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 858 毫秒
1.
Azobenzene derivatives can be used to reversibly photoregulate secondary structure when introduced as intramolecular bridges in peptides and proteins. Here we report the design, synthesis, and characterization of a disubstituted N,N-dialkyl azobenzene derivative that absorbs near 480 nm in aqueous solution and relaxes with a half-life of approximately 50 ms at room temperature. The wavelength of maximum absorbance and the rate of thermal relaxation are solvent-dependent. An increase in the percentage of organic solvent leads, in general, to a blue shift in the absorbance maximum and a slowing of the relaxation rate. In accordance with the design, the thermal relaxation of the azobenzene cross-linker from cis to trans causes an increase in the helix content of one peptide where the linker is attached via cysteine residues spaced at i, i + 11 positions and a decrease in helix content of another peptide with cysteine residues spaced at i, i + 7. This cross-linker design thus expands the possibilities for fast photocontrol of peptide and protein structure.  相似文献   

2.
Inai Y  Komori H 《Biomacromolecules》2004,5(4):1231-1240
The noncovalent chiral domino effect (NCDE), defined as chiral interaction upon an N-terminus of a 3(10)-helical peptide, will provide a unique method for structural control of a peptide helix through the use of external chirality. On the other hand, the NCDE has not been considered to be effective for the helicity control of peptides strongly favoring a one-handed screw sense. We here aim to promote the NCDE on peptide helicity using two types of nonapeptides: H-beta-Ala-Delta(Z)Phe-Aib-Delta(Z)Phe-X-(Delta(Z)Phe-Aib)(2)-OCH(3) [Delta(Z)Phe = alpha,beta-didehydrophenylalanine, Aib = alpha-aminoisobutyric acid], where X as the single chirality is L-leucine (1) or L-phenylalanine (2). NMR, IR, and CD spectroscopy as well as energy calculation revealed that both peptides alone form a right-handed 3(10)-helix. The original CD amplitudes or signs in chloroform, irrespective of a strong screw-sense preference in the central chirality, responded sensitively to external chiral information. Namely added Boc-L-amino acid stabilized the original right-handed helix, while the corresponding d-isomer destabilized it or transformed it into a left-handed helix. These peptides were also shown to bind more favorably to an L-isomer from the racemate. Although similar helicity control was observed for analogous nonapeptides bearing an N-terminal Aib residue (Inai, Y.; et al. Biomacromolecules 2003, 4, 122), the present findings demonstrate that the N-terminal replacement by the beta-Ala residue significantly improves the previous NCDE to achieve more effective control of helicity. Semiempirical molecular orbital calculations on complexation of peptide 2 with Boc-(L or D)-Pro-OH reasonably explained the unique conformational change induced by external chirality.  相似文献   

3.
Ampullosporin A (AmpA), a 15mer peptalbol containing seven Aib residues is able to induce pigmentation on Phoma destructiva and hypothermia in mice, as well as to exhibit a neuroleptic effect. A circular dichroism study of ampullosporin A and its analogues was carried out in organic solvents with different polarities and detergent micelles to determine the relationship between their conformational flexibility and biological activities. The analogues were obtained by modifying the N- and C-termini of ampullosporin A. Furthermore, Gln and Leu were systematically substituted by Ala and Aib residues were replaced by Ala and/or Ac6c. To estimate the helicity of the analogues, the CD spectrum of AmpA recorded in acetonitrile was correlated to its crystal structure. All analogues displayed similar CD curve shapes in organic solvents with the ratio between two negative band intensities R = [theta]n-pi*/[theta]pi-pi* < 1. In acetonitrile, most of the analogues adopted a 70%-85% helical structure, which was higher than the average of 40%-60% obtained in TFE. In detergent micelles, the analogues were distinguishable by their CD profiles. For most of the biologically active analogues, the CD spectra in detergent micelles were characterized by a R ratio > 1 and increased helicity compared with those recorded in TFE, suggesting that the interaction of the peptides with the membrane and peptide association was necessary for their hypothermic effect.  相似文献   

