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.
Short cyclic peptides have a great interest in therapeutic, diagnostic and affinity chromatography applications. The screening of ‘one‐bead‐one‐peptide’ combinatorial libraries combined with mass spectrometry (MS) is an excellent tool to find peptides with affinity for any target protein. The fragmentation patterns of cyclic peptides are quite more complex than those of their linear counterparts, and the elucidation of the resulting tandem mass spectra is rather more difficult. Here, we propose a simple protocol for combinatorial cyclic libraries synthesis and ring opening before MS analysis. In this strategy, 4‐hydroxymethylbenzoic acid, which forms a benzyl ester with the first amino acid, was used as the linker. A glycolamidic ester group was incorporated after the combinatorial positions by adding glycolic acid. The library synthesis protocol consisted in the following: (i) incorporation of Fmoc‐Asp[2‐phenylisopropyl (OPp)]‐OH to Ala‐Gly‐oxymethylbenzamide‐ChemMatrix, (ii) synthesis of the combinatorial library, (iii) assembly of a glycolic acid, (iv) couple of an Ala residue in the N‐terminal, (v) removal of OPp, (vi) peptide cyclisation through side chain Asp and N‐Ala amino terminus and (vii) removal of side chain protecting groups. In order to simultaneously open the ring and release each peptide, benzyl and glycolamidic esters were cleaved with ammonia. Peptide sequences could be deduced from the tandem mass spectra of each single bead evaluated. The strategy herein proposed is suitable for the preparation of one‐bead‐one‐cyclic depsipeptide libraries that can be easily open for its sequencing by matrix‐assisted laser desorption/ionisation MS. It employs techniques and reagents frequently used in a broad range of laboratories without special expertise in organic synthesis. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The present study describes modification of asparagine–glycine–arginine (NGR) peptide at N‐terminally and C‐terminally by introduction of a tridentate chelating scaffold via click chemistry reaction. The N‐terminal and C‐terminal modified peptides were radiometalated with [99mTc(CO)3]+ precursor. The influence of these moieties at the two termini on the targeting properties of NGR peptide was determined by in vitro cell uptake studies and in vivo biodistribution studies. The two radiolabeled constructs did not exhibit any significant variation in uptake in murine melanoma B16F10 cells during in vitro studies. In vivo studies revealed nearly similar tumor uptake of N‐terminally modified peptide construct 5 and C‐terminally construct 6 at 2 h p.i. (1.9 ± 0.1 vs 2.4 ± 0.2% ID/g, respectively). The tumor‐to‐blood (T/B) and tumor‐to‐liver (T/L) ratios of the two radiometalated peptides were also quite similar. The two constructs cleared from all the major organs (heart, lungs, spleen, stomach, and blood) at 4 h p.i. (<1% ID/g). Blocking studies carried out by coinjection of cCNGRC peptide led to approximately 50% reduction in the tumor uptake at 2 h p.i. This work thus illustrates the possibility of convenient modification/radiometalation of NGR peptide at either N‐ or C‐terminus without hampering tumor targeting and pharmacokinetics.  相似文献   

3.
Cyclic Arg‐Gly‐Asp (RGD) peptides show remarkable affinity and specificity to integrin receptors and mediate important physiological effects in tumor angiogenesis. Additionally, they are one of the keyplayers in improving the biocompatibility of biomaterials. The fully biodegradable polymer poly(lactic‐co‐glycolic acid) (PLGA) is frequently used for biomedical implants and can be applied as nanoparticles for drug delivery. The aim of this work was the generation of a lipidated c[RGDfK] peptide including a second functionality for coating of hydrophobic PLGA. Therefore, we established a general and straightforward strategy for the introduction of two different modifications into the same c[RGDfK] peptide. This allowed the generation of a palmitoylated integrin‐binding lipopeptide that shows high affinity to PLGA. Additionally, we coupled 5(6)‐carboxyfluorescein to the second site for modification to enable sensitive quantification of the immobilized lipopeptide on PLGA. In conclusion, we present a synthesis protocol that enables the preparation of c[RGDfK] lipopeptides with a strong affinity to PLGA and an additional site for modifications. This will provide the opportunity to introduce a variety of effector molecules site‐specifically to the c[RGDfK] lipopeptide, which will enable the introduction of multifunctionality into c[RGDfK]‐coated PLGA devices or nanoparticles.  相似文献   

4.
