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.
Amide-hydrogen exchange of three anti-yeast iso-1-cytochrome-c IgG monoclonal antibodies and the Fab, prepared from one of them, were studied by infrared spectrophotometry in the presence and absence of the deuterated immunogen and evolutionarily related species (the deuterated immunogen contained a population of a dimer. Each subunit of the dimer appeared to bind to the antibodies in a manner similar to the monomer). The number of hydrogens of the antibodies whose exchange was suppressed on binding to the immunogen was found to exceed that estimated for the residues shielded by the immunogen. Analysis of the data suggests that such suppression of hydrogen exchange occurs mainly for the Fab domains, but not for the Fc. One of the antibodies showed two distinct classes of amide-hydrogens. Class-1 hydrogens (approx. 36/site) exchange faster than class 2 (approx. 37/site). The exchange of class-1 hydrogens was suppressed by binding to the immunogen, but not to the evolutionarily related species. The exchange of class-2 hydrogens was suppressed by binding to the evolutionarily related species, as well as to the immunogen. Thus, the suppression of exchange of class-1 hydrogens appears to occur by some kind of conformational stabilization, the mechanism of which differentiates between the deuterated immunogen and the evolutionarily related species. Evidence suggests that the trans-interactions of the Fab domains may modulate the hydrogen exchange. If it is assumed that the antigen-binding strengthens the trans-interactions in such a way that the exchange of the slower exchanging hydrogens is suppressed, this could explain the suppression of exchange of class-2 hydrogens.  相似文献   

2.
Hydrogen-deuterium exchange in 2H20 solutions of the two redox states of horse heart cytochrome c was investigated at 20 degrees C, pH 7, by mass spectrometry and infrared spectroscopy. Mass spectrometry indicates that ferricytochrome has 20 hydrogens unexchanged after 24 h, 28 hydrogens exchanging between 10 min and 24 h, and 156 hydrogens exchanging within 10 min; comparative values for ferrocytochrome are 45, 19 and 140. The displacement of the exchange curves obtained by infrared corresponds to 8 to 9 peptide hydrogens. These combined methods show many non-peptide hydrogens exchanging rapidly (87 and 79 for ferricytochrome c and ferrocytochrome c respectively), whereas others, probably buried inside the molecule and involved in hydrogen bonds, are not exchanged, even after 24 h (14 and 30 hydrogens respectively, which is relatively large for a small protein). Infrared results are given in terms of changes of standard free energy for the transconformational reaction which exposes the peptide hydrogens to solvent: in ferricytochrome c and ferrycoytochrome c, 30% and 40% respectively of the peptide hydrogens are protected by conformational transitions stabilized by more than 5 kcal/mol (21 kJ/mol), which implies a large increase in rigidity for the reduced form.  相似文献   

3.
The effects of various cations and cosolvents on the conformation of wheat germ ribosomes and ribosomal subunits have been investigated by using the techniques of circular dichroism and differential hydrogen exchange. A class of hydrogens on both the 40S and 60S subunits exchange out more rapidly as the Mg2+ concentration is raised, indicating that Mg2+ causes a change in subunit conformation. Ca2+ and the polyamines produce a similar increase in the rate of hydrogen exchange. These results suggest that increases in cation concentrations permit a tightening of ribosome structure and a greater degree of internalization of the rRNA. The cosolvent glycerol causes an alteration in the CD spectrum of 80S ribosomes in both the nucleic acid and protein portions of the spectrum. Glycerol also causes a decrease in the rate of exchange of a number of hydrogens on both the 40S and 60S subunits. These results are interpreted to mean that glycerol favors a more native, less denatured structure in the ribosome.  相似文献   

4.
To understand further the weak molecular interactions between inhaled anesthetics and proteins, we studied the character and dynamic consequences of halothane binding to bovine serum albumin (BSA) and myoglobin using photoaffinity labeling and hydrogen-tritium exchange (HX). We find that halothane binds saturably and with submillimolar affinity to BSA, but either nonspecifically or with considerably lower affinity to myoglobin. Titration of halothane binding with guanidine hydrochloride suggested more protection of binding sites from solvent in BSA as compared with myoglobin. Protection factors for slowly exchanging albumin hydrogens are increased in a concentration-dependent manner by up to 27-fold with 10 mM halothane, whereas more rapidly exchanging groups of albumin hydrogens have either unaltered or decreased protection factors. Protection factors for slowly exchanging hydrogens in myoglobin are decreased by halothane, suggesting destabilization through binding to an intermediate or completely unfolded conformer. These results demonstrate the conformation dependence of halothane binding and clear dynamic consequences that correlate with the character of binding in these model proteins. Preferential binding and stabilization of different conformational states may underlie anesthetic-induced protein dysfunction, as well as provide an explanation for heterogeneity of action.  相似文献   

