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 15 毫秒
1.
Cyclic mono-cystinyl active-site fragments of thioredoxin and thioredoxin reductase were synthesized as N-acetyl and C-amide octapeptides by conventional methods of peptide synthesis in solution and on solid supports. Using a side-chain protection based on acid-labile tert-butanol-derived groups and on the S-tert-butylthio unsymmetric disulfide for the thiol functions, in combination with Nα-Z- or Nα-Nps derivatives in the chain elongation steps, the synthesis in solution was carried out in straightforward manner yielding the fully protected octapeptides as well characterized compounds. Upon deprotection with trifluoroacetic acid and reduction of the unsymmetrical disulfides with tri-butylphosphine, the resulting bis-cysteinyl-octapeptides were oxidized in dimethylformamide with azodicarboxylic acid di-tert-butyl ester to produce the desired cyclic compounds in good overall yields. For the synthesis on solid supports a similar acid-labile side-chain protection was applied in combination with the Nα 9-flourenylmethyoxycarbonyl derivatives in the chain elongation steps. Thereby acylations were performed with the related amino acid N-car-boxyanhydrides (UNCAs) or by the O-(1H-benzotriazol-1-yl)-N,N,N′,N′-tetramethyl-uronium-tetrafluoroborate/1-hydroxybenzotriazole (TBTU/HOBt) procedure. The solid phase synthesis of the two octapeptides led to unexpected difficulties in terms of recovery of peptidic material from the resins in the final acidolytic cleavage step as well as of racemization at the level of the cysteine residues by the TBTU/HOBt coupling method. Racemization was efficiently suppressed by employing the related pentafluorophenyl ester and this method led to crude octapeptide products of a degree of purity comparable to those obtained by the synthesis in solution. However, the recovery of the peptides from the resin, i.e., irreversible reattachment of cleaved peptidic material via alkylation of various side-chain functions, could not be avoided even using the most efficient scavengers or their cocktails. © 1994 John Wiley & Sons, Inc. © 1994 John Wiley & Sons, Inc.  相似文献   

2.
Two methionine-containing peptides, endothelin 1 and the 1-16 fragment of the receptor of the plasminogen activator 1 for human urokinase, were synthesized and cyclized by hydrogen peroxide. Endothelin 1 was obtained by using regioselective and random schemes of disulfide bond formation. The conditions of cyclization that provided the target products in high purity were found. The general potential of disulfide bond formation by means of hydrogen peroxide was demonstrated for methionine-containing peptides. The method resulted in target products containing insignificant quantities of the corresponding Met-sulfoxide derivatives.  相似文献   

3.
Kentsis A  Mezei M  Osman R 《Proteins》2005,61(4):769-776
Recent studies have indicated that the unfolded states of polypeptides contain a substantial amount of polyproline type II (P(II)) structure. This energetically and structurally preorganized state may contribute to the reduction of the folding search, as well as to the recognition of intrinsically unstructured proteins and polyproline ligands. Using Monte Carlo simulations of natively unfolded peptides in the presence of explicit aqueous solvation, we observe that residue-specific P(II) conformational propensity is the result of the modulation of polypeptide backbone hydration by a proximal side-chain. Such a mechanism may be unique among those that contribute to the modulation of secondary structures in proteins. The calculated conformational propensities should prove useful for the development of a configurational P(II) scale necessary for the prediction and design of natural-like polypeptides.  相似文献   

4.
Experimental techniques with high temporal and spatial resolution extend our knowledge of how biological macromolecules self-organise and function. Here, we provide an illustration of the convergence between simulation and experiment made possible by techniques such as triplet-triplet energy transfer and fluorescence quenching with long-lifetime and fast-quenching fluorescent probes. These techniques have recently been used to determine the average time needed for two residues in a peptide or protein segment to form a contact. The timescale of this process is accessible to computer simulation, providing a microscopic interpretation of the data and yielding new insight into the disordered state of proteins. Conversely, such experimental data also provide a test of the validity of alternative choices for the molecular models used in simulations, indicating their possible deficiencies. We carried out simulations of peptides of various composition and length using several models. End-to-end contact formation rates and their dependence on peptide length agree with experimental estimates for some sequences and some force fields but not for others. The deviations are due to artefactual structuring of some peptides, which is not observed when an atomistic model for the solvation water is used. Simulations show that the observed experimental rates are compatible with considerably different distributions of the end-to-end distance; for realistic models, these are never Gaussian but indicative of a rugged energy landscape.  相似文献   