Seminal amyloids are well known for their role in enhancing HIV infection. Among all the amyloidogenic peptides identified in human semen, PAP248‐286 was found to be the most active and was termed as semen‐derived enhancer of viral infection (SEVI). Although amyloidogenic nature of the peptide is mainly linked with enhancement of the viral infection, the most active physiological conformation of the aggregated peptide remains inconclusive. Lipids are known to modulate aggregation pathway of a variety of proteins and peptides and constitute one of the most abundant biomolecules in human semen. PAP248‐286 significantly differs from the other known amyloidogenic peptides, including Aβ and IAPP, in terms of critical concentration, surface charge, fibril morphology, and structural transition during aggregation. Hence, in the present study, we aimed to assess the effect of a lipid, 1,2‐dioleoyl‐sn‐glycero‐3‐phosphocholine (DOPC), on PAP248‐286 aggregation and the consequent conformational outcomes. Our initial observation suggested that the presence of the lipid considerably influenced the aggregation of PAP248‐286. Further, ZDOCK and MD simulation studies of peptide multimerization have suggested that the hydrophobic residues at C‐terminus are crucial for PAP248‐286 aggregation and are anticipated to be major DOPC‐interacting partners. Therefore, we further assessed the aggregation behaviour of C‐terminal (PAP273‐286) fragment of PAP248‐286 and observed that DOPC possesses the ability to interfere with the aggregation behaviour of both the peptides used in the current study. Mechanistically, we propose that the presence of DOPC causes considerable inhibition of the peptide aggregation by interfering with the peptide's disordered state to β‐sheet transition.  相似文献   

5.
To prevent aspartimide formation and related side products in Asp‐Xaa, particularly Asp‐Gly‐containing peptides, usually the 2‐hydroxy‐4‐methoxybenzyl (Hmb) backbone amide protection is applied for peptide synthesis according to the Fmoc‐protocols. In the present study, the usefulness of the recently proposed acid‐labile dicyclopropylmethyl (Dcpm) protectant was analyzed. Despite the significant steric hindrance of this bulky group, N‐terminal H‐(Dcpm)Gly‐peptides are quantitatively acylated by potent acylating agents, and alternatively the dipeptide Fmoc‐Asp(OtBu)‐(Dcpm)Gly‐OH derivative can be used as a building block. In contrast to the Hmb group, Dcpm is inert toward acylations, but is readily removed in the acid deprotection and resin‐cleavage step. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
The mechanisms of interfacial folding and membrane insertion of the Alzheimer's amyloid‐β fragment Aβ(25–35) and its less toxic mutant, N27A‐Aβ(25–35) and more toxic mutant, M35A‐Aβ(25–35), are investigated using replica–exchange molecular dynamics in an implicit water‐membrane environment. This study simulates the processes of interfacial folding and membrane insertion in a spontaneous fashion to identify their general mechanisms. Aβ(25–35) and N27A‐Aβ(25–35) peptides share similar mechanisms: the peptides are first located in the membrane hydrophilic region where their C‐terminal residues form helical structures. The peptides attempt to insert themselves into the membrane hydrophobic region using the C‐terminal or central hydrophobic residues. A small portion of peptides can successfully enter the membrane's hydrophobic core, led by their C‐terminal residues, through the formation of continuous helical structures. No detectable amount of M35A‐Aβ(25–35) peptides appeared to enter the membrane's hydrophobic core. The three studied peptides share a similar helical structure for their C‐terminal five residues, and these residues mainly buried within the membrane's hydrophobic region. In contrast, their N‐terminal properties are markedly different. With respect to the Aβ(25–35), the N27A‐Aβ(25–35) forms a more structured helix and is buried deeper within the membrane, which may result in a lower degree of aggregation and a lower neurotoxicity; in contrast, the less structured and more water‐exposed M35A‐Aβ(25–35) is prone to aggregation and has a higher neurotoxicity. Understanding the mechanisms of Aβ peptide interfacial folding and membrane insertion will provide new insights into the mechanisms of neurodegradation and may give structure‐based clues for rational drug design preventing amyloid associated diseases. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

7.
