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.
Raman spectroscopy was employed to examine the secondary structure of the cAMP receptor protein (CRP). Spectra were obtained over the range 400-1900 cm-1 from solutions of CRP and from CRP-cAMP cocrystals. The spectra of CRP dissolved in 30 mM sodium phosphate and 0.15 M NaCl buffered at either pH 6 or pH 8 or dissolved in 0.15-0.2 M NaCl at protein concentrations of 5, 15, and 30 mg/mL were examined. Estimates of the secondary structure distribution were made by analyzing the amide I region of the spectra (1630-1700 cm-1). CRP secondary structure distributions were essentially the same in either pH and at all protein concentrations examined. The amide I analyses indicated a structural distribution of 44% alpha-helix, 28% beta-strand, 18% turn, and 10% undefined for CRP in solution. Raman spectra of CRP-cAMP cocrystals differed from the spectra of CRP in solution. Some differences were assigned to interfering background bands, whereas other spectral differences were attributed to changes in CRP structure. Differences in the amide III region and in the intensity at 935 cm-1 were consistent with alterations in secondary structure. Analysis of the amide I region of the CRP-cAMP cocrystal spectrum indicated a secondary structure distribution of 37% alpha-helix, 33% beta-strand, 17% turn, and 12% undefined. This result is in agreement with a published secondary structure distribution derived from X-ray analysis of CRP-cAMP cocrystals (37% alpha-helix and 36% beta-strand).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Discoidal complexes have been prepared from 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and the apoproteins of HDL3 (apo HDL3) or purified apo A-I. Gel electrophoresis established that apo HDL3 contained 74% apo A-I. Deconvolution and curve-fitting of the infra-red amide I band of the apoprotein in the lipid-protein complex revealed a secondary structure containing approximately 40% alpha-helix and 50% beta-structure. This contrasted with the results from circular dichroism studies (Surewicz et al. (1986) J. Biol. Chem., 261, 16191) of apo A-I/DMPC complexes which predicted 68% alpha-helix and 7% beta-structure. The discrepancy between the two methods and limitations of the two techniques for lipoproteins is discussed.  相似文献   

3.
Anions interact with protein to induce structural changes at ligand binding sites. The effects of anion complexation include structural stabilization and promote cation-protein interaction. This study was designed to examine the interaction of aspirin and ascorbate anions with the Na+, K+-dependent adenosine triphosphatase (Na,K-ATPase) in H2O and D2O solutions at physiological pH, using anion concentrations of 0.1 microM to 1 mM with final protein concentration of 0.5 to 1 mg/ml. Absorption spectra and Fourier transform infrared (FTIR) difference spectroscopy with its self-deconvolution, second derivative resolution enhancement and curve-fitting procedures were applied to characterize the anion binding mode, binding constant, and the protein secondary structure in the anion-ATPase complexes. Spectroscopic evidence showed that the anion interaction is mainly through the polypeptide C=O and C-N groups with minor perturbation of the lipid moiety. Evidence for this came from major spectral changes (intensity variations) of the protein amide I and amide II vibrations at 1651 and 1550 cm(-1). respectively. The anion-ATPase binding constants were K=6.45 x 10(3) M(-1) for aspirin and K=1.04 x 10(4) M(-1) for ascorbate complexes. The anion interaction resulted in major protein secondary structural changes from that of the alpha-helix 19.8%; beta-pleated sheet 25.6%; turn 9.1%; beta-antiparallel 7.5% and random 38% in the free Na,K-ATPase to that of the alpha-helix 24-26%; beta-pleated 17-18%; turn 8%; beta-antiparallel 5-3% and random 45.0% in the anion-ATPase complexes.  相似文献   

