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.
Beta2-microglobulin (beta2-m), a major component of dialysis-related amyloid fibrils, has an intrachain disulfide bond buried inside the native structure. We examined the conformation of beta2-m amyloid fibrils by analyzing the reactivity of the disulfide bond to a reducing reagent, dithiothreitol. Although the disulfide bond in the native structure was highly protected from reduction, the disulfide bonds in the amyloid fibrils prepared at pH 2.5 were progressively reduced at pH 8.5 by 50 mm dithiothreitol. Because beta2-m amyloid fibrils prepared under acidic conditions have been known to depolymerize at a neutral pH, we examined the relation between depolymerization and reduction of the disulfide bond. The results indicate that the disulfide bonds in the amyloid fibrils were protected from reduction, and the reduction occurred during depolymerization. On the other hand, the disulfide bonds of immature filaments, the thin and flexible filaments prepared under conditions of high salt at pH 2.5, were reduced at pH 8.5 more readily than those of amyloid fibrils, suggesting that the disulfide bonds are exposed to the solvent. Taken together, the disulfide bond once exposed to the solvent upon acid denaturation may be progressively buried in the interior of the amyloid fibrils during its formation.  相似文献   

2.
In a previous report (J. Biol. Chem. 258 (1983) 3565–3570) we have demonstrated that the disulfide-reducing agent dithiothreitol has two effects on the sodium-dependent outer cortical brush border membrane d-glucose transporter; the first results in a reversible increase in the affinity of the transporter for the non-transported competitive inhibitor phlorizin, while the second results in a partially reversible loss of phlorizin binding and glucose-transport activity. Evidence was presented that both of these effects are the result of the reduction of disulfide bonds on the transport molecule. In the present paper we extend our observations on the inactivation of the transporter by dithiothreitol. We provide evidence here (i) that the inactivation of the transporter by dithiothreitol is independent of the effect of the reducing agent on the affinity of the transporter, (ii) that this inactivation process is first-order in dithiothreitol and thus presumably due to the reduction of a single disulfide bond essential to the functioning of the transporter. (iii) that it is the reduction of this disulfide bond and not some subsequent conformational or other change in the transporter which results in its inactivation, (iv) that phlorizin and substrates of the transporter provide protection against inactivation by dithiothreitol and that the degree of protection provided correlates well with the known specificity and phlorizin-binding properties of the transporter, and (iv) that the reactivity of the transporter with dithiothreitol is pH-dependent, decreasing with increasing pH over the pH range 6.5–8.5. We conclude that this site of action of dithiothreitol is a single essential disulfide bond intimately associated with the glucose-binding site on the transport molecule.  相似文献   

3.
Bovine plasma albumin (BPA) has approximately one SH group (Cys-34) which catalyzes the intramolecular SH, S-S exchange reaction in the alkaline region at low ionic strength, resulting in the formation of the aged form. So, the N-B transition at ionic strength above 0.20 and below 0.10 was studied using BPA and iodoacetamide-blocked BPA (IA-BPA), respectively. (1) pH profiles of [theta]262 and [theta]268 of BPA in 0.20 M KCl showed the characteristic changes in the pH region 7.0-9.0, corresponding to the N-B transition. On going from pH 7.0 to 9.0 in 0.10 M KCl or NaCl, IA-BPA did not show significant changes in rotational relaxation times of tryptophyl fluorophors, CD-resolved secondary structures, spin-echo 1H-n.m.r. spectra and cross-relaxation times (TIS) between irradiated and observed protein protons, which might reflect the rigidity of the domains and/or subdomains. On the other hand, rotational relaxation times of 1-anilino-8-naphthalenesulfonate-IA-BPA complex (IA-BPA-ANS0.9, molar ratio of ANS to IA-BPA = 0.9/1) showed significant decreases from 131 to 114 ns on going from the N- to the B-forms in 0.10 M KCl. The above results and reported experimental evidence might indicate that on going from the N- to the B-forms in 0.10 M KCl or NaCl, the mutual movement of subdomains, connected with a flexible hinge region (Brown & Shockley (1982)) might increase without loss in the helicity and the rigidity of subdomains. (2) The N-B transition of IA-BPA in the absence of salt was quite different from those in 0.10 M KCl or NaCl. Decreases in the helicity and the intramolecular rigidity, as monitored by TIS-measurements, were observed on going from the N- to the B-forms.  相似文献   