5.
An automated method for the optimal placement of polar hydrogens in a protein structure is described. This method treats the polar, side chain hydrogens of lysine, serine, threonine, and tyrosine and the amino terminus of a protein. The program, called NETWORK, divides the potential hydrogen-bonding pairs of a protein into groups of interacting donors and acceptors. A search is conducted on each of the local groups to find an arrangement which forms the most hydrogen bonds. If two or more arrangements have the same number of hydrogen bonds, the arrangement with the shortest set of hydrogen bonds is selected. The polar hydrogens of the histidyl side chain are specifically treated, and the ionization state of this residue is allowed to change, if this change results in additional hydrogen bonds for the local group. The program will accept Protein Data Bank as well as Biosym-format coordinate files. Input and output routines can be easily modified to accept other coordinate file formats. The predictions from this method are compared to known hydrogen positions for bovine pancreatic trypsin inhibitor, insulin, RNase-A, and trypsin for which the neutron diffraction structures have been determined. The usefulness of this program is further demonstrated by a comparison of molecular dynamics simulations for the enzyme cytochrome P-450cam with and without using NETWORK.  相似文献   

6.
S Segawa  K Kume 《Biopolymers》1986,25(10):1981-1996
The hydrogen-exchange reactions of peptide NH groups in lysozyme were studied by the change in the intensity of the amide II band in the ir spectrum. The slowest exchanging hydrogens, which are involved in intramolecular hydrogen bonding, are further divided into two groups at lower temperatures; half of them are exchanged through local unfolding and the other half through major cooperative unfolding. In order to study the correlation of the change in hydrogen-exchange rates with the change in the unfolding rate constant, we observed the effects of intrachain cross-linking, the addition of denaturant and ligand binding on the exchange rates through local unfolding. Although the exchange rate through major unfolding is greatly decreased by intrachain cross-linking between Glu 35 and Trp 108 (1/22000), the exchange rate through local unfolding is only slightly decreased (1/20). Even at higher temperatures, where most intact lysozyme molecules unfold, the folded conformation of cross-linked lysozyme remains compact, and no intermediate exists in which many side-chain atoms are packed loosely so that the hydrogen-exchange reaction occurs rapidly. Neither the addition of 2-PrOD molecules nor (NAG)3 binding affects the exchange rates through local unfolding. Our experiments confirm that the change in the unfolding rate constant does not correlate with the change in fluctuations in the relatively flexible hydrogen-bonded structure through which the exchange of peptide hydrogens takes place.  相似文献   

7.
The membrane-associated pulmonary surfactant protein C (SP-C), containing a polyvaline alpha-helix, and a synthetic SP-C analogue with a polyleucine helix (SP-C(Leu)) were studied by hydrogen/deuterium exchange matrix-assisted laser desorption ionization (MALDI) mass spectrometry. SP-C, but not SP-C(Leu), formed abundant amyloid fibrils under experimental conditions. In CD(3)OD/D(2)O, 91:9 (v/v), containing 2 mM ammonium acetate, SP-C(Leu) and SP-C exchanged 40% of their exchangeable hydrogens within 1 min. This corresponds to exchange of labile side-chain hydrogen atoms, hydrogens on the N- and C-terminal heteroatoms, and amide hydrogen atoms in the unstructured N-terminal regions. After approximately 300 h, four exchangeable hydrogen atoms in SP-C(Leu) and 10 in SP-C remained unexchanged. During this time period the ion current corresponding to singly charged SP-C decreased to <10% of the initial value due to the formation of insoluble aggregates that are not detected by MALDI mass spectrometry. In contrast, the ion current for SP-C(Leu) was maintained over this time period, although the peptides were incubated together. In combination, hydrogen/deuterium exchange and aggregation data indicate that the polyleucine peptide refolds into a helix after opening, while the unfolded polyvaline peptide forms insoluble beta-sheet aggregates rather than refolding into a helix. The SP-C helix, but not the SP-C(Leu) helix, is thus in a metastable state, which may contribute to the recently observed tendency of SP-C and its precursor to misfold and aggregate in vivo.  相似文献   