4.
The A subunit of cholera toxin contains the ADP-ribosyltransferase activity in its major constituent polypeptide A1 (Mr 23,000) which is responsible for the elevation of cAMP typically observed with most mammalian cell types after exposure to the toxin. The primary structure of the A subunit, recently established by sequence analyses, is presented and used as the basis for the secondary structure prediction according to the method of Chou and Fasman. The results indicated the presence of 27% alpha-helix, 25% beta-structure, 12% beta-turn, and 36% random coil. The majority of the beta-structure consisted of six strands located in the NH2-terminal portion of the molecule (residues 33-106) covering one-half of the region corresponding to the A1 polypeptide portion. The beta-sheet domain led immediately into the active site region characterized by the alternating structures of beta-pleated sheet and alpha-helix (residues 95-140) similar to that reported for other NAD+ binding proteins. The presence of this structural feature in the region was confirmed by the use of another predictive method (J. Garnier et al., J. Mol. Biol. 1978, 120, 97-120). In addition, two regions (residues 14-18 and 200-214), previously identified to contain binding sites for the B subunit as evidenced by chemical modification and monoclonal antibody studies, were found to be in alpha-helix configuration.  相似文献   

5.
Infrared spectroscopy in the amide I region of purified membrane-bound Na,K-ATPase preparation shows that Na+- and K+-bound forms of the enzyme have almost the same secondary structure. No difference is detected in the beta-structure (pleated sheets) content. This is contrary to the statement of the recent paper (Gresalfi, T. J., and Wallace, B. A. (1984) J. Biol. Chem. 259, 2622-2628) where a similar preparation was examined by circular dichroism spectroscopy and it was claimed that net 7% of protein peptide groups undergo a beta-sheet to alpha-helix conformational change upon Na,K-ATPase conversion from the K+ to the Na+ form. The discrepancy of the results is most likely caused by the particulate nature of the enzyme preparations used that could lead to optical artifacts in CD but not in IR measurements. A thorough comparison of IR spectra of these enzyme forms has revealed a very minor spectral difference which could suggest conformational perturbations, if any, of a much lower scale and another type than that claimed by Gresalfi and Wallace. The K+ form tends to absorb slightly more in the region of the alpha-helix band. This could reflect some distortion or a transition to a random coil structure of a small fraction of alpha-helical segments (less than or equal to 2% protein peptide groups) upon the enzyme conversion from the K+ to the Na+ form.  相似文献   

6.
Elongation factor 1alpha from the hyperthermophilic archaeon Sulfolobus solfataricus (SsEF-1alpha) carries the aminoacyl tRNA to the ribosome; it binds GDP or GTP, and it is also endowed with an intrinsic GTPase activity that is triggered in vitro by NaCl at molar concentrations [Masullo, M., De Vendittis, E., and Bocchini, V. (1994) J. Biol. Chem. 269, 20376-20379]. The structural properties of SsEF-1alpha were investigated by Fourier transform infrared spectroscopy. The estimation of the secondary structure of the SsEF-1alpha*GDP complex, made by curve fitting of the amide I' band or by factor analysis of the amide I band, indicated a content of 34-36% alpha-helix, 35-40% beta-sheet, 14-19% turn, and 7% unordered structure. The substitution of the GDP bound with the slowly hydrolyzable GTP analogue Gpp(NH)p induced a slight increase in the alpha-helix and beta-sheet content. On the other hand, the alpha-helix content of the SsEF-1alpha*GDP complex increased upon addition of salts, and the highest effect was produced by 5 M NaCl. The thermal stability of the SsEF-1alpha*GDP complex was significantly reduced when the GDP was replaced with Gpp(NH)p or in the presence of NaBr or NH4Cl, whereas a lower destabilizing effect was provoked by NaCl and KCl. Therefore, the extent of the destabilizing effect of salts depended on the nature of both the cation and the anion. The data suggested that the sodium ion was responsible for the induction of the GTPase activity, whereas the anion modulated the enzymatic activity through destabilization of particular regions of SsEF-1alpha. Finally, the infrared data suggested that, in particular region(s) of the polypeptide chain, the SsEF-1alpha*Gpp(NH)p complex possesses structural conformations which are different from those present in the SsEF-1alpha*GDP complex.  相似文献   

