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 257 毫秒
1.
In order to examine the potential correlation between infrared absorption spectra and 3(10)- and alpha-helices and beta-bend ribbon structures, the secondary structures of synthetic peptides known to contain pure 3(10)-helices, mixed 3(10)/alpha-helices, and pure beta-bend ribbon structures, based upon X-ray diffraction and NMR studies, have been investigated by using FTIR spectroscopy incorporating resolution-enhancement techniques. Studies of the peptides known to contain a stable 3(10)-helix in CDCl3 show the main amide I band of fully stable 3(10)-helices occurs at 1666-1662 cm-1. Resolution-enhancement methods revealed small contributions at 1681-1678 and 1646-1644 cm-1, while the amide II band occurs at 1533-1531 cm-1. Peptides known to contain both alpha- and 3(10)-helices in their structure exhibit bands characteristic of both types of conformation. Peptides known to fold into the beta-bend ribbon structure show an amide I band maximum at 1648-1645 cm-1 with the amide II band at 1538-1536 cm-1. Incorporation of these peptides into model membrane structures, e.g., DMPC vesicles, in aqueous buffer sometimes produces changes in the peptide secondary structure. Those peptides which possess a 3(10)-helical structure in CDCl3 solution change the secondary structure in DMPC vesicles to predominantly alpha-helical, plus a contribution from short, unstable 3(10)-helix and/or beta-turns. Those peptides which contain a combination of alpha- and 3(10)-helical structures in CDCl3 solution tend to retain some 3(10)-helical structure within the lipid environment, although the overall H-bonding pattern is altered. Those peptides which form a beta-bend ribbon structure appear to be largely unaffected in the membrane environment.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
J Cladera  M Sabés  E Padrós 《Biochemistry》1992,31(49):12363-12368
Fourier transform infrared spectroscopy at a resolution of 1 cm-1 has been used to study the conformation of dark-adapted bacteriorhodopsin in the native purple membrane, in H2O and D2O suspensions. A detailed analysis of the amide I bands was made using derivative and deconvolution techniques. Curve-fitting results of four independent experiments indicate, after estimation of the methodological errors, that native bacteriorhodopsin contains 52-73% alpha-helices, 13-19% reverse turns, 11-16% beta-sheets, and 3-7% unordered segments. Our analysis has enabled the identification of several components corresponding to alpha-helices, beta-sheets, and reverse turns. Besides the alpha I- and alpha II-helices (peaking at 1658 and 1665 cm-1), we propose that two more infrared bands arise from alpha-helical structures: one at 1650 cm-1 from alpha I and another one at 1642 cm-1 in H2O suspension, which could originate from type III beta-turns (i.e., one turn of 3(10)-helix). The relatively high content of reverse turns suggests the presence of one reverse turn per loop, plus another one in the C-terminal segment. On the other hand, several reasons argue that the calculated mean beta-sheet content of around 14% should be decreased somewhat. These beta-sheets could be located in the noncytoplasmatic links of the bacteriorhodopsin molecule.  相似文献   