5.
The assembly of reduced pro-alpha chains of type I and type II procollagen into the native triple-helical molecule was examined in vitro in the presence and absence of pure protein disulfide isomerase. The data clearly indicates that protein disulfide isomerase is able to accelerate the formation of native interchain disulfide bonds in these procollagens. It takes about 6 min after disulfide bonding before triple-helical molecules exist, while the time required to produce triple-helical type I procollagen in the presence of protein disulfide isomerase is 9.4 min and that for type II procollagen 17.2 min. These values agree with those obtained for type I and II procollagen in vivo suggesting that protein disulfide isomerase is also an enzyme catalyzing interchain disulfide bond formation in procollagen in vivo. The formation of native disulfide bonds can proceed without any enzyme catalysis but then requires the presence of reduced and oxidized glutathione. Bonding is rather slow in such a case, however, resulting in a delay in the formation of the triple helix.  相似文献   

6.
7.
X Liu  P G Scott  A Otter  G Kotovych 《Biopolymers》1992,32(2):119-130
The importance of side-chain charge interactions in the formation of beta-turns was studied. Sixteen protected NAc-tetrapeptide amides were studied, namely the variants of DEKS: NEKS, EEKS, DDKS, DQKS, NQKS, DERS, NERS, EERS, DDRS, NDRS, DQRS, and DKES. Three tetrapeptides--NPDM, NSDM, and NDDS--were also studied as they have a high probability of forming beta-turns, based on statistical predictions. The results indicate that a small proportion of type I beta-turn exists in solutions of DEKS and DERS in methanol/water (60/40), while NEKS has an even smaller population of this turn. The other tetrapeptides are present in solution only in the extended conformation. These results clearly show the importance of the salt bridge between the side chains of K2 and E3 or R2 and E3, as well as the importance of the charge on the side chain of the first residue in stabilizing the beta-turn. The relevance of statistical predictions for beta-turns in short peptides is discussed.  相似文献   

8.
Hybridization kinetics of DNA sequences with known secondary structures and random sequences designed with similar melting temperatures were studied in solution and when one strand was bound to 5 μm silica microspheres. The rates of hybridization followed second-order kinetics and were measured spectrophotometrically in solution and fluorometrically in the solid phase. In solution, the rate constants for the model sequences varied by almost two orders of magnitude, with a decrease in the rate constant with increasing amounts of secondary structure in the target sequence. The random sequences also showed over an order of magnitude difference in the rate constant. In contrast, the hybridization experiments in the solid phase with the same model sequences showed almost no change in the rate constant. Solid phase rate constants were approximately three orders of magnitude lower compared with the solution phase constants for sequences with little or no single-stranded structure. Sequences with a known secondary structure yielded solution phase rate constants as low as 3 × 103 M−1 s−1 with solid phase rate constants for the same sequences measured at 2.5 × 102 M−1 s−1. The results from these experiments indicate that (i) solid phase hybridization occurs three orders of magnitude slower than solution phase, (ii) trends observed in structure-dependent kinetics of solution phase hybridization may not be applicable to solid phase hybridization and (iii) model probes with known secondary structure decrease reaction rates; however, even random sequences with no known internal single-stranded structure can yield a broad range of reaction rates.  相似文献   

9.
A simple, sensitive, reliable method for determining disulfide groups in peptides is presented. The disulfides are cleaved in a brief treatment with strong alkali. Following neutralization with phosphoric acid, thiol resulting from the alkaline cleavage is estimated colorimetrically with 5,5′-dithio-bis(2-nitrobenzoic acid). In the presence of EDTA, the color yield is stable and is linear with the concentration of oxidized glutathione. The stoichiometry with other peptide disulfides appears to be somewhat variable but not so as to interfere with detection of peptide disulfides in chromatographic fractions. The present method compares favorably with two other proposed disulfide analytical methods. The cleavage assay is chromogenic with disulfides, thiols, and with certain blocked thiols but is not chromogenic with methionine and lanthionine.  相似文献   