8.
To investigate the determinants of protein hydrogen exchange (HX), HX rates of most of the backbone amide hydrogens of Staphylococcal nuclease were measured by NMR methods. A modified analysis was used to improve accuracy for the faster hydrogens. HX rates of both near surface and well buried hydrogens are spread over more than 7 orders of magnitude. These results were compared with previous hypotheses for HX rate determination. Contrary to a common assumption, proximity to the surface of the native protein does not usually produce fast exchange. The slow HX rates for unprotected surface hydrogens are not well explained by local electrostatic field. The ability of buried hydrogens to exchange is not explained by a solvent penetration mechanism. The exchange rates of structurally protected hydrogens are not well predicted by algorithms that depend only on local interactions or only on transient unfolding reactions. These observations identify some of the present difficulties of HX rate prediction and suggest the need for returning to a detailed hydrogen by hydrogen analysis to examine the bases of structure-rate relationships, as described in the companion paper (Skinner et al., Protein Sci 2012;21:996-1005).  相似文献   

9.
Hydrogen exchange kinetic behavior of human erythrocyte glucose transporter protein in vesicles was studied in the absence and in the presence of D-glucose or a well known inhibitor, cytochalasin B. This is to detect a proposed channel of water penetrating into the protein through which the sugar molecule passes and to monitor any conformational changes induced by the substrate or inhibitor. Analyses of the kinetic data revealed several classes of hydrogens which exchange with readily distinguishable rates. Of 660 hydrogens detected per transporter, approximately 30% exchanged with rates generally characterized as those of free amide hydrogens indicating they are interfaced to solvent water. Since the transporter is known to be embedded deep in the hydrophobic area of the membrane with minimum exposure to the outside of the membrane lipid bilayer, a significant portion of these free amide hydrogens must be at the purported channel rather than outside of the membrane. D-Glucose and cytochalasin B affected the exchange kinetics of these presumably channel-associated free amide hydrogens rather differently. D-Glucose reduced the apparent rate constants, but not the total number. Cytochalasin B on the other hand reduced the total number to one-half without significantly changing the apparent rate constants. The remaining 70% of the labeled hydrogens exchanged with much slower rates which vary 10-10,000-fold, indicating that they are internally structured peptide amide and side chain hydrogens. Both D-glucose and cytochalasin B further reduced the rates of these hydrogens, indicating a global stabilization of the protein structure.  相似文献   

10.
《Life sciences》1994,56(4):PL89-PL96
Hydroxylation reactions of aromatic compounds have been used to detect hydroxyl radicals produced by gamma irradiation and ultrasound. The present study investigated the suitability of terephthalic acid (THA) as a hydroxyl radical dosimeter for general use in biologically relevant reactions. Hydroxyl radicals were generated by: (1) irradiating, THA with a 254 nm ultraviolet; (2) irradiating with gamma rays from a cesium source; and (3) generating hydroxyl radicals with 1 mM H2O2 and 10 μM Cu+2. In each of the three experiments, a fluorescent product was generated which exhibited identical fluorescent excitation and emission spectra. THA is non-fluorescent, eliminating the problem of a high initial background. Because THA has four ring hydrogens, only one mon-hydroxylated isomer was formed. The hydrogen peroxide reaction was dependent on the presence of a metal and cupric ions were effective in enhancing the reaction. With a Cu+2 concentration of 10 μM, the reation was linear between 0–30 mM H2O2. Catalase abolished the reaction at a concentration of 100 μg/ml and the effects could still be observed at 10 ng/ml, consistent with the very high rate at which catalase destroys hydrogen peroxide. Tertbutyl- hydroperoxide did not generate any fluorescence in this system which makes THA a very specific detector of hydroxyl radicals.  相似文献   

