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.
Sequence-specific assignments of the 1H-nuclear magnetic resonance (NMR) spectra of the cardiotoxins CTXIIa and CTXIIb from Naja mossambica mossambica were obtained using two-dimensional NMR experiments at 500 MHz and the independently determined amino acid sequences. Assignments were obtained from data at 25 degrees C and 45 degrees C for all but one backbone proton of the 60 residues in each protein. Complete or partial assignments are also reported for the side-chain protons. These assignments supercede those published previously for the toxin preparation VII2 [Hosur, R. V., Wider, G. & Wüthrich K. (1983) Eur. J. Biochem. 130, 497-508]. The 1H/2H-exchange kinetics were measured in 2H2O at 20 degrees C for the amide protons and the N-terminal amino group. These and additional NMR data enabled the determination of the secondary structure in aqueous solution, which is virtually identical in CTXIIa and CTXIIb. Both proteins contain a short double-stranded antiparallel beta-sheet comprising the residues 2-4 and 11-13, and a triple-stranded antiparallel beta-sheet consisting of the residues 20-26, 35-39, and 49-55. The two peripheral strands of the triple-stranded beta-structure were found to be connected by a right-handed cross-over, and the locations of several tight turns were also identified.  相似文献   

2.
When N-[14C] acetylbenzotriazole, presented here as a new agent for the acetylation of proteins, reacted at pH 8 and 25 degrees C with delta-chymotrypsin, 15 amino groups (the epsilon-amino groups of lysing residues and the alpha-amino terminus of half-cystine-1) and two phenolic groups (those of the two exposed tyrosine residues) were acetylated with respective pseudo first-order constants of 0.056 +/- 0.003 and 0.15 +/- 0.03 min(-1). Surprisingly, in contrast with the acetic anhydride reaction, the alpha-amino group of Ile-16 was found to be not acetylated as shown by N-terminus determination and activity measurements: the modified delta-chymotrypsin (or acetylated delta-chymotrypsin) was fully active after neutral dialysis. Only a transient inactivation due to the incorporation of one [14C] acetyl group per mole of catalytic site was observed. The kinetic constant found for reactivation at pH 8.5 was 0.315 +/- 0.005 min(-1) at 25 degrees C. The enzyme-catalyzed hydrolysis of N-acetylbenzotriazole was described by a k(cat) value of 0.093 +/- 0.005 min(-1) at pH 7 and 25 degrees C. Circular dichroism changes observed at 230 nm during the reaction at pH 8.5, of acetylated delta-chymotrypsin with N-acetylbenzotriazole indicated a total conversion of the amount of enzyme molecules which were in the 'inactive' or 'alkaline' conformation at this pH, into the 'active' or 'neutral' one. Benzotriazole alone was unable to induce such a conformational change. The rate constant of the reverse structural process from the 'neutral' to the 'alkaline' conformation was 0.32 +/- 0.02 min(-1): identical to that of the deacetylation of the catalytic site. Thus, the unusual lack of acetylation of Ile-16 alpha-amino group during delta-chymotrypsin treatment with N-acetylbenzotriazole is interpreted as a stabilization of the enzyme 'neutral' conformation where the Ile-16 alpha-amino group is buried, thus inaccessible to the reagent. The properties of the delta-chymotrypsin modification using N-acetylbenzotriazole led to practical uses: direct spectrophotometric titration of chymotrypsin operational normality at pH 7 and rapid preparation of acetylated delta-chymotrypsin. As a protein reagent, N-acetylbenzotriazole is particularly interesting because of its reactivity towards amino and phenolic groups of amino acid residues, its stability at acid pH, i.e., k(hydrolysis=7.38 X 10(-3) min(-1) at 25 degrees C [Ravaux et al. (1971) Tetrahedron Letters, 4013-4015] and its aromaticity, responsible for optical properties.  相似文献   