4.
Reduction of disulfide linkages by dithiothreitol removes LETS (large, external, transformation-sensitive) protein from the cell surface. This process is dependent upon the concentration of dithiothreitol and the time and temperature of reaction. At 0 degrees C the release of LETS protein by dithiothreitol is completely blocked, but this is apparently not due to a requirement for metabolic energy. At this temperature, reduction of LETS protein is incomplete. These results suggest that intact disulfide bonds are involved in the retention of this protein on the cell surface. Furthermore, reduction of purified LETS protein interferes with its ability to confer flattened morphology and increased adhesivity when added to transformed cells. It appears, therefore, that disulfide bonds are functionally important at the cell surface.  相似文献   

5.
Bovine plasma albumin (BPA) has 17 disulfide bonds and approximately one SH group at Cys-34 which catalyzes the intramolecular SH, S-S exchange reaction in the alkaline region at low ionic strength, resulting in the formation of the aged form (A-form). 1) Fractions of alpha-helix (f alpha) and beta-form (f beta) of iodoacetamide-blocked non-aged BPA (IA-BPA) at pH 6.5 (the N-form) and 9.0 (the B-form) in the absence of added salt were 0.70, 0.12 and 0.62, 0.18, respectively (Era et al. (1990]. However, there were no changes in f alpha and f beta of the iodoacetamide-blocked A-form (IA-A-form) over the pH range from 5.5 to 9.1 in the absence of added salt or in 0.10 M KCl (f alpha approximately 0.60, f beta approximately 0.20), indicating that the secondary structure of the IA-A-form might be similar to that of non-aged IA-BPA at pH 9.0 (B-form) in the absence of added salt, that is, the frozen B-form, stabilized covalently by the repairing of disulfide bonds. 2) The rigidity of the A- and IA-A-forms, as monitored by cross-relaxation times between irradiated and observed protein protons, was similar to or slightly higher than that of non-aged IA-BPA or BMA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Incubation of purified phosphoenolpyruvate carboxylase from Zea mays L. leaves with dithiothreitol resulted in an almost 2-fold increase in the enzymic activity. The activated enzyme showed the same affinity for its substrates and the same sensitivity with respect to malate and oxalacetate inhibition. The activation induced by dithiothreitol was reversed by diamide, an oxidant of vicinal dithiols, suggesting that the redox state of disulfide bonds of the enzyme may be important in the expression of the maximal catalytic activity.

Titration of thiol groups before and after activation of maize phosphoenolpyruvate carboxylase by dithiothreitol shows an increase of the accessible groups from 8 to 12 suggesting that the reduction of two disulfide bonds accompanied the activation. The thiols exposed by the treatment with dithiothreitol were available to reagents in nondenatured enzyme and two of them were reoxidized to a disulfide bond by diamide. It is concluded that the mechanism of phosphoenolpyruvate carboxylase activation by dithiothreitol involves the net reduction of two disulfide bonds in the enzyme.

  相似文献   

7.
M H Chau  J W Nelson 《Biochemistry》1992,31(18):4445-4450
Apamin is being studied as a model for the folding mechanism of proteins whose structures are stabilized by disulfide bonds. Apamin consists of 18 amino acid residues and forms a stable structure consisting of a C-terminal alpha-helix and two reverse turns. This structure is stabilized by two disulfide bonds connecting Cys-1 to Cys-11 and Cys-3 to Cys-15. We used glutathione and dithiothreitol as reference thiols to measure the stabilities of the two disulfide bonds as a function of urea concentration and temperature in order to understand what contributes to the stability of the native structure. The results demonstrate modest contributions from secondary structure to the overall stability of the two disulfide bonds. The equilibrium constants for disulfide bond formation between the fully reduced peptide and the native structure with two disulfide bonds at 25 degrees C and pH 7.0 are 0.42 M2 using glutathione and 2.7 x 10(-5) using dithiothreitol. The equilibrium constant decreases by a factor of approximately 4 in 8 M urea and decreases by a factor of 3 between 0 and 60 degrees C. At least three one-disulfide intermediates are found at low concentrations in the equilibrium mixture. Using glutathione, the equilibrium constants for forming the one-disulfide intermediates with respect to the reduced peptide are approximately 0.025 M. The second disulfide bond forms with an equilibrium constant of approximately 17 M. Thus, apamin folding is very cooperative, but the native structure is only modestly stabilized by urea- or temperature-denaturable secondary structure.  相似文献   