11.
The effect of D2O on the kinetic parameters for the hydroperoxide-supported N-demethylation of N,N-dimethylaniline catalyzed by chloroperoxidase and horseradish peroxidase was investigated in order to assess the roles of exchangeable hydrogens in the demethylation reaction. The initial rate of the chloroperoxidase-catalyzed N-demethylation of N,N-dimethylaniline supported by ethyl hydroperoxide exhibited a pL optimum (where L denotes H or D) of 4.5 in both H2O and D2O. The solvent isotope effect on the initial rate of the chloroperoxidase-catalyzed demethylation reaction was independent of pL, suggesting that the solvent isotope effect is not due to a change in the pK of a rate-controlling ionization in D2O. The solvent isotope effect on the Vmax for the chloroperoxidase-catalyzed demethylation reaction was 3.66 +/- 0.62. In contrast, the solvent isotope effect on the Vmax for the horseradish peroxidase catalyzed demethylation reaction was approximately 1.5 with either ethyl hydroperoxide or hydrogen peroxide as the oxidant, indicating that the exchange of hydrogens in the enzyme and hydroperoxide for deuterium in D2O has little effect on the rate of the demethylation reaction. The solvent isotope effect on the Vmax/KM for ethyl hydroperoxide in the chloroperoxidase-catalyzed demethylation reaction was 8.82 +/- 1.57, indicating that the rate of chloroperoxidase compound I formation is substantially decreased in D2O. This isotope effect is suggested to arise from deuterium exchange of the hydroperoxide hydrogen and of active-site residues involved in compound I formation. A solvent isotope effect of 2.96 +/- 0.57 was observed on the Vmax/KM for N,N-dimethylaniline in the chloroperoxidase-catalyzed reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The kinetic properties of purified sheep hepatic pyruvate kinase change upon storage. Assayed at 0.5 mM fructose-1,6-diphosphate and 2 mM ADP, saturation of fresh enzyme with phosphoenolpyruvate is hyperbolic, with KPEP = 0.1 mM (pH 7.5, and 30 degrees C). Under similar conditions enzyme stored at -20 degrees C for 1 week or more yields a nonlinear Lineweaver-Burk plot for PEP. The data may be accounted for by the appearance of two enzymic forms with identical turnover numbers, but different KPEP (0.035 +/- 0.005 and 12.4 +/- 0.6 mM). Storage also increases the concentration of fructose-1,6-diphosphate required for maximal activation from nanomolar to millimolar levels. Assayed at 2 mM ADP and 2 mM PEP, the apparent KFDP is 10 mM. Preincubation of stored enzyme with PEP in the presence of mercaptoethanol leads to significant reversion to original kinetic properties. Available data suggest that the storage-dependent change in kinetic behavior rises from changes in subunit conformation and not from dissociation into subunits.  相似文献   

13.
An expression system has been developed that allows high levels of production of TEM-1 beta-lactamase with ease of biosynthetic incorporation of nuclear isotopes. The gene for mature TEM-1 beta-lactamase fused to the leader sequence of the ompA protein was subcloned into the pET-24a(+) vector by introduction of an NdeI restriction site at the first codon of the fused genes and transformed into Escherichia coli BL21 (DE3) cells. With protein induction at 25 degrees C supported by LB medium supplemented with osmolytes (300 mM sucrose and 2.5 mM betaine), the extracellular, mature form of wild-type TEM-1 beta-lactamase was recovered at a level of 140 mg/L. The production level of E166N, E240C, E104C, and M272C mutants depended on the mutation but was invariably higher than reported by others for expression systems of the wild-type enzyme. Comparison of different carbon sources on the efficiency of biosynthetic incorporation of covalent deuterium showed maximal (90%) incorporation with minimal medium containing 99% (2)H(2)O and sodium d(3)-acetate (99 atom% (2)H). The yield of deuterium-enriched wild-type enzyme was 80 mg/L with yields for mutants proportionally reduced. The high level of protein deuteration achieved with this system allowed detection of the hyperfine coupling between the paramagnetic nitroxyl group of a spin-labeled penicillin substrate and hydrogens on the penicillin moiety in a cryokinetically isolated acylenzyme reaction intermediate because of the decrease in overlapping resonances of active site residues. The overexpression system is readily adaptable for other target proteins and facilitates studies requiring large quantities of protein in isotopically enriched forms.  相似文献   