3.
Amino acid sequencing of a large number of chemical and enzymatic cleavage products of elongation factor 1 alpha purified from rabbit reticulocyte has identified seven post-translationally modified residues. Five of the modifications are methylations of lysine residues yielding dimethyllysine at residues 55 and 165 and trimethyllysine at residues 36, 79, and 318. The two remaining post-translational modifications involve the addition of ethanolamine to glutamic acid residues 301 and 374, as reported previously (Rosenberry, T. L., Krall, J. A., Dever, T. E., Haas, R., Louvard, D., and Merrick, W. C. (1989) J. Biol. Chem. 264, 7096-7099). Fast atom bombardment mass spectrometry and fast atom bombardment tandem mass spectrometry have been used to analyze peptides containing these modified residues. The analyses have determined that glycerylphosphorylethanolamine has been attached to the glutamic acid residues. An analysis of the amino acid sequence surrounding each of the three types of modification has indicated no similarities. Therefore, it seems likely that the modifying enzymes do not recognize a specific amino acid sequence but rather the three-dimensional presentation of either amino or carboxyl residues in the elongation factor 1 alpha structure.  相似文献   

4.
The solution structure of reactive-site hydrolyzed Cucurbita maxima trypsin inhibitor III (CMTI-III*) was investigated by two-dimensional proton nuclear magnetic resonance (2D NMR) spectroscopy. CMTI-III*, prepared by reacting CMTI-III with trypsin which cleaved the Arg5-Ile6 peptide bond, had the two fragments held together by a disulfide linkage. Sequence-specific 1H NMR resonance assignments were made for all the 29 amino acid residues of the protein. The secondary structure of CMTI-III*, as deduced from NOESY cross peaks and identification of slowly exchanging hydrogens, contains two turns (residues 8-12 and 24-27), a 3(10)-helix (residues 13-16), and a triple-stranded beta-sheet (residues 8-10, 29-27, and 21-25). This secondary structure is similar to that of CMTI-I [Holak, T. A., Gondol, D., Otlewski, J., & Wilusz, T. (1989) J. Mol. Biol. 210, 635-648], which has a Glu instead of a Lys at position 9. Sequential proton assignments were also made for the virgin inhibitor, CMTI-III, at pH 4.71, 30 degrees C. Comparison of backbone hydrogen chemical shifts of CMTI-III and CMTI-III* revealed significant changes for residues located far away from the reactive-site region as well as for those located near it, indicating tertiary structural changes that are transmitted through most of the 29 residues of the inhibitor protein. Many of these residues are functionally important in that they make contact with atoms of the enzyme in the trypsin-inhibitor complex, as revealed by X-ray crystallography [Bode, W., Greyling, H. J., Huber, R., Otlewski, J., & Wilusz, T. (1989) FEBS Lett. 242, 285-292].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A calorimetric titration method was used to study ADP binding to native myosin. Data were analyzed by assuming that the myosin molecule has n independent and identical sites for ADP binding. The enthalpy change (deltaH), the binding constant (K), and n were determined. In 0.5 M KCl, 0.01 M MgCl2, and 0.02 M Tris/HCl (pH 7.8), we found: at 0 degrees, deltaH = -57.1 +/- 3.2 kJ-mol-1, log K = 6.42 +/- 0.13, n = 1.49 +/- 0.07; at 12 degrees, deltaH = 73.1 +/- 3.2 kJ-mole-1, log K = 6.08 +/- 0.13, and n = 1.74 +/- 0.07. The average heat capacity change on ADP binding to myosin between 0 and 12 degrees is thus -1.4 +/- 0.4 kJ-mol-1-K-1. Reasonably consistent results were obtained at 25 degrees, suggesting ADP binding to myosin is as strongly exothermic as at lower temperatures, although further interpretation of this result seems unwarranted, mainly because of the instability of myosic at this temperature. The number of protons released on binding of ADP to myosin was determined in separate experiments. The value was 0.19 +/- 0.02 at both 0 and 12 degrees. The reaction of protons with Tris thus contributes about -9.5 kJ-mol-1 to the observed heat on ADP binding.  相似文献   