4.
T Taga  M Itoh  K Machida  T Fujita  T Ichihara 《Biopolymers》1990,29(6-7):1057-1064
The x-ray diffraction study of the C-terminally unprotected tripeptide benzyloxy-carbonyl-L-alanyl-alpha-aminoisobutyl-alpha- aminoisobutylic acid (Z-L-Ala-Aib-Aib-OH) has shown that the molecule adopts a consecutive type III beta-turn, which characterizes a right-handed 3(10) helix. A very weak 4----1 intramolecular hydrogen bond with the long N...O distance of 3.32 A, and a unique "oxy analogue" of the 4----1 hydrogen bond wih the O...O distance of 2.77 A, were observed. The stability of the observed conformation with the asymmetric Aib residues was theoretically evaluated by a conformation-energy calculation. The stereochemical characteristics of Aib and Ala residues were made clear by a comparison of the conformations of the short peptides containing both Aib and Ala residues.  相似文献   

5.
The substitution Ala----Gly has been studied in a unique-sequence peptide (related in sequence to the C-peptide of ribonuclease A) to determine its effect on C-peptide helicity at different residue positions. There is a substantial decrease in helicity for Ala----Gly at residue position 4, 5, or 6 but only a small decrease in helicity for Ala----Gly at end residue 1 and no decrease at end residue 13. The change for Ala----Gly is similar at position 4, 5, or 6; the change is caused chiefly by the difference in s, the helix growth parameter in the Zimm-Bragg model for alpha-helix formation, between Ala and Gly. Thus, the helicity of C-peptide depends sensitively on s at interior positions. The small change in helicity found for Ala----Gly at either end position suggests that the end residues are largely excluded from the helix, with the result that helicity is relatively unaffected by replacement of an end residue. Another possibility is that some helix-stabilizing effect is exerted by Gly only at an end position. Exclusion of an end residue from the helix might be caused either by fraying of the helix ends or by helix termination at an interior residue, resulting from a helix stop signal such as the Glu-2- -Arg-10+ salt bridge or the Phe-8-His-12+ ring interaction.  相似文献   

6.
A number of cell-penetrating peptides (CPPs) have been characterized and their usefulness as delivery tools has been clarified. As one of the CPPs, model amphipathic peptide (MAP) was developed by integrating both hydrophobic and hydrophilic amino acids in its sequence. In our previous work, we designed MAP(Aib) by replacing five alanine (Ala) residues on the hydrophobic face of the helix in the MAP sequence with α-aminoisobutyric acid (Aib) residues, and the replacement resulted in higher helix propensity, stronger resistance to protease, and higher cell membrane permeability than MAP. As a next step, we examined the efficiency of oligonucleotide (ODN) delivery into cells by MAP(Aib) in comparison with that by MAP. The electrostatically formed MAP(Aib)/ODN complex was more easily taken up by cells than the MAP/ODN complex, and the ODN delivery by MAP(Aib) was via an endocytic pathway. We demonstrated that the incorporation of Aib residues into CPPs enhances the delivery of hydrophilic molecules, such as ODN, into cells.  相似文献   

7.
We report a rapid method to synthesize cystine cross-linked heterotrimeric collagenous peptides. They can be engineered to favour one particular axial alignment of the strands, called the register of the helix. Here, the sequence of the constituent peptides contains 18 residues of "guest" collagen type I sequence flanked by N and C-terminal (Gly-Pro-Pro)5 "host" modules which ensure helicity. Further C-terminal residues include appropriately spaced cysteine residues and alanine to provide the necessary flexibility for helix formation. The cross-linking reaction and subsequent separation protocols have been designed for any inserted collagen sequence that does not contain a cysteine residue. Mass spectrometry and ion-exchange chromatography allow us to distinguish between different disulphide-bonded species and to monitor the formation of side-products. Starting peptide can be recovered simply from the reaction mixture by reduction and separation. Yields are typically 30%, working on a 10 mg scale. 15N-1H NMR and platelet adhesion studies show that the peptide heterotrimers presented here can reshuffle to cover all three axial registers. Less flexible spacers between the disulphide linkages and the helix will restrict each heterotrimer to one register only.  相似文献   