10.
Both metalloprotein and flavin-linked sulfhydryl oxidases catalyze the oxidation of thiols to disulfides with the reduction of oxygen to hydrogen peroxide. Despite earlier suggestions for a role in protein disulfide bond formation, these enzymes have received comparatively little general attention. Chicken egg white sulfhydryl oxidase utilizes an internal redox-active cystine bridge and a FAD moiety in the oxidation of a range of small molecular weight thiols such as glutathione, cysteine, and dithiothreitol. The oxidase is shown here to exhibit a high catalytic activity toward a range of reduced peptides and proteins including insulin A and B chains, lysozyme, ovalbumin, riboflavin-binding protein, and RNase. Catalytic efficiencies are up to 100-fold higher than for reduced glutathione, with typical K(m) values of about 110-330 microM/protein thiol, compared with 20 mM for glutathione. RNase activity is not significantly recovered when the cysteine residues are rapidly oxidized by sulfhydryl oxidase, but activity is efficiently restored when protein disulfide isomerase is also present. Sulfhydryl oxidase can also oxidize reduced protein disulfide isomerase directly. These data show that sulfhydryl oxidase and protein disulfide isomerase can cooperate in vitro in the generation and rearrangement of native disulfide pairings. A possible role for the oxidase in the protein secretory pathway in vivo is discussed.  相似文献   

11.
Two cyclic peptide disulfides (Sequence: see text). (X = L-Tyr or L-Phe) have been synthesized as models for the 14-membered redox-active disulfide loop of glutaredoxin. 1H NMR studies at 270 MHz in chloroform solutions establish a type I beta-turn conformation for the Pro-X segment in both peptides, stabilized by a 4----1 hydrogen bond between the Cys(1) CO and Cys(4) NH groups. Nuclear Overhauser effects establish that the aromatic ring in the X = Phe peptide is oriented over the central peptide unit. In dimethyl sulfoxide solutions two conformational species are observed in slow exchange on the NMR time scale, for both peptides. These are assigned to type I and type II beta-turn structures with -Pro-Tyr(Phe)- as the corner residues. The structural assignments are based on correlation of NMR parameters with model 14-membered cyclic cystine peptides with Pro-X spacers. Circular dichroism studies based on the -S-S- n-omega* transition suggest a structural change in the disulfide bridge with changing solvent polarity, establishing conformational coupling between the peptide backbone and the disulfide linkage in these systems.  相似文献   

12.
N L Marky  W K Olson 《Biopolymers》1987,26(3):415-438
The flexibility of hairpin loops containing n bases (residues) has been examined using a theoretical model [N. L. Marky and W. K. Olson (1982), Biopolymers, 21 , 2329–2344] of oligonucleotide loop closure. The study is based on correlated probabilities of chain separation and terminal residue orientation as outlined previously. The probabilities are calculated using standard statistical mechanical methods as functions of local conformational changes of the chain backbone. Our results for an RNA chain of 9 residues suggest that the anticodon loop is a dynamic structure capable of assuming a variety of different spatial conformations. Free energy values related to the various conformations span a narrow range of values (2–4 kcal/mole) and compare well with experimental observations in aqueous solution. Conformational transitions between the loop conformations are within less than 0.5 kcal/mole in free energy. The different spatial loop conformations and the likely pathways between them may have potential relevance to the molecular translation of the genetic code.  相似文献   