6.
Amino acid sequences were analyzed for two isoforms of myosin essential light chain, LC17a and LC17b [Hasegawa, Y., Ueno, H., Horie, K., & Morita, F. (1988) J. Biochem. 103, 15-18] from porcine aorta smooth muscle. Both LC17a and LC17b consisted of 150 amino acid residues and their N-terminal Cys residues were blocked by an acetyl group. The amino acid sequences of LC17a and LC17b were common from the N-terminal to Glu-141 and five amino acid substitutions were observed within the remaining C-terminal 9 residues. The amino acid sequences of LC17a and LC17b were identical to those deduced from the nucleotide sequences of bovine aortic cDNAs encoding the two isoforms [Lash, J. A., Helper, D.J., Klug, M., Nicolozakes, A.W., & Hathaway, D.R. (1990) Nucleic Acids Res. 18, 7176].  相似文献   

7.
The equilibrium and kinetic properties for the urea-induced unfolding of the alpha subunit of tryptophan synthase from Escherichia coli, Salmonella typhimurium, and five interspecies hybrids were compared to determine the role of protein folding in evolution. The parent proteins differ at 40 positions in the sequence of 268 amino acids, and the hybrids differ by up to 15 amino acids from the Escherichia coli alpha subunit. The results show that all the proteins follow the same folding mechanism and are consistent with a previously proposed hypothesis [Hollecker, M., & Creighton, T. E. (1983) J. Mol. Biol. 168, 409; Krebs, H., Schmid, F. X., & Jaenicke, R. (1983) J. Mol. Biol. 169, 619] that the folding mechanisms are conserved in homologous proteins. Analysis of the kinetic data suggests that the 15 positions at which the parent proteins differ in the amino folding unit, residues 1-188, do not play a role in a rate-limiting step in folding that has been previously identified as the association of the amino and carboxyl folding units [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T. S., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965]. One or more of the 25 positions at which the parent proteins differ in the carboxyl folding unit, residues 189-268, do appear to play a role in this same rate-limiting step.  相似文献   

8.
Equilibrium and kinetic studies on the folding of a series of amino acid replacements at position 211 in the alpha subunit of tryptophan synthase from Escherichia coli were performed in order to determine the role of this position in the rate-limiting step in folding. Previous studies [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965-2974] have shown that the rate-limiting step corresponds to the association/dissociation of the amino (residues 1-188) and carboxy (residues 189-268) folding units. In terms of the secondary structure, the amino folding unit consists of the first six strands and five alpha helices of this alpha/beta barrel protein. The carboxy folding unit comprises the remaining two strands and three alpha helices; position 211 is in strand 7. Replacement of the wild-type glycine at position 211 with serine, valine, and tryptophan at most alters the rate of dissociation of the folding units; association is not changed significantly. In contrast, glutamic acid and arginine dramatically decelerate and accelerate, respectively, both association and dissociation. The difference in effects is attributed to long-range electrostatic interactions for these charged side chains; steric effects and/or hydrogen bonding play lesser roles. When considered with previous data on replacements at other positions in the alpha subunit [Hurle, M. R., Tweedy, N. B., & Matthews, C. R. (1986) Biochemistry 25, 6356-6360], it is clear that beta strands 6 (in the amino folding unit) and 7 (in the carboxy folding unit and containing position 211) dock late in the folding process.  相似文献   