8.
The amino acid Aib predisposes a peptide to be helical with context‐dependent preference for either 310‐ or α‐ or a mixed helical conformation. Short peptides also show an inherent tendency to be unfolded. To characterize helical and unfolded states adopted by water‐soluble Aib‐containing peptides, the conformational preference of Ac‐Ala‐Aib‐Ala‐Lys‐Ala‐Aib‐Lys‐Ala‐Lys‐Ala‐Aib‐Tyr‐NH2 was determined by CD, NMR and MD simulations as a function of temperature. Temperature‐dependent CD data indicated the contribution of two major components, each an admixture of helical and extended/polyproline II structures. Both right‐ and left‐handed helical conformations were detected from deconvolution of CD data and 13C NMR experiments. The presence of a helical backbone, more pronounced at the N‐terminal, and a temperature‐induced shift in α‐helix/310‐helix equilibrium, more pronounced at the C‐terminal, emerged from NMR data. Starting from polyproline II, the N‐terminal of the peptide folded into a helical backbone in MD simulations within 5 ns at 60°C. Longer simulations showed a mixed‐helical backbone to be stable over the entire peptide at 5°C while at 60°C the mixed‐helix was either stable at the N‐terminus or occurred in short stretches through out the peptide, along with a significant population of polyproline II. Our results point towards conformational heterogeneity of water‐soluble Aib‐based peptide helices and the associated subtleties. The problem of analyzing CD and NMR data of both left‐ and right‐handed helices are discussed, especially the validity of the ellipticity ratio [θ]222/[θ]207, as a reporter of α‐/310‐ population ratio, in right‐ and left‐handed helical mixtures. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

9.
A P Todd  G L Millhauser 《Biochemistry》1991,30(22):5515-5523
A series of short alanine-based synthetic peptides (16 or 17 residues) have previously been shown to exhibit an anomalously high degree of alpha-helicity [Marqusee, S., et al. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 5286-5290; Marqusee, S., & Baldwin, R.L. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 8898-8902]. These peptides are ideal models for extracting position-dependent structural and dynamic information. Using the methanethiosulfonate nitroxide spin label (MTSSL), we labeled an analogue of the salt-bridge-stabilized "i+4" peptide, called the "i+4c", which has a specific attachment site created by replacing the central alanine with a cysteine. Circular dichroism (CD) spectra demonstrate that the i+4c-MTSSL peptide retains nearly the same helicity as the original i+4 peptide. The ESR spectra of the labeled peptide indicate no significant aggregation. ESR spectra were acquired throughout the helix-coil transition by temperature variation. From the motionally narrowed spectra, we extracted the rotational correlation times of the nitroxide label. Parallel measurements with circular dichroism enabled us to relate these parameters directly to the fractional helicity. For comparison, we followed a similar procedure with MTSSL-labeled glutathione (GS-MTSSL), a tripeptide that does not form an alpha-helix. Our results are interpreted in terms of a local tumbling volume, V(L), which reflects the portion of the peptide that reorients with the nitroxide label. At high fractional helicity, V(L) is similar to the volume expected for a 17-residue helix.  相似文献   

10.
The folding of short alanine-based peptides with different numbers of lysine residues is simulated at constant temperature (274 K) using the rigid-element Monte Carlo method. The solvent-referenced potential has prevented the multiple-minima problem in helix folding. From various initial structures, the peptides with three lysine residues fold into helix-dominated conformations with the calculated average helicity in the range of 60-80%. The peptide with six lysine residues shows only 8-14% helicity. These results agree well with experimental observations. The intramolecular electrostatic interaction of the charged lysine side chains and their electrostatic hydration destabilize the helical conformations of the peptide with six lysine residues, whereas these effects on the peptides with three lysine residues are small. The simulations provide insight into the helix-folding mechanism, including the beta-bend intermediate in helix initiation, the (i, i + 3) hydrogen bonds, the asymmetrical helix propagation, and the asymmetrical helicities in the N- and C-terminal regions. These findings are consistent with previous studies.  相似文献   

