formation for most proteins, Ins(A)/IGF-1(B) cannot form and maintain its native disulfides just like that of IGF-1, while the disulfides of Ins(B)/IGF-1(A) are stable in the same condition. One major equilibrium intermediate with two disulfides of Ins(A)/IGF-1(B) was purified and characterized. V8 endoproteinase cleavage and circular dichroism analysis suggested that the intra A-chain/domain disulfide was reduced in the intermediate. Our present results suggested that the energetic state of the intra A-chain/domain disulfide of insulin and IGF-1 was not controlled by the A-chain/domain sequence close to this disulfide but was mainly controlled by the sequence of the B-chain/domain.  相似文献   

8.
Cigić B  Dahl SW  Pain RH 《Biochemistry》2000,39(40):12382-12390
The 13.5 kDa N-terminal part of the propeptide remains associated with mature cathepsin C after proteolytic activation and excision of the activation peptide. This residual pro-part, isolated from the recombinant enzyme, folds spontaneously and rapidly to a stable, compact monomer with secondary structure and stable tertiary interactions. Folding and unfolding kinetics of the residual pro-part with intact disulfides are complex, and accumulation of transient intermediates is observed. The cleaved form of the pro-part isolated from natural human cathepsin C also folds, suggesting that the intact form comprises two folding domains. The linkages of the two disulfide bridges have been established as 30-118 and 54-136 for the native enzyme. The native disulfide bonds can be re-formed from the fully reduced and denatured state by oxidative refolding, resulting in a domain that is spectroscopically indistinguishable from the original refolded residual pro-part. Both disulfides are solvent-exposed and can be reduced in the absence of denaturant. The reduced form retains most or all of the native tertiary structure and is only approximately 2 kcal.mol(-1) less stable than the oxidized form. It folds fast relative to the rate of biosynthesis, to the same conformation as the oxidized form. Folding and disulfide formation are sequential. These results indicate that the proenzyme folds sequentially in vivo and that the residual pro-part constitutes a rapidly and independently folding domain that stabilizes the mature enzyme. It thus fulfills the criteria required of an intramolecular chaperone. It may also be involved in stabilizing the tetrameric structure of the mature enzyme.  相似文献   

9.
Disulfide exchange folding of insulin-like growth factor I.   总被引:11,自引:0,他引:11  
The disulfide exchange folding properties of insulin-like growth factor I (IGF-I) have been analyzed in a redox buffer containing reduced (10 mM) and oxidized (1 mM) glutathione. Under these conditions, the 3 disulfide bridges of the 70 amino acid peptide were not quantitatively formed. Instead, five major forms of IGF-I were detected, and these components were concluded to be in equilibrium as their relative amounts were similar starting from either reduced, native, or a mismatched variant of IGF-I containing two non-native disulfides. The different components in the mixtures were trapped by thiol alkylation using vinylpyridine and subsequently isolated by reverse-phase HPLC. The purified variants were further characterized using plasma desorption mass spectrometry and peptide mapping. Two of the five different forms were identified as native and mismatched IGF-I. One form was a variant with only one disulfide bond, and the other two major components had two disulfides formed. In a separate experiment, early refolding intermediates were trapped by pyridylethylation after only 90 s of refolding in the glutathione buffer, starting from reduced IGF-I. The intermediates were identical to the components observed at equilibrium, but at different relative concentrations. On the basis of the disulfide bond patterns of the different components in the equilibrium mixtures, we conclude that the disulfide between cysteines-47 and -52 in IGF-I is an unfavorable high-energy bond that may exist in the native molecule in a strained configuration.  相似文献   

10.
The 52-residue alpha/beta chimera of the epidermal growth factor-like domain in neu differentiation factor (NDFealpha/beta) has been synthesized and folded to form a three disulfide bridge (Cys182-Cys196, Cys190-Cys210, Cys212-Cys221) containing peptide. We investigated two general strategies for the formation of the intramolecular disulfide bridges including, the single-step approach, which used fully deprotected and reduced peptide, and a sequential approach that relied on orthogonal cysteine protection in which specific pairs are excluded from the first oxidation step. Because there are 15 possible disulfide bridge arrangements in a peptide with six cysteines, the one-step approach may not always provide the desired disulfide pairing. Here, we compare the single-step approach with a systematic evaluation of the sequential approach. We employed the acetamidomethyl group to protect each pair of cysteines involved in disulfide bridges, i.e. Cys182 to Cys196, Cys190 to Cys210 and Cys212 to Cys221. This reduced the number of possible disulfide patterns from 15 to three in the first folding step. We compared the efficiencies of folding for each protected pair using RP-HPLC, mapped the disulfide connectivity of the predominant product and then formed the final disulfide from the partially folded intermediate via 12 oxidation. Only the peptide having the Cys182-Cys196 pair blocked with acetamidomethyl forms the desired disulfide isomer (Cys190-Cys210/Cys212-Cys221) as a single homogeneous product. By optimizing both approaches, as well as other steps in the synthesis, we can now rapidly provide large-scale syntheses of NDFealpha/beta and other novel EGF-like peptides.  相似文献   