9.
Glucosyltransferases (GTFs) secreted by mutans streptococci and some other lactic acid bacteria catalyze glucan synthesis from sucrose, and possess a C-terminal glucan-binding domain (GBD) containing homologous, directly repeating units. We prepared a series of C-terminal truncated forms of the GBD of Streptococcus sobrinus GTF-I and studied their binding to dextran by isothermal titration calorimetry. The binding of all truncates was strongly exothermic. Their titration curves were analyzed assuming that the GBD recognizes and binds to a stretch of dextran chain, not to a whole dextran molecule. Both the number of glucose units constituting the dextran stretch (n) and the accompanying enthalpy change (DeltaH degrees ) are proportional to the molecular mass of the GBD truncate, with which the Gibbs energy change calculated by the relation DeltaG degrees = -RT ln K (R, the gas constant; T, the absolute temperature; K, the binding constant of a truncate for a dextran stretch of n glucose units) also increases linearly. For the full-length GBD (508 amino acid residues), n = 33.9, K = 4.88 x 10(7) M-1, and DeltaH degrees = -289 kJ mol-1 at 25 degrees C. These results suggest that identical, independent glucose-binding subsites, each comprising 14 amino acid residues on average, are arranged consecutively from the GBD N-terminus. Thus, the GBD binds tightly to a stretch of dextran chain through the adding up of individually weak subsite/glucose interactions. Furthermore, the entropy change accompanying the GBD/dextran interaction as given by the relation DeltaS degrees = (DeltaG degrees - DeltaH degrees)/T has a very large negative value, probably because of a loss of the conformational freedom of dextran and GBD after binding.  相似文献   

10.
Liu X  Chan K  Chu IK  Li J 《Carbohydrate research》2008,343(17):2870-2877
Nonspecific proteolytic digestion of glycoproteins is an established technique in glycomics and glycoproteomics. In the presence of pronase E, for example, glycoproteins are digested to small glycopeptides having one to six amino acids residues, which can be analyzed with excellent sensitivity using mass spectrometry. Unfortunately, the long digestion times (1-3 days) limit the analytical throughput. In this study, we used controlled microwave irradiation to accelerate the proteolytic cleavage of glycoproteins mediated by pronase E. We used ESI-MS and MALDI-MS analyses to evaluate the microwave-assisted enzymatic digestions at various digestion durations, temperatures, and enzyme-to-protein ratios. When digesting glycoproteins, pronase E produced glycopeptides within 5 min under microwave irradiation; glycopeptides having one or two amino acids were the major products. Although analysis of peptides containing multiple amino acid residues offers the opportunity for peptide sequencing and provides information regarding the sites of glycosylation, the signals of Asn-linked glycans were often suppressed by the glycopeptides containing basic amino acids (Lys or Arg) in MALDI-MS experiments. To minimize this signal-to-content dependence, we converted the glycopeptides into their sodiated forms and then methylated them using methyl iodide. This controlled methylation procedure resulted in quaternization of the amino group of the N-terminal amino acid residue. Using this approach, the mass spectrometric response of glyco-Asn was enhanced, compensating for the poorer ionization efficiency associated with the basic amino acids residues. The methylated products of glycopeptides containing two or more amino acid residues were more stable than those containing only a single Asn residue. This feature can be used to elucidate glycan structures and glycosylation sites without the need for MS/MS analysis.  相似文献   

11.
J W Nelson  N R Kallenbach 《Biochemistry》1989,28(12):5256-5261
alpha-Helix formation in the S-peptide (residues 1-19 of ribonuclease A) was studied in detail by use of two-dimensional 1H nuclear magnetic resonance to monitor the effects of 2,2,2-trifluoroethanol (TFE) at 0 degrees C and pH* 2.07. TFE stabilizes the S-peptide alpha-helix. Helix formation by a particular amino acid was monitored by the chemical shifts of the C alpha, C beta, and C gamma protons while increasing the concentration of TFE: large changes in chemical shift of a particular residue indicate that it is induced to go helical, whereas small chemical shift changes indicate little helix formation. Residues Thr-3 to Met-13 undergo chemical shift changes consistent with helix formation, whereas the other residues do not. Earlier work [Kim, P. S., & Baldwin, R. L. (1984) Nature 307, 329-334] reported that residues Thr-3 to His-12 become helical in aqueous solution. The existence of a "helix stop signal" was inferred from this behavior. We thus conclude that this helix stop signal persists in TFE solutions.  相似文献   