11.
We have investigated the solution conformation of the functionally relevant C-terminal extremes of alpha- and beta-tubulin, employing the model recombinant peptides RL52alpha3 and RL33beta6, which correspond to the amino acid sequences 404-451(end) and 394-445(end) of the main vertebrate isotypes of alpha- and beta-tubulin, respectively, and synthetic peptides with the alpha-tubulin(430-443) and beta-tubulin(412-431) internal sequences. Alpha(404-451) and beta(394-445) are monomeric in neutral aqueous solution (as indicated by sedimentation equilibrium), and have circular dichroism (CD) spectra characteristic of nearly disordered conformation, consistent with low scores in peptide helicity prediction. Limited proteolysis of beta(394-445) with subtilisin, instead of giving extensive degradation, resulted in main cleavages at positions Thr409-Glu410 and Tyr422-Gln423-Gln424, defining the proteolysis resistant segment 410-422, which corresponds to the central part of the predicted beta-tubulin C-terminal helix. Both recombinant peptides inhibited microtubule assembly, probably due to sequestration of the microtubule stabilizing associated proteins. Trifluoroethanol (TFE)-induced markedly helical CD spectra in alpha(404-451) and beta(394-445). A substantial part of the helicity of beta(394-445) was found to be in the CD spectrum of the shorter peptide beta(412-431) with TFE. Two-dimensional 1H-NMR parameters (nonsequential nuclear Overhauser effects (NOE) and conformational C alphaH shifts) in 30% TFE permitted to conclude that about 25% of alpha(404-451) and 40% of beta(394-451) form well-defined helices encompassing residues 418-432 and 408-431, respectively, flanked by disordered N- and C-segments. The side chains of beta(394-451) residues Leu418, Val419, Ser420, Tyr422, Tyr425, and Gln426 are well defined in structure calculations from the NOE distance constraints. The apolar faces of the helix in both alpha and beta chains share a characteristic sequence of conserved residues Ala,Met(+4),Leu(+7),Tyr(+11). The helical segment of alpha(404-451) is the same as that described in the electron crystallographic model structure of alphabeta-tubulin, while in beta(394-451) it extends for nine residues more, supporting the possibility of a functional coil --> helix transition at the C-terminus of beta-tubulin. These peptides may be employed to construct model complexes with microtubule associated protein binding sites.  相似文献   

12.
Understanding the secondary structure of peptides is important in protein folding, enzyme function, and peptide‐based drug design. Previous studies of synthetic Ala‐based peptides (>12 a.a.) have demonstrated the role for charged side chain interactions involving Glu/Lys or Glu/Arg spaced three (i, i + 3) or four (i, i + 4) residues apart. The secondary structure of short peptides (<9 a.a.), however, has not been investigated. In this study, the effect of repetitive Glu/Lys or Glu/Arg side chain interactions, giving rise to E‐R/K helices, on the helicity of short peptides was examined using circular dichroism. Short E‐R/K–based peptides show significant helix content. Peptides containing one or more E‐R interactions display greater helicity than those with similar E‐K interactions. Significant helicity is achieved in Arg‐based E‐R/K peptides eight, six, and five amino acids long. In these short peptides, each additional i + 3 and i + 4 salt bridge has substantial contribution to fractional helix content. The E‐R/K peptides exhibit a strongly linear melt curve indicative of noncooperative folding. The significant helicity of these short peptides with predictable dependence on number, position, and type of side chain interactions makes them an important consideration in peptide design.  相似文献   