7.
M G Zagorski  C J Barrow 《Biochemistry》1992,31(24):5621-5631
Beta-peptide is a major component of amyloid deposits in Alzheimer's disease. We report here a proton nuclear magnetic resonance (NMR) spectroscopic investigation of a synthetic peptide that is homologous to residues 1-28 of beta-peptide [abbreviated as beta-(1-28)]. The beta-(1-28) peptide produces insoluble beta-pleated sheet structures in vitro, similar to the beta-pleated sheet structures of beta-peptide in amyloid deposits in vivo. For peptide solutions in the millimolar range, in aqueous solution at pH 1-4 the beta-(1-28) peptide adopts a monomeric random coil structure, and at pH 4-7 the peptide rapidly precipitates from solution as an oligomeric beta-sheet structure, analogous to amyloid deposition in vivo. The NMR work shown here demonstrates that the beta-(1-28) peptide can adopt a monomeric alpha-helical conformation in aqueous trifluoroethanol solution at pH 1-4. Assignment of the complete proton NMR spectrum and the determination of the secondary structure were arrived at from interpretation of two-dimensional (2D) NMR data, primarily (1) nuclear Overhauser enhancement (NOE), (2) vicinal coupling constants between the amide (NH) and alpha H protons, and (3) temperature coefficients of the NH chemical shifts. The results show that at pH 1.0 and 10 degrees C the beta-(1-28) peptide adopts an alpha-helical structure that spans the entire primary sequence. With increasing temperature and pH, the alpha-helix unfolds to produce two alpha-helical segments from Ala2 to Asp7 and Tyr10 to Asn27. Further increases in temperature to 35 degrees C cause the Ala2-Asp7 section to become random coil, while the His13-Phe20 section stays alpha-helical. A mechanism involving unfavorable interactions between charged groups and the alpha-helix macrodipole is proposed for the alpha-helix----beta-sheet conversion observed at midrange pH.  相似文献   

8.
Three equations describing sedimentation equilibrium are examined and tested for their ability to analyze data. The testing procedure using simulated data is similar to that described previously (Holladay, L. A., and Sophianopoulos, A. J. (1972) J. Biol. Chem.247, 427–439) and used with another equation. The equations examined here are found to be of much less statistical reliability and of a more restricted range of application than the previously examined equation. The equation described previously, (Holladay, L. A., and Sophianopoulos, A. J. (1972) J. Biol. Chem.247, 427–439) is also used here to examine the conditions necessary to detect isodesmic systems of more than four components. The self-association of lysozyme reported previously (Sophianopoulos, A. J., and Van Holde, K. E. (1964) J. Biol. Chem.239, 2516–2524) is reexamined at pH 8.2, 0.15 ionic strength, and 13°C. The tentative conclusion is that the system is mainly a monomer-dimer, with a small, uncertain amount of tetramer possibly present. Under the above conditions the second virial coefficient, B, is estimated to lie in the range 0–4.4 × 10?6 mole·dl·g?2, the dimerization constant. K21, lies in the range 2.3–2.7 × 10?3m, and the tetramerdimer constant, K42, is in the range 1.5–15 × 10?3m.  相似文献   

9.
Secondary and tertiary structural parameters of type E botulinum neurotoxin in the unactivated single-chain and activated two-chain (i.e., after proteolytic cleavage) forms were analyzed using circular dichroism, derivative absorption and fluorescence spectroscopy. The estimated secondary structures (22 and 20% alpha-helix, 44 and 44% beta-pleated sheets, and 34 and 36% random coils for the single- and two-chain neurotoxins, respectively) indicated that virtually no change occurred upon nicking of the single-chain neurotoxin. About 57% of the 70 Tyr residues were exposed in the single-chain form, which increased to 62% in the two-chain form. Fluorescence quenching experiments with neutral, anionic and cationic quenchers indicated that about 40% of the maximum accessible fluorescent Trp residues were exposed on the surface of the single-chain neurotoxin as compared to only 20% in the case of the two-chain neurotoxin. Acrylamide was the most effective quencher with a fraction accessibility of 0.56 and 0.48 of maximum accessible Trp fluorescence residues in the single and two-chain forms of the neurotoxin, respectively. Native polyacrylamide gel electrophoresis of the two forms of the neurotoxin revealed greater mobility for the two chain form. This indicates that the surface charges in the single-chain neurotoxin were altered upon nicking. These observations suggest that nicking of the single-chain type E neurotoxin results in refolding and redistribution of the surface charges of the neurotoxin.  相似文献   