3.
Fourier-transform infrared spectroscopy is a valuable method for the study of protein conformation in solution primarily because of the sensitivity to conformation of the amide I band (1700-1620 cm-1) which arises from the backbone C = O stretching vibration. Combined with resolution-enhancement techniques such as derivative spectroscopy and self-deconvolution, plus the application of iterative curve-fitting techniques, this method provides a wealth of information concerning protein secondary structure. Further extraction of conformational information from the amide I band is dependent upon discerning the correlations between specific conformational types and component bands in the amide I region. In this paper, we report spectra-structure correlations derived from conformational perturbations in bovine trypsin which arise from autolytic processing, zymogen activation, and active-site inhibition. IR spectra were collected for the single-chain (beta-trypsin) and once-cleaved, double-chain (alpha-trypsin) forms as well as at various times during the course of autolysis and also for zymogen, trypsinogen, and beta-trypsin inhibited with diisopropyl fluorophosphate. Spectral differences among the various molecular forms were interpreted in light of previous biochemical studies of autolysis and the known three-dimensional structures of the zymogen, the active enzyme, and the DIP-inhibited form. Our spectroscopic results from these proteins in D2O imply that certain loop structures may absorb in the region of 1655 cm-1. Previously, amide I' infrared bands near 1655 cm-1 have been interpreted as arising solely from alpha-helices. These new data suggest caution in interpreting this band. We have also proposed that regions of protein molecules which are known from crystallographic experiments to be disordered absorb in the 1645 cm-1 region and that type II beta-turns absorb in the region of 1672-1685 cm-1. Our results also corroborate assignment of the low-frequency component of extended strands to bands below 1636 cm-1. Additionally, the results of multiple measurements have allowed us to estimate the variability present in component band areas calculated by curve fitting the resolution-enhanced IR spectra. We estimate that this approach to data analysis and interpretation is sensitive to changes of 0.01 unit or less in the relative integrated intensities of component bands in spectra whose peaks are well resolved.  相似文献   

4.
A Blume  W Hübner  G Messner 《Biochemistry》1988,27(21):8239-8249
Fourier transform infrared spectroscopy has been used to characterize the carbonyl stretching vibration of DMPC, DMPE, DMPG, and DMPA, all labeled with 13C at the carbonyl group of the sn-2 chain. Due to the vibrational isotope effect, the 13C = O and the 12C = O vibrational bands are separated by ca. 40-43 cm-1. This frequency difference does not change when the labeling is reversed with the 13C = O group at the sn-1 chain. For lipids in organic solvents possible conformational differences between the sn-1 and sn-2 ester groups have no effect on the vibrational frequency of the C = O groups. In aqueous dispersion unlabeled phospholipids always show a superposition of two bands for the C = O vibration located at ca. 1740 and 1727 cm-1. These two bands have previously been assigned to the sn-1 and sn-2 C = O groups. FT-IR spectra of 13C-labeled phospholipids show that the vibrational bands of both, the sn-1 as well as the sn-2 C = O group, are clearly superpositions of at least two underlying components of different frequency and intensity. Band frequencies were determined by Fourier self-deconvolution and second-derivative spectroscopy. The difference between the component bands is ca. 11-17 cm-1. Again, the conformational effect as shown by reversed labeling is negligible with only 1-2 cm-1. The splitting of the C = O vibrational bands in H2O and D2O is caused by hydrogen bonding of water molecules to both C = O groups as shown by a comparison with spectra of model ester compounds in different solvents. To extract quantitative information about changes in hydration, band profiles were stimulated with Gaussian-Lorentzian functions. The chemical nature of the head group and its electronic charge have distinctive effects on the extent of hydration of the carbonyl groups. In the gel and liquid-crystalline phase of DMPC the sn-2 C = O group is more hydrated than the sn-1 C = O. This is accord with the conformation determined by X-ray analysis. In DMPG the sn-1 C = O group seems to be more accessible to water, indicating a different conformation of the glycerol backbone.  相似文献   

5.
Vibrational circular dichroism (VCD) studies are reported for two unrelated recombinant growth factor proteins: epidermal growth factor and basic fibroblast growth factor (bFGF). NMR, electronic CD, and bFGF X-ray studies indicate that these two proteins are primarily composed of beta-sheet and loop secondary structure elements with no detectable alpha-helices. Two reports on solution conformation of these proteins using FTIR absorption spectroscopy with subsequent resolution enhancement confirmed the presence of a large fraction of a beta-sheet conformation but in addition indicated the presence of large absorption bands in the 1650-1656 cm-1 region, which are typically assigned to alpha-helices. The VCD spectra of both proteins have band shapes that strongly resemble those of other high beta-sheet fraction proteins, such as the trypsin family of proteins. Quantitative analysis of the VCD spectra also indicates that these proteins are predominantly in beta-sheet and extended ("other") conformations with very little alpha-helix fraction. These results agree with the CD interpretation and affirm that the FTIR peaks in the region 1650-1656 cm-1 can be assigned to loops. This study provides an example of the limitations of using FTIR frequencies alone for examination of protein secondary structure.  相似文献   