13.
The introduction into peptide chains of alpha-aminoisobutyric acid (Aib) has proven to stabilize the helical structure in short peptides by restricting the available range of polypeptide backbone conformations. In order to evaluate the potential stabilizing effect of Aib at the protein level, we have studied the conformational and stability properties of Aib-containing analogs of the carboxy-terminal subdomain 255-316 of thermolysin. Previous NMR studies have shown that this disulfide-free 62-residue fragment forms a dimer in solution and that the global 3D structure of each monomer (3 alpha-helices encompassing residues 260-274, 281-295, and 301-311) is largely coincident with that of the corresponding region in the X-ray structure of intact thermolysin. The Aib analogs of fragment 255-316 were prepared by a semisynthetic approach in which the natural fragment 255-316 was coupled to synthetic analogs of peptide 303-316 using V8-protease in 50% (v/v) aqueous glycerol [De Filippis, V., and Fontana, A. (1990) Int. J. Pept. Protein Res. 35, 219-227]. The Ala residue in position 304, 309, or 312 of fragment 255-316 was replaced by Aib, leading to the singly substituted fragments Ala304Aib, Ala309Aib, and Ala312Aib. Moreover, fragment Ala304Aib/Ala309Aib with a double Ala --> Aib exchange in positions 304 and 309 was produced. Far- and near-UV circular dichroism measurements demonstrated that both secondary and tertiary structures of the natural fragment 255-316 are fully retained upon Ala --> Aib substitution(s). Thermal unfolding measurements, carried out by recording the ellipticity at 222 nm upon heating, showed that the melting temperatures (Tm) of analogs Ala304Aib and Ala309Aib were 2.2 and 5.4 °C higher than that of the Ala-containing natural species (Tm = 63.5 °C), respectively, whereas the Tm of the Ala312Aib analog was lowered by -0.6 °C. The enhanced stability of the Ala304Aib analog can be quantitatively explained on the basis of a reduced backbone entropy of unfolding due to the restriction of the conformational space allowed to Aib in respect to Ala, while the larger stabilization observed for the Ala309Aib analog can be accounted for by both entropic and hydrophobic effects. In fact, whereas Ala304 is a surface residue, Ala309 is shielded from the solvent, and thus the enhanced stability of fragment Ala309Aib is also due to the burial of an additional -CH3 group with respect to the natural fragment. The slightly destabilizing effect of the Ala --> Aib exchange in position 312 appears to derive from unfavorable strain energy effects, since phi and psi values for Ala312 are out of the allowed angles for Aib. Of interest, the simultaneous incorporation of Aib at positions 304 and 309 leads to a significant and additive increase of +8 °C in Tm. The results of this study indicate that the rational incorporation of Aib into a polypeptide chain can be a general procedure to significantly stabilize proteins.  相似文献   

14.
The conformational behaviour of delta Ala has been investigated by quantum mechanical method PCILO in the model dipeptide Ac-delta Ala-NHMe and in the model tripeptides Ac-X-delta Ala-NHMe with X = Gly, Ala, Val, Leu, Abu and Phe and is found to be quite different. The computational results suggest that in the model tripeptides the most stable conformation corresponds to phi 1 = -30 degrees, psi 1 = 120 degrees and phi 2 = psi 2 = 30 degrees in which the > C = 0 of the acetyl group is involved in hydrogen bond formation with N-H of the amide group. Similar results were obtained for the conformational behaviour of D-Ala in Ac-D-Ala-NHMe and Ac-Ala-D-Ala-NHMe. The conformational behaviour of the amino acids delta Ala, D-Ala, Val and Aib in model tripeptides have been utilized in the designing of left handed helical peptides. It is shown that the peptide HCO-(Ala-D-Ala)3-NHMe can adopt both left and right handed helix whereas in the peptide Ac-(Ala-delta Ala)3-NHMe the lowest energy conformer is beta-bend ribbon structure. Left handed helical structure with phi = 30 degrees, psi = 60 degrees for D-Ala residues and phi = psi = 30 degrees for delta Ala is found to be more stable by 4 kcal mole-1 than the corresponding right handed helical structure for the peptide Ac-(D-Ala-delta Ala)3-NHMe. In both the peptides Ac-(Val-delta Ala)3-NHMe and Ac-(D-Val-delta Ala)3-NHMe the most stable conformer is the left handed helix. Comparisons of results for Ac-(Ala-delta Ala)3-NHMe and Ac(Val-delta Ala)3-NHMe and Ac-(D-Ala-delta Ala)3-NHMe and Ac-(D-Val-delta Ala)3-NHMe also reveal that the Val residues facilitate the population of 3(10) left handed helix over the other conformers. It is also shown that the conformational behaviour of Aib residue depends on the chirality of neighbouring amino acids, i.e. Ac-(Aib-Ala)3-NHMe adopts right handed helical structure whereas Ac-(Aib-D-Ala)3-NHMe is found to be in left handed helical structure.  相似文献   

15.
The CD spectra of the peptides Boc-X-(Aib-X)n-OMe (n = 1, 2, 3) and Boc-(Aib-X)5-OMe, where X = L -Ala or L -Val have been examined in several solvents. The X = Ala and Val peptides behave similarly in all solvents, suggesting that the Aib residues dominate the folding preferences of these peptides. The decapeptides adopt helical conformations in methanol and trifluoroethanol, with characteristic negative CD bands at 222 and 205 nm. In the heptapeptides, similar spectra with reduced intensities are observed. Comparison with nmr studies suggest that estimates of helical content in oligopeptides by CD methods may lead to erroneous conclusions. The pentapeptides yield solvent-dependent spectra indicative of conformational perturbations. Peptide association in dioxane results in an unusual spectrum with a single negative band at 210 nm for the decapeptides. Disaggregation is induced by the addition of methanol or water to dioxane solutions. Aggregation of the heptapeptides is less pronounced in dioxane, suggesting that a critical helix length may be necessary to promote association stabilized by helix dipole–dipole interactions.  相似文献   