13.
A comparative study of modern coupling reactions involving Boc-protected amino acid derivatives and dipeptides with N-terminal alpha,alpha-dialkylation and N-methylation was carried out. The coupling reactions were run using either equimolar amounts of the amino and activated carboxyl components or an excess of the activated carboxyl component. Yields of the target tripeptide Boc-Phe-Xaa-Phe-OBzl (Xaa = (NMe)Ala, (NMe)Aib, or (NMe) alpha Ac5c) were compared. Less than 10% of the product was obtained from methods utilizing pivaloyl mixed anhydride, pentafluorophenyl ester or acyl fluoride activation when Xaa = (NMe)Aib and (NMe) alpha Ac5c. At room temperature, significant yields of these two products were obtained from reactions which utilized an excess of the HBTU reagent (O-benzotriazol-1-yl-N,N,N',N'-tetramethyluronium hexafluorophosphate), the PyBroP reagent (bromo-tris-pyrrolidino-phosphonium hexafluorophosphate) or Boc-Phe-NCA (Boc-protected phenylalanine N-carboxyanhydride). Moreover, the Boc-Phe-NCA method was superior when used over a prolonged reaction time or at elevated temperature.  相似文献   

14.
To investigate the effect of “secondary interaction” on hydrolysis by various acid proteinases from molds and yeasts, synthetic peptides
amino acid residues) were used as substrates. Pepsin was used for the comparative study. These peptides were split at the peptide bonds indicated by the arrows, permitting examination of the effect of residue X distant by two or three amino acid residues from the hydrolytic site in the peptides. According to the system of Schechter and Berger (Biochem. Biophys. Res. Commun. 27; 157, 1967), the amino acid residues in peptide substrates were numbered P1, P2, etc. toward the N-terminal direction from the site of hydrolysis, and P1′, P2′, etc. toward the C-terminal direction. The results indicated that hydrolysis by these microbial enzymes is affected by at least six amino acid residues (P1-P3 and P1′-P3′) in peptide substrates, as is seen with pepsin. Elongation of the peptide chain with suitable amino acid residues from P1 to P2 or P3 and from P1′ to P2′ or P3′ in peptide substrates resulted in much or less increase of hydrolysis depending upon the species of the enzyme producers.  相似文献   

15.
The hallmark of T cell responses to staphylococcal enterotoxins (SE) and other super-Ag is a selective stimulation of cells expressing particular TCR-V beta segments. Our previous studies suggested that the disulfide loop in SE is critical for their interaction with the TCR. To investigate this concept in further detail we constructed disulfide loop mutants of staphylococcal enterotoxin A (SEA), and examined these altered toxins for mitogenicity, class II MHC binding, and V beta specificity. We found that substitutions of either Cys-96 or Cys-106 decreased mitogenicity by 100-fold without significantly affecting class II binding or resistance of the molecule to proteolysis. Several mutants lost the capacity to stimulate V beta 11+ cells, except a Cys-106----Gln mutant for which V beta 11-stimulatory activity was increased. By contrast, mutants containing Cys----Ala substitutions acquired the capacity to stimulate V beta 6+ cells. Despite these effects of V beta specificity, all mutants retained the predominant preference of SEA for V beta 3+ cells. Neither exchange of regions flanking the loop in SEA with corresponding residues in SEB, nor conversion of the entire loop region of SEA to that of SEE, were associated with transfers of V beta specificity. Our results suggest that the disulfide loop in SEA contributes to toxin avidity for the TCR, rather than specificity for particular V beta.  相似文献   

16.
We have developed a statistical mechanics algorithm, TANGO, to predict protein aggregation. TANGO is based on the physico-chemical principles of beta-sheet formation, extended by the assumption that the core regions of an aggregate are fully buried. Our algorithm accurately predicts the aggregation of a data set of 179 peptides compiled from the literature as well as of a new set of 71 peptides derived from human disease-related proteins, including prion protein, lysozyme and beta2-microglobulin. TANGO also correctly predicts pathogenic as well as protective mutations of the Alzheimer beta-peptide, human lysozyme and transthyretin, and discriminates between beta-sheet propensity and aggregation. Our results confirm the model of intermolecular beta-sheet formation as a widespread underlying mechanism of protein aggregation. Furthermore, the algorithm opens the door to a fully automated, sequence-based design strategy to improve the aggregation properties of proteins of scientific or industrial interest.  相似文献   