11.
The hepatic, microsomal, thiol:protein disulfide oxidoreductase catalyzes the glutathione (GSH) reduction of protein disulfides to sulfhydryl groups. In the presence of physiological concentrations of glucagon this activity increased from 2.3 to 6.4 fold in isolated microsomes. The stimulation had a P50 for glucagon of 7.8 X 10(-10) M which was only observed at microsomal protein concentrations of less than 100 micrograms/ml and in the presence of a GSH reducing system. This latter observation suggests that the stimulation may be inhibited by the presence of oxidized glutathione. These data support the hypothesis that glucagon may act in part by stimulating the reduction of protein disulfides by the thiol:protein disulfide oxidoreductase.  相似文献   

12.
A novel strategy combining Edman degradation and thiol modification was developed to assign the three disulfides of huwentoxin-II (HWTX-II), an insecticidal peptide purified from the venom of the spider Selenocosmia huwena. Phenylthiohydantoin (Pth) derivatives of Cys and the elimination product, dehydroalanine (DeltaSer), can be observed in the Cys cycles during Edman degradation of native HWTX-II. The appearance of two products indicates that the disulfides of HWTX-II were split and that the free thiol group of the second half cystine has been generated. Information about the nature of the disulfide bridges of HWTX-II could be obtained from the sequencing signal if the nascent thiols were modified stepwise by 4-vinylpyridine. Using this method the disulfide bridges of HWTX-II were assigned as Cys4-Cys18, Cys8-Cys29 and Cys23-Cys34, which is different from that seen in HWTX-I, a neurotoxic peptide from the same spider. Using this strategy, one can assign the disulfide bonds of small proteins by sequencing and modification n - 1 times, where n is the number of disulfide bonds in the protein. The above assignment of the disulfide bonds of HWTX-II was confirmed by MALDI-TOF MS of tryptic fragments of HWTX-II. Some disulfide interchanging during proteolysis was observed by monitoring the kinetics of proteolysis of HWTX-II by MALDI-TOF MS.  相似文献   

13.
Decorsin is a 39-residue RGD-protein crosslinked by three disulfide bridges isolated from the leech Macrobdella decora belonging to the family of GPIIb-IIIa antagonists and acting as a potent inhibitor of platelet aggregation. Here we report the solid-phase synthesis of decorsin using the Fmoc strategy. The crude polypeptide was purified by reverse-phase HPLC in its reduced form and allowed to refold in the presence of glutathione. The homogeneity of the synthetic oxidized decorsin was established by reverse-phase HPLC and capillary zone electrophoresis. The results of amino acid analysis after acid hydrolysis of the synthetic protein, NH2-terminal sequencing and mass determination (4,377 Da) by electrospray mass spectrometry were in full agreement with this theory. The correct pairing of the three disulfide bridges in synthetic decorsin was determined by a combined approach of both peptide mapping using proteolytic enzymes and analysis of the disulfide chirality by CD spectroscopy in the near-UV region. Synthetic decorsin inhibited human platelet aggregation with an IC50 of approximately 0.1 microM, a figure quite similar to that determined utilizing decorsin from natural source. In particular, the synthetic protein was 2,000-fold more potent than a model RGD-peptide (e.g., Arg-Gly-Asp-Ser) in inhibiting platelet aggregation. Thermal denaturation experiments of synthetic decorsin, monitored by CD spectroscopy, revealed its high thermal stability (Tm approximately 74 degrees C). The features of the oxidative refolding process of reduced decorsin, as well as the thermal stability of the oxidized species, were compared with those previously determined for the NH2-terminal core domain fragment 1-41 or 1-43 from hirudin. This fragment shows similarity in size, pairing of the three disulfides and three-dimensional structure with those of decorsin, even if very low sequence similarity. It is suggested that the less efficient oxidative folding and the enhanced thermal stability of decorsin in respect to those of hirudin core domain likely can be ascribed to the presence of the six Pro residues in the decorsin chain, whereas none is present in the hirudin domain. The results of this study indicate that decorsin can be obtained by solid-phase methodology in purity and quantities suitable for structural and functional studies and thus open the way to prepare by chemical methods novel decorsin derivatives containing unusual amino acids or even non-peptidic moieties.  相似文献   