12.
Plasma desorption mass spectrometry (PD-MS) was used to assess the molecular weight heterogeneity of glycopeptides (6-12 amino acids) from each of the three N-linked glycosylation sites of bovine fetuin (R.G. Spiro (1962) J. Biol. Chem. 237, 382-388). The glycopeptides were purified by a combination of anion exchange chromatography and reverse-phase HPLC. Since no detectable fragmentation was observed in the PD-MS of these asialoglycopeptides, the observation of multiple molecular ions could be attributed to either carbohydrate or peptide heterogeneity. Assignment of molecular ions, within 3 to 5 amu of the theoretical mass, of glycopeptides from each glycosylation site was made from amino acid composition, peptide sequence around the glycosylation sites, and previously reported triantennary oligosaccharide structures (B. Nilsson, N.E. Nordén, and S. Svensson (1979) J. Biol. Chem. 254, 4545-4553). Ion groups differing in mass by one N-acetyllactosamine unit were observed in glycopeptides from the Asn-Asp and Asn-Cys sites, localizing these previously observed biantennary oligosaccharide structures (R.R. Townsend, M.R. Hardy, T.C. Wong, and Y.C. Lee (1986) Biochemistry 25, 5716-5725; S. Takasaki and A. Kobata (1986) Biochemistry 25, 5709-5715) to these two sites. The presence of biantennary oligosaccharides at the Asn-Asp sites could be substantiated using 1H NMR but were not detected in the Asn-Cys glycopeptides. PD-MS was also implemented in the purification protocol for these glycopeptides and proved to be useful in assessing purity of chromatographic fractions which were mixtures of glycopeptides displaying both carbohydrate and peptide heterogeneity. A preparation scheme was developed to obtain molecular ions of desialylated glycopeptides by PD-MS.  相似文献   

13.
The purification and characterization of six isoforms of ornatin, potent glycoprotein IIb-IIIa (GP IIb-IIIa) antagonists and platelet aggregation inhibitors are described. These isoforms were purified from whole leech homogenates of the leech Placobdella ornata, a North American leech commonly known as the turtle leech, by trichloroacetic acid precipitation, Sephadex G-50 size exclusion chromatography, GP IIb-IIIa affinity chromatography, and C18 reverse-phase HPLC. Each of the five completely sequenced isoforms, which range from 41 to 52 residues in length, contains the Arg-Gly-Asp (RGD) sequence, a common recognition sequence in adhesion proteins, as well as 6 cysteine residues; the positions of both of these features are conserved in the primary sequences. The amino acid sequences of ornatin isoforms B, C, D, and E are highly conserved, whereas ornatin A2 and A3 are less similar and lack 9 residues at the N-terminus. The ornatins are approximately 40% identical with decorsin, a GP IIb-IIIa antagonist isolated from the leech Macrobdella decora [Seymour, J. L., Henzel, W. J., Nevins, B., Stults, J. T. & Lazarus, R. A. (1990) J. Biol. Chem. 265, 10143-10147]; furthermore, the RGD sequence and 5 out of 6 cysteine residues are maintained in the same relative positions in both decorsin and ornatin. The ornatin isoforms do not exhibit significant similarity to any members of the snake-venom-derived family of GP IIb-IIIa antagonists [Dennis, M. S., Henzel, W. J., Pitti, R. M., Lipari, M. T., Napier, M. A., Deisher, T. A., Bunting, S. & Lazarus, R. A. (1990) Proc. Natl Acad. Sci. USA 87, 2471-2475] except in the RGD region of these proteins. The ornatin isoforms inhibit the binding of GP IIb-IIIa to immobilized fibrinogen with IC50 values ranging over 2.9-5.3 nM; ornatin isoforms A2, C, and E inhibit ADP-induced human platelet aggregation with IC50 values of about 130, 280, and 440 nM, respectively.  相似文献   