DKP formation is a serious side reaction during the solid‐phase synthesis of peptide acids containing either Pro or Gly at the C‐terminus. This side reaction not only leads to a lower overall yield, but also to the presence in the reaction crude of several deletion peptides lacking the first amino acids. For the preparation of protected peptides using the Fmoc/tBu strategy, the use of a ClTrt‐Cl‐resin with a limited incorporation of the C‐terminal amino acid is the method of choice. The use of resins with higher loading levels leads to more impure peptide crudes. The use of HPLC‐ESMS is a useful method for analysing complex samples, such as those formed when C‐terminal Pro peptides are prepared by non‐optimized solid‐phase strategies. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The design and synthesis of cyclic mimetics of VCAM‐1 protein that reproduce the integrin‐binding domain are presented. The unprotected peptide precursor 37 – 43 , Thr‐Gln‐Ile‐Asp‐Ser‐Pro‐Leu, was grafted onto functional templates of type naphthalene, biphenyl and benzyl through the chemoselective formation of C‐ and N‐terminal oximes resulting in a mixture of four isomeric forms due to synanti isomerism of the oxime bonds. Some isomers could be monitored by HPLC and identified by NMR. The molecule containing a naphthalene‐derived template was found to inhibit the VCAM‐1/VLA‐4 interaction more efficiently than previously reported for sulfur‐bridged cyclic peptides containing similar sequences. The finding confirms the importance of incorporating conformational constraints between the terminal ends of the peptide loop 37 – 43 in the design of synthetic inhibitors of the VCAM‐1/integrin interaction. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
The protein tyrosine phosphatase SHP‐1 plays an important role in many physiological and pathophysiological processes. This phosphatase is activated through binding of ligands to its SH2‐domains, mainly to the N‐terminal one. Based on a theoretical docking model, backbone‐to‐side chain cyclized octapeptides were designed as ligands. Assembly of such modelled structures required the synthesis of N‐functionalized tyrosine derivatives and their incorporation into the sequence. Because of difficulties encountered in the condensation of N‐protected amino acids to the N‐alkylated tyrosine‐peptide we synthesized and used preformed dipeptide building units. As all attempts to obtain phosphorylated dipeptide units failed, the syntheses had to be performed with a free phenolic function. Use of different N‐alkyl or cycloalkyl residues in the N‐functionalized side chains allowed to investigate the effect of ring size, flexibility and hydrophobicity of formed lactam bridges on stimulatory activity. All tested linear and cyclic octapeptides stimulate the phosphatase activity of SHP‐1. Stimulatory activities of cyclic ligands increase with the chain length of the lactam bridges resulting in increased flexibility and better entropic preformation of the binding conformation. The strong activity of some cyclic octapeptides supports the modelled structure. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Co‐administration of beta‐lactam antibiotics and beta‐lactamase inhibitors has been a favored treatment strategy against beta‐lactamase‐mediated bacterial antibiotic resistance, but the emergence of beta‐lactamases resistant to current inhibitors necessitates the discovery of novel non‐beta‐lactam inhibitors. Peptides derived from the Ala46–Tyr51 region of the beta‐lactamase inhibitor protein are considered as potent inhibitors of beta‐lactamase; unfortunately, peptide delivery into the cell limits their potential. The properties of cell‐penetrating peptides could guide the design of beta‐lactamase inhibitory peptides. Here, our goal is to modify the peptide with the sequence RRGHYY that possesses beta‐lactamase inhibitory activity under in vitro conditions. Inspired by the work on the cell‐penetrating peptide pVEC, our approach involved the addition of the N‐terminal hydrophobic residues, LLIIL, from pVEC to the inhibitor peptide to build a chimera. These residues have been reported to be critical in the uptake of pVEC. We tested the potential of RRGHYY and its chimeric derivative as a beta‐lactamase inhibitory peptide on Escherichia coli cells and compared the results with the action of the antimicrobial peptide melittin, the beta‐lactam antibiotic ampicillin, and the beta‐lactamase inhibitor potassium clavulanate to get mechanistic details on their action. Our results show that the addition of LLIIL to the N‐terminus of the beta‐lactamase inhibitory peptide RRGHYY increases its membrane permeabilizing potential. Interestingly, the addition of this short stretch of hydrophobic residues also modified the inhibitory peptide such that it acquired antimicrobial property. We propose that addition of the hydrophobic LLIIL residues to the peptide N‐terminus offers a promising strategy to design novel antimicrobial peptides in the battle against antibiotic resistance. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
During the final step of t‐Boc/Bzl, solid‐phase peptide synthesis (SPPS)‐protecting groups from amino acids (aa) side chains must be removed from the target peptides during cleavage from the solid support . These reaction steps involve hydrolysis with hydrogen fluoride (HF) in the presence of a nucleophile (scavenger), whose function is to trap the carbocations produced during SN1‐type reactions. Five peptide sequences were synthesised for evaluating p‐methoxyphenol effectiveness as a potent scavenger. After the synthesis, the resin–peptide was then separated into two equal parts to be cleaved using two scavengers: conventional reactive p‐cresol (reported in the literature as an effective acyl ion eliminator) and p‐methoxyphenol (hypothesised as fulfilling the same functions as the routinely used scavenger). Detailed analysis of the electrostatic potential map (EPM) revealed similarities between these two nucleophiles, regarding net atomic charge, electron density distribution, and similar pKa values. Good scavenger efficacy was observed by chromatography and mass spectrometry results for the synthesised molecules, which revealed that p‐methoxyphenol can be used as a potent scavenger during SPPS by t‐Boc/Bzl strategy, as similar results were obtained using the conventional scavenger.  相似文献   

12.