6.
Raman spectra of the parallel-stranded duplex formed from the deoxyoligonucleotides 5'-d-[(A)10TAATTTTAAATATTT]-3' (D1) and 5'-d[(T)10ATTAAAATTTATAAA]-3' (D2) in H2O and D2O have been acquired. The spectra of the parallel-stranded DNA are then compared to the spectra of the antiparallel double helix formed from the deoxyoligonucleotides D1 and 5'-d(AAATATTTAAAATTA-(T)10]-3' (D3). The Raman spectra of the antiparallel-stranded (aps) duplex are reminiscent of the spectra of poly[d(A)].poly[d(T)] and a B-form structure similar to that adopted by the homopolymer duplex is assigned to the antiparallel double helix. The spectra of the parallel-stranded (ps) and antiparallel-stranded duplexes differ significantly due to changes in helical organization, i.e., base pairing, base stacking, and backbone conformation. Large changes observed in the carbonyl stretching region (1600-1700 cm-1) implicate the involvement of the C(2) carbonyl of thymine in base pairing. The interaction of adenine with the C(2) carbonyl of thymine is consistent wtih formation of reverse Watson-Crick base pairing in parallel-stranded DNA. Phosphate-furanose vibrations similar to those observed for B-form DNA of heterogenous sequence and high A,T content are observed at 843 and 1092 cm-1 in the spectra of the parallel-stranded duplex. The 843-cm-1 band is due to the presence of a sizable population of furanose rings in the C2'-endo conformation. Significant changes observed in the regions from 1150 to 1250 cm-1 and from 1340 to 1400 cm-1 in the spectra of the parallel-stranded duplex are attributed to variations in backbone torsional and glycosidic angles and base stacking.  相似文献   

7.
The conformation of the milk protein alpha-lactalbumin has been studied using vibrational circular dichroism (VCD) and compared to parallel studies on lysozyme. These proteins have been shown by Acharya et al. [(1989) J. Mol. Biol. 208, 99-127] to have very similar three-dimensional crystal structures. However, their VCD spectra in D2O solution are quite different. The VCD of lysozyme in D2O more resembles that of alpha-lactalbumin in 33% propanol/D2O, under which conditions alpha-lactalbumin has conformationally transformed to a structure with increased helical fraction. These results can be seen to be consistent with UVCD and resolution-enhanced FTIR spectra of alpha-lactalbumin and lysozyme in both D2O and H2O environments. The solvent sensitivity of the alpha-lactalbumin spectra and hence of its conformation contrasted with the lack of such sensitivity for lysozyme suggest that the alpha-lactalbumin crystal structure represents a conformation different from that which is dominant in aqueous solution.  相似文献   

8.
9.
1H-NMR spectra of bleomycin A2 recorded at 500 MHz in D2O and H2O at 24 degrees C and 3 degrees C were investigated. Resonances of the individual spin systems were identified by using two-dimensional correlation spectroscopy (COSY), two-dimensional spin echo correlated spectroscopy (SECSY) and by the application of two-dimensional Nuclear Overhauser Effect spectroscopy (NOESY). Employment of these techniques allowed the assignment of 113 exchangeable and 59 non-exchangeable protons in the 1H NMR spectrum of bleomycin A2. By means of 2D NOE spectroscopy also interresidual connectivities could be observed. Comparison of the NOESY spectra at 3 degrees C and 24 degrees C suggest that at low temperatures the central party of the bleomycin A2 molecule tends to adopt an extended conformation.  相似文献   