14.
The recent identification of the alpha-subunit of mitochondrial F1-ATPase complex in rat liver peroxisomes suggests another functional role for this protein in both organelles in addition to its involvement in mitochondrial oxidative phosphorylation. We report here that a very rapid response (15 min) in the induction of the alpha-regulatory subunit of the mitochondrial F1-ATPase complex is observed in 37 degrees C heat-shocked larvae of Drosophila hydei. Under the same heat-shock treatment, a similar-fold induction for the heat-shock protein hsp-70 was less rapid (45 min). Although the amino acid sequence identities between the "chaperonine" and the alpha-subunit protein families are very low (less than 20%), two amino acid sequences, of 12 and 13 residues each, are found in the alpha-subunits of the F1-ATPase complex from various eukaryotes which show a highly conserved identity (over 50%) with amino acid sequences found in molecular chaperones. We suggest that the nuclear coded alpha-subunit belongs to the family of stress proteins hsp-60 and thus, that it could perform similar functional role(s) to those recently described for mitochondrial hsp-60 (Cheng, M. Y., Hartl, F. U., Martin, J., Pollock, R. A., Kalousek, F., Neupert, W., Hallberg, E. M., Hallberg, R. L., and Horwich, A. L. (1989) Nature 337, 620-625 and Ostermann, J., Horwich, A. L., Neupert, W., and Ultrich-Hartl, F. (1989) Nature 341, 125-130) in both the mitochondria and the peroxisomes. Furthermore, we suggest that the two conserved elements among the chaperonines and the alpha-subunits could putatively be involved in the chaperonine function of these proteins.  相似文献   

15.
R T Lee  T C Wong  R Lee  L Yue  Y C Lee 《Biochemistry》1989,28(4):1856-1861
A heterobifunctional linking reagent containing a masked aldehydo group and acyl hydrazide was synthesized for coupling of glycopeptides and other amino-containing compounds to proteins. After conversion to acyl azide, the reagent reacts with the amino group of a glycopeptide, and the modified glycopeptide is deacetalized with a weak acid to unmask the aldehydo group, which is then conjugated to bovine serum albumin (BSA) by reductive alkylation with pyridine-borane. The overall reaction scheme proceeds under relatively mild conditions. When the protein amino group was in a large excess (greater than 6-fold) of the aldehyde reagent, the efficiency of conjugation was as high as 88% even at submicromole levels. As a test case for application of this reagent, 6-aminohexyl beta-D-galactopyranoside (Gal-AH) was attached to the linking reagent and conjugated to BSA at various aldehyde-to-protein molar ratios ranging from 25 to 200. The level of O-galactosyl residue incorporated into BSA by this reagent far exceeded that observed in a similar reductive alkylation involving S-galactoside reagents [Lee, R. T., & Lee, Y. C. (1980) Biochemistry 19, 156-163]. By use of the present conjugating procedure, as many as 112 mol of Gal-AH residues were incorporated per mole of BSA, which represents near total modification of the amino groups. Some binding characteristics of the new BSA derivatives were studied in the mammalian hepatic galactose/N-acetylgalactosamine specific lectin system along with other types of BSA derivatives (containing S-galactosyl residues). In general, the behavior of the new derivatives was similar to that of other types. For instance, the affinity increased exponentially at low sugar substitution levels (up to 30 mol of galactosyl residues/mol of BSA), and the slope of exponential increase and affinity at a given sugar substitution level was similar to those of other types.  相似文献   