14.
The structure and 1H/2H exchange kinetics of affinity-purified nAChR reconstituted into egg phosphatidylcholine membranes with increasing levels of either dioleoylphosphatidic acid (DOPA) or cholesterol (Chol) have been examined using infrared spectroscopy. All spectra of the reconstituted nAChR membranes recorded after 72 h in 2H2O exhibit comparable amide I band shapes, suggesting a similar secondary structure for the nAChR in each lipid environment. Increasing levels of either DOPA or Chol, however, lead to an increasing intensity of the amide II band, indicating a decreasing proportion of nAChR peptide hydrogens that have exchanged for deuterium. Spectra recorded as a function of time after exposure of the nAChR to 2H2O show that the presence of either lipid slows down the 1H/2H exchange of those peptide hydrogens that normally exchange on the minutes to hours time scale. The slowing of peptide 1H/2H exchange correlates with both an increasing ability of the nAChR to undergo agonist-induced conformational change [Baenziger, J. E., Morris, M.-L., Darsaut, T. E., and Ryan, S. E. (1999) in preparation] and possibly a decreasing membrane fluidity. Our data suggest that lipid composition dependent changes in nAChR peptide 1H/2H exchange kinetics reflect altered internal dynamics of the nAChR. Lipids may influence protein function by changing the internal dynamics of integral membrane proteins.  相似文献   

15.
A procedure to measure exchange rates of fast exchanging protein amide hydrogens by time-resolved NMR spectroscopy following in situ initiation of the reaction by diluting a native protein solution into an exchanging deuterated buffer is described. The method has been used to measure exchange rates of a small set of amide hydrogens of reduced cytochrome c, maintained in a strictly anaerobic atmosphere, in the presence of an otherwise inaccessible range of guanidinium deuterochloride concentrations. The results for the measured protons indicate that hydrogen exchange in the unfolding transition region of cytochrome c reach the EX2 limit, but emphasize the difficulty in interpretation of the exchange mechanism in protein hydrogen exchange studies. Comparison of free energies of structure opening for the measured hydrogens with the global unfolding free energy monitored by far-UV CD measurements has indicated the presence of at least one partially unfolded equilibrium species of reduced cytochrome c. The results provide the first report of measurement of free energy of opening of structure to exchange in the 0–2-kcal/mol range. Proteins 32:241–247, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

16.
Freshly prepared human red blood cells incubated with 5 mM ferricyanide, 0.2 mM iodoacetate and 2 mM adenosine in the presence of 5 mM EGTA demonstrate comparable increases in Na+ and K+ permeability (ferricyanide effect). This effect is unrelated to the Ca2+-activated K+ channel (Gardos effect) since influx of Ca2+ from outside the cell is excluded. Also this effect is different from the non-specific Na+ and K+ permeability change elicited by PCMBS. These differences become obvious by using various reagents. For example, A23187 and quinidine exert opposite effects in Gardos and ferricyanide experiments, where A23187 and atebrin react oppositely in the latter and in PCMBS experiments. The ferricyanide effect described here does not involve formation of nonspecific channels. The change in Na+ permeability separately from K+ permeability under certain circumstances suggests a more specific effect.  相似文献   

17.
To determine the source(s) of blood and very low density lipoprotein (VLDL)-triglyceride glycerol during fasting, four men ingested (2)H(2)O from 14 to 20 h into a 60-h fast to achieve ~0.5% body water enrichment. At 60 h of fasting, glycerol flux was measured using [2-(14)C]glycerol. Blood was taken for measurement of (2)H enrichment at carbon 6 of glucose and at carbon 3 of free glycerol and VLDL-triglyceride glycerol. (2)H enrichment of the 2 hydrogens bound to carbon 3 of VLDL-triglyceride glycerol was 105 +/- 2% of the (2)H enrichment of the 2 hydrogens bound to carbon 6 of glucose, indicating isotopic equilibrium between hepatic glyceraldehyde 3-P and glycerol 3-P. The (2)H enrichment of the 2 hydrogens bound to carbon 3 of free glycerol was 17 +/- 3% of VLDL-triglyceride glycerol, indicating that a significant percentage of free glycerol in blood originated from the hydrolysis of circulating VLDL-triglyceride or a pool of glycerol with similar (2)H enrichment. Glycerol flux was 6.3 +/- 1.1 micromol. kg(-1). min(-1). Glycerol appearing from nonadipose tissue sources was then approximately 1.1 micromol. kg(-1). min(-1). Seven other subjects were fasted for 12, 42, and 60 h. A small percentage of glycerol in the circulation after 12 h of fasting was enriched with (2)H. The enrichment of the 2 hydrogens bound to carbon 3 of free glycerol in the longer periods of fasting was approximately 16% of the enrichment of the 2 hydrogens bound to carbon 6 of glucose. Therefore, as much as 15-20% of systemic glycerol turnover during fasting is not from lipolysis of adipose tissue triglyceride.  相似文献   