10.
Human airway lysozyme, purified from pathological bronchial secretions, is characterized by a specific activity 3-fold higher than that of hen egg-white lysozyme. The amino acid composition of human airway lysozyme is identical to that of other human lysozymes. The laser Raman spectra of human airway lysozyme and hen egg-white lysozyme in phosphate buffer solution (pH 7.2) are recorded in the range 300-1900 cm-1 at 488 nm. Drastic intensity differences are observed between the spectra analyzed in the ranges characteristic of the peptide backbone (e.g., beta-sheet; C alpha-C, C alpha-N), and of the aromatic side-chain vibrations (tyrosine, tryptophan). The deconvolution of the Raman amide I band gives secondary structures of 38% and 39% alpha-helix, 25% and 20% beta-sheet, and 37% and 41% undefined structure for the human and hen lysozymes, respectively.  相似文献   

11.
The present work describes the dynamics of the apo form of cytochrome b(562), a small soluble protein consisting of 106 amino acid residues [Itagaki, E., and Hager, L. P. (1966) J. Biol. Chem. 241, 3687-3695]. The presence of exchange in the millisecond time scale is demonstrated for the last part of helix IV (residues 95-105 in the holo form). The chemical shift index analysis [Wishart, D. S., and Sykes, B. D. (1994) J. Biomol. NMR 4, 171-180] based on H(alpha), C(alpha), C(beta), and C' chemical shifts suggests a larger helical content than shown in the NMR structure based on NOEs. These results indicate the presence of helical-like conformations participating in the exchange process. This hypothesis is consistent with amide deuterium exchange rates and the presence of some hydrogen bonds identified from amide chemical shift temperature coefficients [Baxter, N. J., and Williamson, M. P. (1997) J. Biomol. NMR 9, 359-369]. (15)N relaxation indicates limited mobility for the amide protons of this part of the helix in the picosecond time scale. A 30 ns stochastic dynamics simulation shows small fluctuations around the helical conformation on this time scale. These fluctuations, however, do not result in a significant decrease of the calculated order parameters which are consistent with the experimental (15)N relaxation data. These results resolve an apparent discrepancy in the NMR structures between the disorder observed in helix IV due to a lack of NOEs and the secondary structure predictions based on H(alpha) chemical shifts [Feng, Y., Wand, A. J., and Sligar, S. G. (1994) Struct. Biol. 1, 30-35].  相似文献   

12.
cis-Pt(NH(3))(2)Cl(2) (cisplatin) is an antitumor drug with many severe toxic side effects including enzymatic changes associated with its mechanism of action. This study was designed to examine the interaction of cisplatin drug with the Na(+), K(+)-dependent adenosine triphosphatase (Na,K-ATPase) in H(2)O and D(2)O solutions at physiological pH, using drug concentrations of 0.1 microM to 1 mM. UV absorption spectra and Fourier transform infrared difference spectroscopy with its self-deconvolution, second derivative resolution enhancement and curve-fitting procedures were applied to characterize the drug binding mode, the drug binding constant and the protein secondary structure in the cisplatin-ATPase complexes. Spectroscopic evidence showed that at low drug concentration (0.1 microM), cisplatin binds mainly to the lipid portion of the enzyme, whereas at higher drug contents, the Pt cation interaction is through the polypeptide C==O and C-N groups with overall binding constant of K=1.93 x 10(4) M(-1). At high cisplatin concentration (1 mM), drug binding results in protein secondary structural changes from that of the alpha-helix 19.8%; beta-pleated 25.6%; turn 9.1%; beta-antiparallel 7.5% and random 38%, in the free Na,K-ATPase to that of the alpha-helix 22.2%; beta-pleated 23.2%; turn 9.4%; beta-antiparallel 2.2% and random 43%, in the cis-Pt-ATPase complexes.  相似文献   

13.
The secondary structure of streptokinase (Sk) in aqueous solution was quantitatively examined by using Fourier transform infrared (FT-IR) spectroscopy. Resolution enhancement techniques, including Fourier deconvolution and derivative spectroscopy, were combined with band curve-fitting procedures to quantitate the spectral information from the amide I bands. Nine component bands were found under the broad, nearly featureless amide I bands which reflect the presence of various substructures. The relative areas of these component bands indicate an amount of beta-sheet between 30 and 37% and an alpha-helix content of only 12-13% in Sk. Further conformational substructures are assigned to turns (25-26%) and to "random" structures (15-16%). Additionally, the correlation of a pronounced component band near 1640 cm-1 (10-16% fractional area) with the possible presence of 3(10)-helices is discussed.  相似文献   