16.
Halobacterial flagellins are sulfated glycoproteins   总被引:21,自引:0,他引:21  
The cell-surface glycoprotein of Halobacteria contains oligosaccharides of the type Glc4----1GlcA4----1GlcA4----1GlcA (where GlcA indicates glucuronic acid) with a sulfate group attached to each of the GlcA residues. We report here that in addition to this cell-surface glycoprotein, the halobacterial flagellar proteins (recently described by Alam, M., and Oesterhelt, D. (1984) J. Mol. Biol. 176, 459-475) also contain the same type of sulfated oligosaccharides. These flagellins have the following features. All of the individual flagellar proteins contain identical sulfated saccharide moieties linked to the amido nitrogen of Asn through a Glc residue (the novel type of N-glycosidic linkage that has been found in the cell-surface glycoprotein from Halobacteria (Wieland, F., Heitzer, R., and Schaefer, W. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 5470-5474)). The amino acid sequence of one carbohydrate-binding region is Gln-Ala-Ala-Gly-Ala-Asp-Asn-Jle-Asn-Leu-Thr-Lys. This surrounding sequence CHO is consistent with the general formula Asn-X-Thr(Ser), common to all N-linked glycopeptides determined so far. Biosynthesis of flagellar glycoconjugates involved sulfated oligosaccharides linked to dolichol monophosphate. The individual glycoproteins making up the flagella are structurally closely related to one another.  相似文献   

17.
The effects of four single amino acid replacements on the stability and folding of the alpha subunit of tryptophan synthase from Escherichia coli have been investigated by ultraviolet differences spectroscopy. In previous studies [Miles, E. W., Yutani, K., & Ogasahara, K. (1982) Biochemistry 21, 2586], it had been shown that the urea-induced unfolding at pH 7.8, 25 degrees C, proceeds by the initial unfolding of the less stable carboxyl domain (residues 189-268) followed by the unfolding of the more stable amino domain (residues 1-188). The effects of the Phe-22----Leu, Glu-49----Met, Gly-234----Asp, and Gly-234----Lys mutants on the equilibrium unfolding process can all be understood in terms of the domain unfolding model. With the exception of the Glu-49----Met replacement, the effects on stability are small. In contrast, the effects of three of the four mutations on the kinetics of interconversion of the native form and one of the stable partially folded intermediates are dramatic. The results for the Phe-22----Leu and Gly-234----Asp mutations indicate that these residues play a key role in the rate-limiting step. The Glu-49----Met mutation increases the stability of the native form with respect to that of the intermediate but does not affect the rate-limiting step. The Gly-234----Lys mutation does not affect either the stability or the kinetics of folding for the transition between native and intermediate forms. The changes in stability calculated from the unfolding and refolding rate constants agree quantitatively with those obtained from the equilibrium data. When considered with the results from a previous study on the Gly-211----Glu replacement [Matthews, C. R., Crisanti, M. M., Manz, J. T., & Gepner G. L. (1983) Biochemistry 22, 1445], it can be concluded that the rate-limiting step in the conversion of the intermediate to the native conformation involves either domain association or some other type of molecule-wide phenomenon.  相似文献   

18.
NMR titration curves are reported for the 4 histidine residues of ribonuclease A in sodium acetate and for ribonuclease S in sodium acetate, phosphate, and sulfate solutions. Evidence is presented that the imidazole side chain of histidine residue 48 undergoes a conformational change, probably also involving the carboxyl side chain of aspartic acid residue 14. This group is considered to be responsible for the low pH inflection with pKa 4.2 present in the NMR titration curve of the C-2 proton resonance of histidine 48. The NMR titration curves of the active site histidine residues 12 and 119 also exhibit inflections at low pH values, although there is no carboxyl group within 9 A of the imidazole side chain of histidine residue 12 in the structure of ribonuclease S determined by x-ray crystallography (Wyckoff, H. W., Tsernoglou, D., Hanson, A. W. Knox, J. R., Lee, B., and Richards, F. M. (1970) J. Biol. Chem. 245, 305-328). Curve fitting was carried out on 11 sets of NMR titration data using a model in which the 3 histidine residues 12, 119, and 48 are assumed to be affected by a common carboxyl group. The results obtained indicate that such a model with fewer parameters gives as good a representation of the data as the model in which each histidine residue is assumed to interact separately with a different carboxyl group. Therefore, it is concluded that the ionization of aspartic acid residue 14 is indirectly experienced by the active site histidine residues through the conformational change at histidine 48. A model assuming mutual interaction of the active site histidine residues does not account for the low pH inflections in these curves.  相似文献   