8.
Tris(2-carboxyethyl)phosphine (TCEP) is a widely used substitute for dithiothreitol (DTT) in the reduction of disulfide bonds in biochemical systems. Although TCEP has been recently shown to be a substrate of the flavin-dependent sulfhydryl oxidases, there is little quantitative information concerning the rate by which TCEP reduces other peptidic disulfide bonds. In this study, mono-, di-, and trimethyl ester analogues of TCEP were synthesized to evaluate the role of carboxylate anions in the reduction mechanism, and to expand the range of phosphine reductants. The effectiveness of all four phosphines relative to DTT has been determined using model disulfides, including a fluorescent disulfide-containing peptide (H(3)N(+)-VTWCGACKM-NH(2)), and with protein disulfide bonds in thioredoxin and sulfhydryl oxidase. Mono-, di-, and trimethyl esters exhibit phosphorus pK values of 6.8, 5.8, and 4.7, respectively, extending their reactivity with the model peptide to correspondingly lower pH values relative to that of TCEP (pK = 7.6). At pH 5.0, the order of reactivity is as follows: trimethyl- > dimethyl- > monomethyl- > TCEP > DTT; tmTCEP is 35-fold more reactive than TCEP, and DTT is essentially unreactive. Esterification also increases lipophilicity, allowing tmTCEP to penetrate phospholipid bilayers rapidly (>30-fold faster than DTT), whereas the parent TCEP is impermeant. Although more reactive than DTT toward small-molecule disulfides at pH 7.5, all phosphines are markedly less reactive toward protein disulfides at this pH. Molecular modeling suggests that the nucleophilic phosphorus of TCEP is more sterically crowded than the thiolate of DTT, contributing to the lower reactivity of the phosphine with protein disulfides. In sum, these data suggest that there is considerable scope for the synthesis of phosphine analogues tailored for specific applications in biological systems.  相似文献   

9.
Hydrogen-bonded multilayer thin films were constructed using poly(vinylpyrrolidone) and poly(methacrylic acid) functionalized with cysteamine. The resulting films included thiol moieties that were cross-linked to render the films stable at physiological pH. Film buildup was followed using quartz crystal microgravimetry, which was also used to demonstrate the improved stability imparted by reacting the thiol moieties to form disulfide bonds. Films without disulfide bonds were readily deconstructed at physiological pH, while those with disulfide bonds were swollen upon exposure to this pH (7) but remained intact. Addition of a common thiol-disulfide exchange reagent, dithiothreitol (DTT) at pH 7 led to disassembly of the multilayer films. The films were also prepared on colloidal substrates (as demonstrated using confocal microscopy) and were used to retain a model drug (fluorescently labeled transferrin) and release this molecule when triggered by the addition of DTT. This approach has potential for the in vivo applications of hollow capsules, as thiol-disulfide exchange leading to deconstruction of the capsules can occur with the assistance of intracellular proteins.  相似文献   