16.
The secondary structure and membrane-associated conformation of a synthetic peptide corresponding to the putative membrane-binding C-terminal 38 residues of the bovine milk component PP3 was determined using 1H NMR in methanol, CD in methanol and SDS micelles, and 15N solid-state NMR in planar phospholipid bilayers. The solution NMR and CD spectra reveal that the PP3 peptide in methanol and SDS predominantly adopts an alpha-helical conformation extending over its entire length with a potential bend around residue 19. 15N solid-state NMR of two PP3 peptides 15N-labelled at the Gly7 and Ala32 positions, respectively, and dissolved in dimyristoylphosphatidylcholine/dimyristoylphosphatidylglycerol phospholipid bilayers shows that the peptide is associated to the membrane surface with the amphipathic helix axis oriented parallel to the bilayer surface.  相似文献   

17.
Asn deamidation was monitored in Ala-based octadecapeptides of varying alpha-helicity. Gly was substituted for Ala residues at positions 6 and 16 to create a peptide with less helicity. Ala --> Gly substitutions were made at three or more residues from the Asn to negate known primary sequence effects on deamidation rates. The extent of helicity and rate of Asn deamidation for alkaline aqueous solutions of each peptide was measured as a function of temperature by circular dichroism and reversed-phase high-performance liquid chromatography, respectively. The rate of deamidation in the peptides was inversely proportional to the extent of alpha-helicity. The results support the conclusion that Asn deamidation only occurs in the nonhelical population of conformers.  相似文献   

18.
19.
Cluster determinant 4 (CD4) is a type I transmembrane glycoprotein of 58 kDa. It consists of an extracellular domain of 370 amino acids, a short transmembrane region, and a cytoplasmic domain of 40 amino acids at the C-terminal end. We investigated the structure of the 62 C-terminal residues of CD4, comprising its transmembrane and cytoplasmic domains. The five cysteine residues of this region have been replaced with serine and histidine residues in the polypeptide CD4mut. Uniformly 15N and 13C labeled protein was recombinantly expressed in E. coli and purified. Functional binding activity of CD4mut to protein VpU of the human immunodeficiency virus type 1 (HIV-1) was verified. Close to complete NMR resonance assignment of the 1H, 13C, and 15N spins of CD4mut was accomplished. The secondary structure of CD4mut in membrane simulating dodecylphosphocholine (DPC) micelles was characterized based on secondary chemical shift analysis, NOE-based proton-proton distances, and circular dichroism spectroscopy. A stable transmembrane helix and a short amphipathic helix in the cytoplasmic region were identified. The fractional helicity of the cytoplasmic helix appears to be stabilized in the presence of DPC micelles, although the extension of this helix is reduced in comparison to previous studies on synthetic peptides in aqueous solution. The role of the amphipathic helix and its potentially variable length is discussed with respect to the biological functions of CD4.  相似文献   

20.
The alpha-aminoisobutyric (Aib) residue has generally been considered to be a strongly helicogenic residue as evidenced by its ability to promote helical folding in synthetic and natural sequences. Crystal structures of several peptide natural products, peptaibols, have revealed predominantly helical conformations, despite the presence of multiple helix-breaking Pro or Hyp residues. Survey of synthetic Aib-containing peptides shows a preponderance of 3(10)-, alpha-, and mixed 3(10)/alpha-helical structures. This review highlights the examples of Aib residues observed in nonhelical conformations, which fall 'primarily' into the polyproline II (P(II)) and fully extended regions of conformational space. The achiral Aib residue can adopt both left (alpha(L))- and right (alpha(R))-handed helical conformations. In sequences containing chiral amino acids, helix termination can occur by means of chiral reversal at an Aib residue, resulting in formation of a Schellman motif. Examples of Aib residues in unusual conformations are illustrated by surveying a database of Aib-containing crystal structures.  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