19.
Energetically favored conformations of glycopeptide 1 were calculated using the newly developed force-field program, GEGOP (geometry of glycopeptides). The three-dimensional structure of glycopeptide 1, which is part of the Fc fragment of IgG1, has been calculated. 1 contains 27 amino acid residues from Pro291 to Lys317 and a biantennary decasaccharide N-linked to Asn297. The conformations of the peptide and the carbohydrate parts are shown to be mutually dependent. Single glycosyl residues of 1 exhibit interaction energies of up to -31.8 kJ/mol with the peptide portion. Generally, only a few of the glycosyl residues of the oligosaccharide moiety express significant interaction energies with the peptide part. No easy prediction is possible of glycosyl residues which exhibit favorable interaction energies. However, in all of the calculated structures, the glycosyl residues of the 1-6-linked branches show strong attractive forces for the peptide part. 1-6-glycosidically linked branches can adopt a larger number of conformations than other linkages due to their high flexibility which allows more favorable interactions with proteins. We developed the GEGOP program in order to be able to study the preferred conformations of large glycopeptides. The program is based on the GESA (geometry of saccharides) program and utilizes the HSEA (hard sphere exo anomeric) force field for the carbohydrate part and the ECEPP/2 (empirical conformation energy program for peptides) force field [Némethy, G., Pottle, M. S. & Scheraga, H. A. (1983) J. Phys. Chem. 87, 1883-1887] for the peptide part. The GEGOP program allows the simultaneous relaxation of all rotational degrees of freedom of these glycoconjugates during the energy optimization process. Thus, mutual interactions between glycosyl and amino acid residues can be studied in detail.  相似文献   

20.
D L Blithe  C A Buck  L Warren 《Biochemistry》1980,19(14):3386-3395
Glucosamine-labeled glycopeptides from control and virus-transformed BHK fibroblasts were characterized by size, lectin affinity, charge, and composition. As already demonstrated, on the basis of elution position on a column of Sephadex G-50, transformed cells contained a greater proportion of large glycopeptides than did control cells. Transformed cells also contained a larger proportion of glycopeptides which do not bind to Con A-Sepharose. By sequential chromatography on Sephadex G-50, Con A-Sepharose, and DEAE-Sephadex, approximately 40 individual peaks were partially or completely resolved. If sialic acid was removed from the glycopeptides prior to analysis by ion-exchange chromatography, 95% of the glycopeptides from control cells and 85% of the glycopeptides from transformed cells were no longer bound by DEAE-Sephadex. It was concluded that the DEAE-Sephadex elution properties of the glycopeptides are determined almost entirely by the sialic acid content of the molecules. A comparison of the profiles of control and transformed cell glycopeptides simultaneously eluting from columns of DEAE-Sephadex revealed that the differences between the two cells were largely quantitative; however, the possibility of the existence of qualitative differences as well cannot be excluded. In particular, there was one component present on the surface of transformed cells that was virtually absent in control cells. It was degraded by nitrous acid hydrolysis and heparinase and appeared to be heparan sulfate like material. After fractionation, each isolated glycopeptide population was analyzed for carbohydrate and, in some cases, amino acid content. The apparently larger glycopeptides, group A, the dominant population in transformed cells, were found to contain 3 to 4 mannose residues/glycopeptide when the sugars were normalized to sialic acid content. On the basis of the same criteria, group B glycopeptides contained 4-6 mannose residues/glycopeptide. The carbohydrate and amino acid compositions of the glycopeptides from transformed cells were, with a few exceptions, similar to those from control cells. Some isolated glycopeptides appeared to contain both O-glycosidic anad N-glycosidic linkages on the same oligopeptide.  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