14.
The 45-residue C-terminal EGF-like domain in human blood coagulation factor IX has been synthesized by a 2-step method to form selectively 3 disulfide bridges. Four out of 6 cysteines are blocked with either trityl or 4-methyl-benzyl, and the remaining 2 cysteines are blocked with acetamidomethyl (Acm). In the first step, 4 free cysteinyl thiols are released concurrently with the removal of all protecting groups except Acm and are oxidized to form 1 of the 3 possible isomers containing 2 pairs of disulfides. In the second step, iodine is used to remove the Acm groups to yield the third disulfide bridge. This approach reduces the number of possible disulfide bridging patterns from 15 to 3. To determine the optimal protecting group strategy, 3 peptides are synthesized, each with Acm blocking 1 of the 3 pairs of cysteines involved in disulfide bridges: Cys5 to Cys16 (Cys 1-3), Cys12 to Cys26 (Cys 2-4), or Cys28 to Cys41 (Cys 5-6). Only the peptide having the Cys 2-4 pair blocked with Acm forms the desired disulfide isomer (Cys 1-3/5-6) in high yield after the first step folding, as identified by proteolytic digestion in conjunction with mass spectrometric peptide mapping. Thus, the choice of which pair of cysteines to block with Acm is critically important. In the case of EGF-like peptides, it is better to place the Acm blocking groups on one of the pairs of cysteines involved in the crossing of disulfide bonds.  相似文献   

15.
Studies on the mechanism of the glutathione regeneration (Saxena, V.P., and Wetlaufer, D.B. (1970) Biochemistry 9, 5015-5023) of hen egg lysozyme have been carried out. The first two stoichiometric disulfides in lysozyme are formed about 8 times more rapidly than the second two. Almost no enzymic activity is regained until the first two disulfides are formed, thus ruling out an all-or-none mechanism. The disulfide peptides formed early in the regeneration have been isolated and identified. The results show a limited search of folding intermediates, and outline a folding pathway. The early disulfides involve cysteinyl residues III, IV, V, and VI. At the same time cysteinyl residues I, II, VII, and VIII are still reduced, as demonstrated by their isolation as S-alkylated derivatives. At slightly later times a peptide is found which contains the (native) disulfide between cysteinyl residues II and VII. It is likely, but as yet unproven, that formation of disulfide I-VIII completes the cross-linking of lysozyme.  相似文献   

16.
Current evidence indicates that the ligand-facilitated dimerization of neurophysin is mediated in part by dimerization-induced changes at the hormone binding site of the unliganded state that increase ligand affinity. To elucidate other contributory factors, we investigated the potential role of neurophysin's short interdomain loop (residues 55-59), particularly the effects of loop residue mutation and of deleting amino-terminal residues 1-6, which interact with the loop and adjacent residues 53-54. The neurophysin studied was bovine neurophysin-I, necessitating determination of the crystal structures of des 1-6 bovine neurophysin-I in unliganded and liganded dimeric states, as well as the structure of its liganded Q58V mutant, in which peptide was bound with unexpectedly increased affinity. Increases in dimerization constant associated with selected loop residue mutations and with deletion of residues 1-6, together with structural data, provided evidence that dimerization of unliganded neurophysin-I is constrained by hydrogen bonding of the side chains of Gln58, Ser56, and Gln55 and by amino terminus interactions, loss or alteration of these hydrogen bonds, and probable loss of amino terminus interactions, contributing to the increased dimerization of the liganded state. An additional intersubunit hydrogen bond from residue 81, present only in the liganded state, was demonstrated as the largest single effect of ligand binding directly on the subunit interface. Comparison of bovine neurophysins I and II indicates broadly similar mechanisms for both, with the exception in neurophysin II of the absence of Gln55 side chain hydrogen bonds in the unliganded state and a more firmly established loss of amino terminus interactions in the liganded state. Evidence is presented that loop status modulates dimerization via long-range effects on neurophysin conformation involving neighboring Phe22 as a key intermediary.  相似文献   

17.
Size-exclusion high performance liquid chromatography was used to compare the Stokes radius of the mixed disulfide of trypsinogen refolded for 10 min with the Stokes radius of denatured trypsinogen in high concentrations of urea. After folding for 10 min, rechromatography of a collection of sequential fractions of an initial separation showed that the fractions display microheterogeneity as seen in the value of the Stokes radius of each fraction. These intermediate species differed in their Stokes radius, and each had a globular structure cross-linked by disulfide bonds. In contrast, when trypsinogen with the native disulfides intact was equilibrated at different concentrations of urea (0-8 M), a progressive increase in Stokes radius was observed with extent of unfolding. Rechromatography of a series of fractions collected at a specific urea concentration showed that each had the same Stokes radius as the fraction in the initial separation. Urea-denatured trypsinogen and partially refolded trypsinogen must therefore differ in the disulfide pairing that links regions of the polypeptide chain. These observations support the suggestion that non-native disulfide bonds are responsible for the many stable conformations that form early in the folding of the mixed disulfide of trypsinogen (Light, A., and Higaki, J.N. (1987) Biochemistry 26, 5556-5564). These intermediates initially are loose structures (large Stokes radius) that become more compact with time (decreasing Stokes radius). The intermediates must therefore undergo a continuing disulfide interchange until native disulfides form late in the process when the stable conformation of the native molecule is reached.  相似文献   