10.
Disulfide bridge-mediated folding of Sindbis virus glycoproteins.   总被引:3,自引:3,他引:0       下载免费PDF全文
The Sindbis virus envelope is composed of 80 E1-E2 (envelope glycoprotein) heterotrimers organized into an icosahedral protein lattice with T=4 symmetry. The structural integrity of the envelope protein lattice is maintained by E1-E1 interactions which are stabilized by intramolecular disulfide bonds. Structural domains of the envelope proteins sustain the envelope's icosahedral lattice, while functional domains are responsible for virus attachment and membrane fusion. We have previously shown that within the mature Sindbis virus particle, the structural domains of the envelope proteins are significantly more resistant to the membrane-permeative, sulfhydryl-reducing agent dithiothreitol (DTT) than are the functional domains (R. P. Anthony, A. M. Paredes, and D. T. Brown, Virology 190:330-336, 1992). We have used DTT to probe the accessibility of intramolecular disulfides within PE2 (the precursor to E2) and E1, as these proteins fold and are assembled into the spike heterotrimer. We have determined through pulse-chase analysis that intramolecular disulfide bonds within PE2 are always sensitive to DTT when the glycoproteins are in the endoplasmic reticulum. The reduction of these disulfides results in the disruption of PE2-E1 associations. E1 acquires increased resistance to DTT as it folds through a series of disulfide intermediates (E1alpha, -beta, and -gamma) prior to assuming its native and most compact conformation (E1epsilon). The transition from a DTT-sensitive form into a form which exhibits increased resistance to DTT occurs after E1 has folded into its E1beta conformation and correlates temporally with the dissociation of BiP-E1 complexes and the formation of PE2-E1 heterotrimers. We propose that the disulfide bonds within E1 which stabilize the protein domains required for maintaining the structural integrity of the envelope protein lattice form early within the folding pathway of E1 and become inaccessible to DTT once the heterotrimer has formed.  相似文献   

11.
Experimental determination of the number of thiols in a protein requires methodology that combines high sensitivity and reproducibility with low intrinsic thiol oxidation disposition. In detection of disulfide bonds, it is also necessary to efficiently reduce disulfides and to quantify the liberated thiols. Ellman's reagent (5,5'-dithiobis-[2-nitrobenzoic acid], DTNB) is the most widely used reagent for quantification of protein thiols, whereas dithiothreitol (DTT) is commonly used for disulfide reduction. DTNB suffers from a relatively low sensitivity, whereas DTT reduction is inconvenient because the reagent must be removed before thiol quantification. Furthermore, both reagents require a reaction pH > 7.0 where oxidation by ambient molecular oxygen is significant. Here we describe a quick and highly sensitive assay for protein thiol and dithiol quantification using the reducing agent sodium borohydride and the thiol reagent 4,4'-dithiodipyridine (4-DPS). Because borohydride is efficiently destroyed by the addition of acid, the complete reduction and quantification can be performed conveniently in one tube without desalting steps. Furthermore, the use of reverse-phase high-performance liquid chromatography for the thiol quantification by 4-DPS reduces the detection limit to the picomolar range (equivalent to 1 microg of a 50-kDa protein containing 1 thiol) while at the same time maintaining low pH throughout the procedure.  相似文献   

12.
Hong J  Lee A  Han H  Kim J 《Analytical biochemistry》2009,384(2):368-920
Time-dependent reduction of the disulfide bonds of immunoglobulin G (IgG) using dithiothreitol (DTT) was used to characterize the structure of IgG. After treatment with DTT, gel electrophoresis was used to separate the resulting IgG fragments, which were subsequently quantified. Using this approach, the disulfide bond between a light chain and a heavy chain was determined to be cleaved faster than the disulfide bonds between two heavy chains.  相似文献   

13.
Human adamalysin 19 (a disintegrin and metalloproteinase 19, hADAM19) is activated by furin-mediated cleavage of the prodomain followed by an autolytic processing within the cysteine-rich domain at Glu586-Ser587, which occurs intramolecularly, producing an NH2 terminal fragment (N-fragment) associated with its COOH-terminal fragment (C-fragment), most likely through disulfide bonds. When stable Madin-Darby canine kidney (MDCK) transfectants overexpressing soluble hADAM19 were treated with dithiothreitol (DTT) or with media at pH 6.5, 7.5, or 8.5, the secretion and folding of the enzyme were not affected. Autolytic processing was blocked by DTT and pH 6.5 media, which favor disulfide reduction, but was increased by pH 8.5 media, which promotes disulfide formation. Cys605, Cys633, Cys639, and Cys643 of the C-fragment appear to be partially responsible for the covalent association between the C-fragment and the N-fragment. A new autolytic processing site at Lys543-Val544 was identified in soluble mutants when these cysteine residues were individually mutated to serine residues. Shed fragments were also detectable in the media from MDCK cells stably expressing the full-length Cys633Ser mutant. Ilomastat/GM6001 inhibited hADAM19 with an IC50 of 447 nM, but scarcely affected the shedding process. The cysteine-rich domain likely forms disulfide bonds to regulate the autolytic processing and shedding of hADAM19.  相似文献   