10.
Model peptides based on -(Aib-Ala)(n)-, and (Aib)(n)-Leu-(Aib)(2) sequences, which have varying amounts of 3(10)-helical character, were studied by use of vibrational and electronic circular dichroism (VCD and ECD) and Fourier transform infrared (FTIR) absorption spectroscopies to test the correlation of spectral response and conformation. The data indicate that these peptides, starting from a length of about four to six residues, predominantly adopt a 3(10)-helical conformation at room temperature. The longest model peptides, depending on the series, may evidence some alpha-helical contribution to the spectra, while the shorter ones, with less than six residues, have much less order. The IR absorption spectra (as supported by theory) showed only small frequency changes between 3(10)- and alpha-helices. By contrast, solvent effects are a source of much bigger perturbations. The ECD results show that the intensity ratio for the approximately 222-nm to approximately 208-nm bands, while useful for distinguishing between these two helical types in some sequences, may have a narrower range of application than VCD. However, the VCD data presented here continue to support the proposed discrimination between alpha- and 3(10)-helices based on qualitative amide I and II bandshape differences. The present study shows the intensities of the 3(10)-helical amide I (peak-to-peak) to its amide II VCD to be of the same order and useful for discriminating them from alpha-helices, whose amide I dominates the amide II in intensity. This qualitative result is experimentally independent of the amount of alphaMe-substituted residues in the sequence. These experimental VCD results are consistent in detail with theoretical spectral simulations for Ac-(Ala)(8)-NH(2), Ac-(Aib-Ala)(4)-NH(2), and Ac-(Aib)(8)-NH(2) in 3(10)- and alpha-helical conformations.  相似文献   

11.
Time-resolved infrared difference spectra of the ATP-induced phosphorylation of the sarcoplasmic reticulum Ca2+-ATPase have been recorded in H2O and 2H2O at pH 7.0 and 1 degrees C. The reaction was induced by ATP release from P3-1-(2-nitro)phenylethyladenosine 5'-triphosphate (caged ATP) and from [gamma-18O3]caged ATP. A band at 1546 cm-1, not observed with the deuterated enzyme, can be assigned to the amide II mode of the protein backbone and indicates that a conformational change associated with ATPase phosphorylation takes place after ATP binding. This is also indicated between 1700 and 1610 cm-1, where bandshifts of up to 10 cm-1 observed upon protein deuteration suggest that amide I modes of the protein backbone dominate the difference spectrum. From the band positions it is deduced that alpha-helical, beta-sheet, and probably beta-turn structures are affected in the phosphorylation reaction. Model spectra of acetyl phosphate, acetate, ATP, and ADP suggest the tentative assignment of some of the bands of the phosphorylation spectrum to the molecular groups of ATP and Asp351, which participate directly in the phosphate transfer reaction: a positive band at 1719 cm-1 to the C==O group of aspartyl phosphate, a negative band at 1239 cm-1 to the nuas(PO2-) modes of the bound ATP molecule, and a positive band at 1131 cm-1 to the nuas(PO32-) mode of the phosphoenzyme phosphate group, the latter assignment being supported by the band's sensitivity toward isotopic substitution in the gamma-phosphate of ATP. Band positions and shapes of these bands indicate that the alpha- and/or beta-phosphate(s) of the bound ATP molecule become partly dehydrated when ATP binds to the ATPase, that the phosphoenzyme phosphate group is unprotonated at pH 7.0, and that the C==O group of aspartyl phosphate does not interact with bulk water. The Ca2+ binding sites seem to be largely undisturbed by the phosphorylation reaction, and a functional role of the side chains of Asn, Gln, and Arg residues was not detected.  相似文献   

12.
Fourier transform infrared (FTIR) spectra have been obtained of human low-density lipoproteins (LDL) in H2O and 2H2O buffers. The absorption bands are assigned to vibrations of the lipid and apolipoprotein B-100 components. The analysis of second-derivative spectra allowed an assignment of individual protein bands to alpha-helical, random, coil or beta-structure and beta-turn conformations. Changes in the FTIR spectra after Cu2+-catalysed oxidation of the LDL particles indicate that the structure of apolipoprotein B-100 becomes less ordered, with some alterations of alpha-helical and beta-turn conformation. The main beta-structure absorption at 1620 cm-1 is unaffected by oxidation. Taking into account the resistance to oxidation and the slow H-2H exchange it is suggested that the beta-structure is hidden from external factors whereas other structures are mostly present on the surface of the LDL particle. Oxidation affects mainly the surface region of apolipoprotein B-100 and leads to a structural rearrangement which consequently changes the receptor specificity of the LDL.  相似文献   