14.
15.
Normal human plasma alpha 2HS-glycoprotein has earlier been shown to be comprised of two polypeptide chains. Recently, the amino acid and carbohydrate sequences of the short chain were elucidated (Gejyo, F., Chang, J.-L., Bürgi, W., Schmid, K., Offner, G. D., Troxler, R.F., van Halbeck, H., Dorland, L., Gerwig, G. J., and Vliegenthart, J.F.G. (1983) J. Biol. Chem. 258, 4966-4971). In the present study, the amino acid sequence of the long chain of this protein, designated A-chain, was determined and found to consist of 282 amino acid residues. Twenty-four amino acid doublets were found; the most abundant of these are Pro-Pro and Ala-Ala which each occur five times. Of particular interest is the presence of three Gly-X-Pro and one Gly-Pro-X sequences that are characteristic of the repeating sequences of collagens. Chou-Fasman evaluation of the secondary structure suggested that the A-chain contains 29% alpha-helix, 24% beta-pleated sheet, and 26% reverse turns and, thus, approximately 80% of the polypeptide chain may display ordered structure. Four glycosylation sites were identified. The two N-glycosidic oligosaccharides were found in the center region (residues 138 and 158), whereas the two O-glycosidic heterosaccharides, both linked to threonine (residues 238 and 252), occur within the carboxyl-terminal region. The N-glycans are linked to Asn residues in beta-turns, while the O-glycans are located in short random segments. Comparison of the sequence of the amino- and carboxyl-terminal 30 residues with protein sequences in a data bank demonstrated that the A-chain is not significantly related to any known proteins. However, the proline-rich carboxyl-terminal region of the A-chain displays some sequence similarity to collagens and the collagen-like domains of complement subcomponent C1q.  相似文献   

16.
The formation of closed icosahedral capsids from a single species of coat protein subunit requires that the subunits assume different conformations at different lattice positions. In the double-stranded DNA bacteriophage P22, formation of correctly dimensioned capsids is mediated by interaction between coat protein subunits and scaffolding protein. Raman spectroscopy has been employed to compare the conformations of coat protein subunits which have been polymerized to form capsids in the presence and absence of the of scaffolding protein display a Raman spectrum characterized by a broad amide I band centered at 1665 cm-1 with a discernible shoulder near 1653 cm-1, and a broad amide III profile centered at 1238 cm-1 but asymmetrically skewed to higher frequency. These spectral features indicate that the protein conformation in procapsid shells is rich in beta-sheet secondary structure but contains also a significant distribution of alpha-helix. When biologically active, purified subunits assemble in the absence of scaffolding protein, they form polydisperse multimers lacking the proper dimensions of procapsid closed shells. We designate these multimers as "associated subunits" (AS). The Raman spectrum of associated subunits indicates a narrower distribution of secondary structure. The associated subunits are characterized by a sharper and more intense Raman amide I band at 1666 cm-1, with no prominent amide I shoulder of lower frequency. An analogous narrowing of the Raman amide III profile is also observed for AS particles, with an accompanying shift of the amide III band center to 1235 cm-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The interaction of melittin with monolayers of 1,2-dipalmitoylphosphatidylcholine and 1,2-dipalmitoylphosphatidylserine has been investigated with infrared external reflection-absorption spectroscopy. Improved instrumentation permits determination of acyl chain conformation and peptide secondary structure in situ at the air/water interface. The IR frequency of the 1,2-dipalmitoylphosphatidylcholine antisymmetric acyl chain CH2 stretching vibration decreases by 1.3 cm-1 upon melittin insertion, consistent with acyl chain ordering, whereas the same vibrational mode increases by 0.5 cm-1 upon peptide interaction with the 1,2-dipalmitoylphosphatidylserine monolayer, indicative of chain disordering. Thus the peptide interacts quite differently with zwitterionic compared with negatively charged monolayer surfaces. Melittin in the monolayer adopted a secondary structure with an amide l(l') frequency (1635 cm-1) dramatically different from the alpha-helical motif (amide l frequency 1656 cm-1 in a dry or H2O hydrated environment, amide l' frequency 1645 cm-1 in an H-->D exchanged alpha-helix) assumed in bilayer or multibilayer environments. This work represents the first direct in situ spectroscopic indication that peptide secondary structure in lipid monolayers may differ from that in bilayers.  相似文献   