14.
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   

15.
Human immunoglobulin G1 (IgG1) contains 12 domains, and each has an intrachain disulfide bond that connects the two layers of antiparallel β-sheets. These intrachain disulfide bonds are shielded from solvents under native conditions. Therefore, accessibility of the disulfide bonds to reduction under conditions that unfold antibody has the potential to be a good indicator of the thermodynamic stability of each domain. The stability of a recombinant monoclonal antibody at the domain level was investigated using a novel method involving reduction of the disulfide bonds in the presence of increasing amounts of guanidine hydrochloride and alkylation with [12C]iodoacetic acid, which was followed by reduction of the remaining disulfide bonds and alkylation with [13C]iodoacetic acid. The percentage of modification by [12C]iodoacetic acid of each cysteine residue was calculated using mass spectra of the cysteine-containing tryptic peptides and used to follow the unfolding of each domain. It demonstrated that the CH2 domain was the least stable domain of the antibody, whereas the CH3 domain was the most stable domain of the antibody. Other domains showed intermediate resistance to the denaturant concentration, similar to the overall unfolding transition monitored by the intrinsic tryptophan fluorescence wavelength shift.  相似文献   

16.
N S Reading  S D Aust 《Biochemistry》2001,40(27):8161-8168
Phanerochaete chrysosporium manganese peroxidase (MnP) [isoenzyme H4] was engineered with additional disulfide bonds to provide structural reinforcement to the proximal and distal calcium-binding sites. This rational protein engineering investigated the effects of multiple disulfide bonds on the stabilization of the enzyme heme environment and oxidase activity. Stabilization of the heme environment was monitored by UV-visible spectroscopy based on the electronic state of the alkaline transition species of ferric and ferrous enzyme. The optical spectral data confirm an alkaline transition to hexacoordinate, low-spin heme species for native and wild-type MnP and show that the location of the engineered disulfide bonds in the protein can have significant effects on the electronic state of the enzyme. The addition of a single disulfide bond in the distal region of MnP resulted in an enzyme that maintained a pentacoordinate, high-spin heme at pH 9.0, whereas MnP with multiple engineered disulfide bonds did not exhibit an increase in stability of the pentacoordinate, high-spin state of the enzyme at alkaline pH. The mutant enzymes were assessed for increased stability by incubation at high pH. In comparison to wild-type MnP, enzymes containing engineered disulfide bonds in the distal and proximal regions of the protein retained greater levels of activity when restored to physiological pH. Additionally, when assayed for oxidase activity at pH 9.0, proteins containing engineered disulfide bonds exhibited slower rates of inactivation than wild-type MnP.  相似文献   

17.
Effects of photooxidation of membrane thiol groups on lysosomal proton permeability were studied by measuring intralysosomal pH with fluorescein isothiocyanate-dextran and monitoring proton leakage with p-nitrophenol. Methylene blue-mediated photooxidation of lysosomes decreased their membrane thiol groups and produced cross-linking of the membrane proteins, which was established by the measurement of residual membrane thiol groups with 5,5'-dithio-bis(2-nitrobenzoic acid) and sodium dodecyl sulfate-polyacrylamide gel electrophoresis, respectively. The cross-linking of proteins could be abolished by subsequent treatment of the photodamaged lysosomes with dithiothreitol, indicating that the proteins were linked via disulfide bonds. In addition, the photodamage of lysosomes raised the intralysosomal pH and caused leakage of the lysosomal protons, which could also be reversed by subsequent dithiothreitol treatment. This indicates that lysosomal proton permeability can be increased by photooxidation of the membrane thiol groups and recovered to the normal level by reduction of the groups.  相似文献   