13.
Local determinants of 3(10)-helix stabilization have been ascertained from the analysis of the crystal structure data base. We have clustered all 5-length substructures from 51 nonhomologous proteins into classes based on the conformational similarity of their backbone dihedral angles. Several clusters, derived from 3(10)-helices and multiple-turn conformations, had strong amino acid sequence patterns not evident among alpha-helices. Aspartate occurred over twice as frequently in the N-cap position of 3(10)-helices as in the N-cap position of alpha-helices. Unlike alpha-helices, 3(10)-helices had few C-termini ending in a left-handed alpha conformation; most 3(10) C-caps adopted an extended conformation. Differences in the distribution of hydrophobic residues among 3(10)- and alpha-helices were also apparent, producing amphipathic 3(10)-helices. Local interactions that stabilize 3(10)-helices can be inferred both from the strong amino acid preferences found for these short helices, as well as from the existence of substructures in which tertiary interactions replace consensus local interactions. Because the folding and unfolding of alpha-helices have been postulated to proceed through reverse-turn and 3(10)-helix intermediates, sequence differences between 3(10)- and alpha-helices can also lend insight into factors influencing alpha-helix initiation and propagation.  相似文献   

14.
The alternative structures of the synthetic poly(amino2dA-dT) duplex have been studied using infrared spectroscopy in films and in solution (D2O and H2O) in the presence and in the absence of magnesium salt. In solution without magnesium salt, the polynucleotide exists in a B genus conformation with some of the sugar puckers possibly in the C3'-endo/anti geometry. In magnesium-containing solution (66 mM MgCl2), however, we report infrared spectra of Mg(2+)-poly(amino2dA-dT) which have characteristic marker bands of the A form. Film samples in 70% relative humidity (RH) give similar infrared spectra to those of the polynucleotide obtained using Mg2+. Thus, when analyzed in comparison with previously reported infrared spectra of other oligo and polynucleotides, our data show that double helical poly(amino2dA-dT) goes into the same (or very closely related) conformation in dehydrated films as in solutions containing Mg2+.  相似文献   