18.
We have recorded the circular dichroism spectra of the cellular and the viral H-ras gene products both in the absence and in the presence of guanine nucleotides and analyzed these spectra in terms of the secondary structure composition of these proteins. It is shown that the GTP complex of the ras proteins has a different secondary structure composition than the GDP complex and, furthermore, that there are differences in the secondary structure of the viral ras protein and the cellular ras protein. We have also recorded and analyzed the circular dichroism spectrum of the isolated guanine nucleotide binding domain of the Escherichia coli elongation factor Tu (EF-Tu), which has been considered as a model for the tertiary structure of the ras proteins [McCormick, F., Clark, B. F. C., LaCour, T. F. M., Kjeldgaard, M., Norskov-Lauritsen, L., & Nyborg, J. (1985) Science (Washington, D.C.) 230, 78-82]. Our data show that the guanine nucleotide binding domain of EF-Tu (30% alpha-helix and 16% beta-pleated sheet for the GDP complex) has quite a different secondary structure composition than the ras proteins (e.g., the cellular ras protein has 47% alpha-helix and 22% beta-pleated sheet for the GDP complex), indicating that the protein core comprising the guanine nucleotide binding site might be similar but that major structural differences must exist at the portion outside this core. Normal and transforming ras proteins also differ slightly in their hydrodynamic properties as shown by sedimentation velocity runs in the analytical ultracentrifuge.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Peptide-chain secondary structure of bacteriorhodopsin.   总被引:7,自引:3,他引:4       下载免费PDF全文
Ultraviolet circular dichroism spectroscopy in the interval from 190 to 240 nm and infrared spectroscopy in the region of the amide I band (1,600 cm-1 to 1,700 cm-1) has been used to estimate the alpha-helix content and the beta-sheet content of bacteriorhodopsin. Circular dichroism spectroscopy strongly suggests that the alpha-helix content is sufficient for only five helices, if each helix is composed of 20 or more residues. It also suggests that there is substantial beta-sheet conformation in bacteriorhodopsin. The presence of beta-sheet secondary structure is further suggested by the presence of a 1,639 cm-1 shoulder on the amide I band in the infrared spectrum. Although a structural model consisting of seven alpha-helical rods has been generally accepted up to this point, the spectroscopic data are more consistent with a model consisting of five alpha-helices and four strands of beta-sheet. We note that the primary amino acid sequence can be assigned to segments of alpha-helix and beta-sheet in a way that does not require burying more than two charged groups in the hydrophobic membrane interior, contrary to the situation for any seven-helix model.  相似文献   

20.
Infrared spectroscopy in the interval from 1800 to 1300 cm-1 has been used to investigate the secondary structure and the hydrogen/deuterium exchange behavior of bacteriorhodopsin and bovine rhodopsin in their respective native membranes. The amide I' and amide II' regions from spectra of membrane suspensions in D2O were decomposed into constituent bands by use of a curve-fitting procedure. The amide I' bands could be fit with a minimum of three theoretical components having peak positions at 1664, 1638, and 1625 cm-1 for bacteriorhodopsin and 1657, 1639, and 1625 cm-1 for rhodopsin. For both of these membrane proteins, the amide I' spectrum suggests that alpha-helix is the predominant form of peptide chain secondary structure, but that a substantial amount of beta-sheet conformation is present as well. The shape of the amide I' band was pH-sensitive for photoreceptor membranes, but not for purple membrane, indicating that membrane-bound rhodopsin undergoes a conformation change at acidic pH. Peptide hydrogen exchange of bacteriorhodopsin and rhodopsin was monitored by observing the change in the ratio of integrated absorbance (Aamide II'/Aamide I') during the interval from 1.5 to 25 h after membranes were introduced into buffered D2O. The fraction of peptide groups in a very slowly exchanging secondary structure was estimated to be 0.71 for bacteriorhodopsin at pD 7. The corresponding fraction in vertebrate rhodopsin was estimated to be less than or equal to 0.60. These findings are discussed in relationship to previous studies of hydrogen exchange behavior and to structural models for both proteins.  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