18.
The conformation of the N-linked complex glycopeptide of fetuin was examined with hydrogen-exchange techniques. The glycopeptide molecule contains eight acetamido hydrogens stemming from five N-acetylglucosamine residues and three N-acetylneuraminic acid residues and also one from the remaining sugar-peptide linkage. The hydrogen-exchange rates of these secondary amides were compared with small molecule model compounds having identical primary structures at their exchangeable hydrogen sites. Differences between the model rates and glycopeptide rates therefore cannot be accounted for by primary structure effects but reflect conformational features of the glycopeptide. Two glycopeptide hydrogens exhibit significantly hindered exchange; the rest exchange at the model rates. Removal of the three N-acetylneuraminic acid residues from terminal positions on the three branches of the glycopeptide removes the slowed hydrogens. The remaining ones continue to exchange at the model rate. These results indicate that two of the eight sugar acetamido hydrogens are involved in intramolecular hydrogen bonds. A likely structure includes two hydrogen bonds between the three N-acetylneuraminic acid residues. These two hydrogens, slowed to a moderate degree, reflect a preferred conformation stabilized by about 1 kcal/mol in free energy. The solution conformation of the glycopeptide suggested by these results is one that is partially ordered and can be easily modulated, owing to the relatively small amount of energy stabilizing the preferred conformation.  相似文献   

19.
Peptide NH resonances in the 250 MHZ 1H nuclear magnetic resonance (NMR) spectrum of oxytocin in H2O were assigned to specific amino acid residues by the "underwater decoupling" technique (i.e., decoupling from corresponding CalphaH resonances, which are buried beneath the intense water peak). These experiments confirm previous assignments of A. I. Brewster an V. J. Hruby ((1973), Proc. Natl. Acad. Sci. U.S.A. 70, 3806) and A. F. Bradbury et al. ((1974), FEBS Lett. 42, 179). Three methods of assigning NH resonances of peptides--solvent titration, underwater decoupling, and isotopic labeling--are compared. As the solvet composition is gradually changed from dimethyl sulfoxide to H2O, oxytocin undergoes a conformational change at 70-90 mol % of H2O. Exposure to solvent of specific hydrogens of oxytocin in H2O was studied by monitoring intensity changes of solute resonances when the solvent peak was saturated. Positive nuclear Overhauser effects (NOE's) of 14 +/- 5 were observed for the Tyr ortho CH and meta CH resonances, respectively. Comparative studies with deamino-oxytocin indicate that these effects result predominantly from intermolecular dipoledipole interaction between aromatic side chain CH protons and protons of the solvent. The NOE's therefore indicate intimate contact between water and the aromatic CH hydrogens of the Tyr side chain. The extent of saturation transferred by proton exchange between water and NH group varies with Ph in a manner which appears to reflect the acid-base catalysis of the protolysis reaction. There is no indication that any NH protons are substantially shiedled from the solvent.  相似文献   

20.
Quasielastic neutron scattering experiments performed on yeast phosphoglycerate kinase in the native form and denatured in 1.5 M guanidinium chloride reveal a change in the fast (picosecond time scale) diffusive internal dynamics of the protein. The momentum and energy transfer dependences of the scattering for both states are fitted by an analytical model in which, on the experimentally accessible picosecond time scale and angstrom length scale, the dynamics of a fraction of the nonexchangeable hydrogens in the protein is described as a superposition of vibrations with uniform diffusion in a sphere, the rest of the hydrogens undergoing only vibrational motion. The fraction diffusing changes, from ≈60% in the native protein to ≈82% in the denatured protein. The radius of the sphere also changes slightly, from ≈1.8 Å in the native protein to ≈2.2 Å in the denatured protein. Possible implications of these results for the general protein folding problem are discussed. Proteins 28:380–387, 1997 © 1997 Wiley-Liss, Inc.  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