15.
F Dousseau  M Pézolet 《Biochemistry》1990,29(37):8771-8779
A method for estimating protein secondary structure from infrared spectra has been developed. The infrared spectra of H2O solutions of 13 proteins of known crystal structure have been recorded and corrected for the spectral contribution of water in the amide I and II region by using the algorithm of Dousseau et al. [Dousseau, F., Therrien, M., & Pézolet, M. (1989) Appl. Spectrosc. 43, 538-542]. This calibration set of proteins has been analyzed by using either a classical least-squares (CLS) method or the partial least-squares (PLS) method. The pure-structure spectra calculated by the classical least-squares method are in good agreement with spectra of poly(L-lysine) in the alpha-helix, beta-sheet, and undefined conformations. The results show that the best agreement between the secondary structure determined by X-ray crystallography and that predicted by infrared spectroscopy is obtained when both the amide I and II bands are used to generate the calibration set, when the PLS method is used, and when it is assumed that the secondary structure of proteins is composed of only four types of structure: ordered and disordered alpha-helices, beta-sheet, and undefined conformation. Attempts to include turns in the secondary structure estimation have led to a loss of accuracy. The standard deviation of the difference between X-ray and infrared secondary structure estimates with this method is 4.8% for the alpha-helix, 3.7% for the beta-sheet, and 5.1% for the undefined structure, whereas the regression coefficients are 0.95, 0.96, and 0.56, respectively. The spectra of the calibration proteins were also recorded in 2H2O solution.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
We have examined the influence of monovalent and divalent cations on the secondary structure of bovine alpha-lactalbumin at neutral pH using Fourier-transform infrared spectroscopy. Our present studies are based on previously reported amide I' component band assignments for this protein [Prestrelski, S. J., Byler, D. M., & Thompson, M. P. (1991) Int. J. Pept. Protein Res. 37, 508-512]. The results indicate that upon dissolution, alpha-lactalbumin undergoes a small, but significant, time-dependent conformational change, regardless of the ions present. Additionally, these studies provide the first quantitative measure of the well-known secondary structural change which accompanies calcium binding. Results indicate that removal of Ca2+ from holo alpha-lactalbumin results in local unfolding of the Ca(2+)-binding loop; the spectra indicate that approximately 16% of the backbone chain changes from a rigid coordination complex to an unordered loop. We have also examined the effects of binding of several other metal ions. Our studies have revealed that binding of Mn2+ to apo alpha-lactalbumin (Ca(2+)-free), while inducing a small, but significant, conformational change, does not cause the alpha-lactalbumin backbone conformation to change to that of the holo (Ca(2+)-bound) form as characterized by infrared spectroscopy. Similar changes to those induced by Mn2+ are observed upon binding of Na+ to apo alpha-lactalbumin, and furthermore, even at very high concentrations (0.2 M), Na+ does not stabilize a structure similar to the holo form. Binding of Zn2+ to the apo form of alpha-lactalbumin does not result in significant backbone conformational changes, suggesting a rigid Zn(2+)-binding site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Infrared spectroscopy of a single cell--the human erythrocyte   总被引:1,自引:0,他引:1  
Methods for obtaining the infrared spectrum of a single erythrocyte by infrared microscopy have been developed. The spectrum contains the amide I, II, and III bands characteristic of protein secondary structure near 1650, 1550, and 1300 cm-1, respectively. Bound carbon monoxide exhibits a readily measured band at 1951 cm-1 for 12C16O and 1907 cm-1 for 13C16O. Both amide and CO bands are similar to those found for purified hemoglobin A. Spectra can be obtained in H2O or D2O media under physiologically relevant conditions. Single cell infrared spectroscopy (SCIR) permits the qualitative and quantitative determination of differences among individual red cells. These results suggest many potential applications for SCIR for the measurements of properties of individual cells at the molecular level under physiologically relevant conditions.  相似文献   

18.
IR spectra (1600-1800 and 3000-3650 cm-1) of lincomycin base solutions in inert (CCl4 and C2Cl4), proton acceptor (dioxane, dimethylsulfoxide and triethyl amine) and proton donor (CHCl3, CD3OD and D2O) solvents were studied. Analysis of the concentration and temperature changes in the spectra revealed that association in lincomycin in the inert solvents was due to intramolecular hydrogen linkage involving amide and hydroxyl groups. Disintegration of the associates after the solution dilution and temperature rise was accompanied by formation of intramolecular bonds stabilizing the stable conformation structure of the lincomycin molecule. The following hydrogen linkage in the conformation was realized: NH...N (band v NH...N at 3340 cm-1), OH...O involving the hydroxyl at C-7 and O atoms in the D-galactose ring (band v OH...O at 3548 cm-1), a chain of the hydrogen bonds OH...OH...OH in the lincomycin carbohydrate moiety (band v OH...O at 3593 cm-1 and v OH of the end hydroxyl group at 3625 cm-1). Bonds NH and C-O of the amide group were located in transconformation. Group C-O did not participate in the intramolecular hydrogen linkage.  相似文献   

19.
Infrared spectra for carbon monoxide bound to alpha and beta subunits of human hemoglobin A have subunit differences near 1950 cm-1 and indicate that 92% of the alpha subunits exist in one conformer and 5% in a second conformer under conditions where 99% of the beta subunit is in only one conformation. The sum of the separated subunit spectra is equivalent to the alpha 2 beta 2 tetramer spectrum. CO infrared spectra indicate that CO displaces O2 from HbO2 in red cells or in solution preferentially at the beta subunits. The measurement of C-O stretch bands provides a direct method for characterization of ligand binding sites within intact cells.  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