17.
Models of F-actin structure predict the importance of hydrophobic loop 262-274 at the interface of subdomains 3 and 4 to interstrand interactions in filaments. If this premise is correct, prevention of the loop conformational change--its swinging motion--should abort filament formation. To test this hypothesis, we used site-directed mutagenesis to create yeast actin triple mutant (LC)2CA (L180C/L269C/C374A). This mutation places two cysteine residues in positions potentially enabling the locking of loop 262-274 to the monomer surface via disulfide formation. Exposure of the purified mutant to oxidation catalysts resulted in an increased electrophoretic mobility of actin on SDS PAGE and a loss of two cysteines by DTNB titrations, consistent with disulfide formation. The polymerization of un-cross-linked mutant actin by MgCl2 was inhibited strongly but could be restored to wild type actin levels by phalloidin and improved greatly through copolymerization with the wild-type actin. Light scattering measurements revealed nonspecific aggregation of the cross-linked actin under the same conditions. Electron microscopy confirmed the absence of filaments and the presence of amorphous aggregates in the cross-linked actin samples. Reduction of the disulfide bond by DTT restored normal actin polymerization in the presence of MgCl2 and phalloidin. These observations provide strong experimental support for a critical role of the hydrophobic loop 262-274 in the polymerization of actin into filaments.  相似文献   

18.
Bioactive peptides have an important multifunctional role in the gastrointestinal tract. In the present study we have investigated the dynamism of the appearance of PACAP (pituitary adenylate cyclase activating polypeptide), VIP (vasoactive intestinal polypeptide), gastrin, and secretin immunoreactivities in human foregut derivates during the ontogenesis using an immunohistochemical approach. None of these peptides were observed in the foregut derivates of an 8-week-old embryo. VIP immunoreactive nerve fibers appeared by the 11th week in the smooth muscle layers of the stomach. No other peptide immunoreactivities were observed of this stage. In 18- and 20-week old fetuses PACAP, secretin, and gastrin immunoreactive cells appeared in the developing glands of the stomach. In the duodenum gastrin immunoreactivity was present in the Lieberkühn's glands and secretin immunoreactive cells were seen between the surface epithelial cells. In the pancreas secretin immunoreactivity was found in the Langerhans islets; however, PACAP immunreactivity was observed in the exocrine portion. The distribution of VIP fibers did not change during the fetal life and it was similar to the adult pattern. According to our results the appearance of PACAP, secretin, and gastrin in the developing glands suggests their role in the proliferation and differentiation of the epithelial derivates.  相似文献   

19.
Microscopic disulfide-exchange rate constants have been measured for the formation and opening of small disulfide loops in reactions between glutathione and peptides containing 2 cysteines. Twelve cysteine-Xm-cysteine peptides have been studied, where X is an amino acid and m is the number of amino acids between the cysteines. Homopolymers of alanine for m equaling 0-5 are evaluated, as well as X1 and X2 series employing glycine, valine, or proline. Equilibrium constants Kc for loop closing are only slightly dependent on the nature of X. Loops with even values of m generally are favored relative to loops with odd values. Kc increases in the rank order X1, X3, X0, X5, X4, and X2. Formation of a disulfide between sequentially adjacent cysteines therefore is not especially difficult. The dependence of Kc on the odd-even nature of m is compared with similar patterns observed both in statistics of disulfide formation in naturally occurring proteins and in theoretical studies of peptide cyclization. The relative equilibrium populations of intramolecular disulfides in peptides containing cysteine-cysteine-cysteine and cysteine-serine-cysteine-serine-cysteine clusters are consistent with predictions based on the values of Kc in the two-cysteine peptides.  相似文献   

20.
Native protein disulfide bond formation in the endoplasmic reticulum (ER) requires protein disulfide isomerase (PDI) and Ero1p. Here we show that oxidizing equivalents flow from Ero1p to substrate proteins via PDI. PDI is predominantly oxidized in wild-type cells but is reduced in an ero1-1 mutant. Direct dithiol-disulfide exchange between PDI and Ero1p is indicated by the capture of PDI-Ero1p mixed disulfides. Mixed disulfides can also be detected between PDI and the ER precursor of carboxypeptidase Y (CPY). Further, PDI1 is required for the net formation of disulfide bonds in newly synthesized CPY, indicating that PDI functions as an oxidase in vivo. Together, these results define a pathway for protein disulfide bond formation in the ER. The PDI homolog Mpd2p is also oxidized by Ero1p.  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