18.
H Kikuchi  Y Goto  K Hamaguchi 《Biochemistry》1986,25(8):2009-2013
The constant (CL) fragment of the immunoglobulin light chain contains only one intrachain disulfide bond buried in the interior of the molecule. The kinetics of reduction with dithiothreitol of the disulfide bond were studied at various concentrations of guanidine hydrochloride at pH 8.0 and 25 degrees C. It was found that the disulfide bond is reduced even in the absence of guanidine hydrochloride. The results of the reduction kinetics were compared with those of the unfolding and refolding kinetics of the CL fragment previously reported [Goto, Y., & Hamaguchi, K. (1982) J. Mol. Biol. 156, 891-910]. It was shown that the reduction of the disulfide bond proceeds through a species with a conformation very similar to that of the fully unfolded one and that the CL fragment undergoes global unfolding transition even in water.  相似文献   

19.
The steroidogenic acute regulatory protein (StAR) simulates steroid biosynthesis by increasing the flow of cholesterol from the outer mitochondrial membrane (OMM) to the inner membrane. StAR acts exclusively on the OMM, and only StAR's carboxyl-terminal alpha-helix (C-helix) interacts with membranes. Biophysical studies have shown that StAR becomes a molten globule at acidic pH, but a physiologic role for this structural transition has been controversial. Molecular modeling shows that the C-helix, which forms the floor of the sterol-binding pocket, is stabilized by hydrogen bonding to adjacent loops. Molecular dynamics simulations show that protonation of the C-helix and adjacent loops facilitates opening and closing the sterol-binding pocket. Two disulfide mutants, S100C/S261C (SS) and D106C/A268C (DA), designed to limit the mobility of the C-helix but not disrupt overall conformation, were prepared in bacteria, and their correct folding and positioning of the disulfide bonds was confirmed. The SS mutant lost half, and the DA mutant lost all cholesterol binding capacity and steroidogenic activity with isolated mitochondria in vitro, but full binding and activity was restored to each mutant by disrupting the disulfide bonds with dithiothreitol. These data strongly support the model that StAR activity requires a pH-dependent molten globule transition on the OMM.  相似文献   

20.
R P Miller  R A Farley 《Biochemistry》1990,29(6):1524-1532
Previous studies of titratable (Na+ + K+)-ATPase sulfhydryl groups have indicated the presence of one disulfide bond per mole of holoenzyme. This single disulfide cross-link was assigned to the beta subunit on the basis of the difference between the number of titrated "free" sulfhydryl groups and the total number of titrated sulfhydryl groups for each subunit [Esmann, M. (1982) Biochim. Biophys. Acta 688, 251; Kawamura, M., & Nagano, K. (1984) Biochim. Biophys. Acta 694, 27]. In the present study, beta-subunit tryptic peptides containing disulfide cross-links were identified and purified by HPLC. Two new peptides were generated from each disulfide-bonded peptide by reduction with dithiothreitol, and the amino acid compositions of these reduced peptides were determined. The data demonstrate that there are three disulfide bonds in the native beta subunit: 125Cys-148Cys, 158Cys-174Cys, and 212Cys-275Cys. The number of disulfide bonds in the beta subunit was also estimated by titration of sulfhydryl groups with [14C]iodoacetamide. Six sulfhydryl groups were identified: two sulfhydryl groups were titrated without prior reduction, and four were identified only after reduction of the protein with dithiothreitol. These data, suggesting that the beta subunit contains two disulfide bonds, are inconsistent with the peptide isolation experiments, which directly identified three disulfide bonds in the beta subunit. This inconsistency was resolved by demonstrating that approximately 20% of each disulfide bond in the beta subunit was reduced prior to the start of the experiment, resulting in an underestimation of the number of disulfide-bonded sulfhydryl groups in the beta subunit from the titration experiments.(ABSTRACT TRUNCATED AT 250 WORDS)  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