18.
S Eubanks  M Lu  D Peyton  E Breslow 《Biochemistry》1999,38(41):13530-13541
Earlier thermodynamic studies of the intermolecular interactions between mature oxytocin and neurophysin, and of the effects of these interactions on neurophysin folding, raised questions about the intramolecular interactions of oxytocin with neurophysin within their common precursor. To address this issue, the disulfide-rich precursor of oxytocin-associated bovine neurophysin was expressed in Escherichia coli and folded in vitro to yield milligram quantities of purified protein; evidence of significant impediments to yield resulting from damage to Cys residues is presented. The inefficiency associated with the refolding of reduced mature neurophysin in the presence of oxytocin was found not to be alleviated in the precursor. Consistent with this, the effects of pH on the spectroscopic properties of the precursor and on the relative stabilities of the precursor and mature neurophysin to guanidine denaturation indicated that noncovalent intramolecular bonding between oxytocin and neurophysin in the precursor had only a small thermodynamic advantage over the corresponding bonding in the intermolecular complex. Loss of the principal interactions between hormone and protein, and of the enhanced stability of the precursor relative to that of the mature unliganded protein, occurred reversibly upon increasing the pH, with a midpoint at pH 10. Correlation of these results with evidence from NMR studies of structural differences between the precursor and the intermolecular complex, which persist beyond the pH 10 transition, suggests that the covalent attachment of the hormone in the precursor necessitates a conformational change in its neurophysin segment and leads to properties of the system that are distinct from those of either the liganded or unliganded mature protein.  相似文献   

19.
The structural and functional roles of the two disulfide bridges in interleukin-8 (IL-8) were addressed using IL-8 analogues with covalently modified disulfide bridges. The analogues were prepared using chemical synthesis by replacement of a cysteine for either homocysteine, penicillamine, or selenocysteine and on folding resulted in a covalently modified disulfide. Deletion of either of the two disulfide bridges by replacement of either cysteine pair with alanine resulted in loss of both structure and function. In contrast, all of the analogues with modified disulfide bridges had native tertiary fold as determined by nuclear magnetic resonance spectroscopic methods. Their structural similarity provided a rational basis for assessing the functional effects of the changes to the disulfide. Modification to the disulfide bridge between cysteines 9 and 50 had only a modest effect on IL-8 function. In contrast, alterations to the 7-34 disulfide bridge resulted in a dramatic reduction in biological potency. Thus, although both disulfide bridges are required for maintenance of the native tertiary fold, their role in determining IL-8 activity is distinct. We propose that 7-34 disulfide has a direct role in determining receptor binding and activation, whereas the 9-50 was not directly involved. The synthesis of non-natural disulfide analogues is a novel general approach to structure-activity relationships of disulfide bridges. The demonstration that the participation of disulfide bridges in function can be dissociated from their effects on the stability of the tertiary structure suggests that this method will lead to increased understanding of the roles of disulfide bridges in proteins.  相似文献   

20.
The in vitro refolding process of the double-chain insulin was studied based on the investigation of in vitro single-chain insulin refolding. Six major folding intermediates, named P1A, P2B, P3A, P4B, P5B, and P6B, were captured during the folding process. The refolding experiments indicate that all of these intermediates are on-pathway. Based on these intermediates and the formation of hypothetic transients, we propose a two-stage folding pathway of insulin. (1) At the early stage of the folding process, the reduced A chain and B chain individually formed the intermediates two A chain intermediates (P1A and P3A), and four B chain intermediates (P2B, P4B, P5B, and P6B). (2) In the subsequent folding process, transient Ⅰ was formed from P3A through thiol/disulfide exchange reaction; then, transients Ⅱ and Ⅲ, each containing two native disulfides, were formed through the recognition and interaction of transient Ⅰ with P4B or P6B and the thiol group's oxidation reaction mainly using GSSG as oxidative reagent; finally, transients Ⅱ and Ⅲ, through thiol/mixture disulfide exchange reaction, formed the third native disulfide of insulin to complete the folding.  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