Recombinant protein expression and purification remains a central need for biotechnology. Herein, the authors report a streamlined protein and peptide purification strategy using short self‐assembling peptides and a C‐terminal cleavage intein. In this strategy, the fusion protein is first expressed as an aggregate induced by the self‐assembling peptide. Upon simple separation, the target protein or peptide with an authentic N‐terminus is then released in the solution by intein‐mediated cleavage. Different combinations of four self‐assembling peptides (ELK16, L6KD, FK and FR) with three inteins (Sce VMA, Mtu ΔI‐CM and Ssp DnaB) were explored. One protein and two peptides were used as model polypeptides to test the strategy. The intein Mtu ΔI‐CM, which has pH‐shift inducible cleavage, was found to work well with three self‐assembling peptides (L6KD, FR, FK). Using this intein gave a yield of protein or peptide comparable with that from other more established strategies, such as the Trx‐strategy, but in a simpler and more economical way. This strategy provides a simple and efficient method by which to prepare proteins and peptides with an authentic N‐terminus, which is especially effective for peptides of 30‐100 amino acids in length that are typically unstable and susceptible to degradation in Escherichia coli.  相似文献   

13.
N‐terminal modification of peptides by unnatural amino acids significantly affects their enzymatic stability, conformational properties and biological activity. Application of N‐amidino‐amino acids, positively charged under physiological conditions, can change peptide conformation and its affinity to the corresponding receptor. In this article, we describe synthesis of short peptides, containing a new building block—N‐amidino‐pyroglutamic acid. Although direct guanidinylation of pyroglutamic acid and oxidation of N‐amidino‐proline using RuO4 did not produce positive results, N‐amidino‐Glp‐Phe‐OH was synthesized on Wang polymer by cyclization of α‐guanidinoglutaric acid residue. In the course of synthesis, it was found that literature procedure of selective Boc deprotection using TMSOTf/TEA reagent is accompanied by concomitant side reaction of triethylamine alkylation by polymer linker fragment. It should be mentioned that independently from cyclization time and coupling agent (DIC or HCTU), the lactam formation was incomplete. Separation of the cyclic product from the linear precursor was achieved by HPLC in ammonium formate buffer at pH 6. HPLC analysis showed N‐amidino‐Glp‐Phe‐OH stability at acidic and physiological pH and fast ring opening in water solution at pH 9. The suggested method of N‐amidino‐Glp residue formation can be applied in the case of short peptide chains, whereas synthesis of longer ones will require fragment condensation approach. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
A core‐shell type polymer support for solid‐phase peptide synthesis has been developed for high coupling efficiency of peptides and versatile applications such as on‐bead bioassays. Although various kinds of polymer supports have been developed, they have their own drawbacks including poor accessibility of reagents and incompatibility in aqueous solution. In this paper, we prepared hydrophilic tri(ethylene glycol) (TEG) grafted core‐shell type polymer supports (TEG SURE) for efficient solid‐phase peptide synthesis and on‐bead bioassays. TEG SURE was prepared by grafting TEG derivative on the surface of AM PS resin via biphasic diffusion control method and subsequent acetylation of amine groups which are located at the core region of AM PS resin. The performance of TEG SURE was evaluated by synthesizing several peptides. Three points can be highlighted: (1) easy control of loading level of TEG, (2) improved efficiency of peptide synthesis compared with the conventional resins, and (3) applicability of on‐bead bioassays.  相似文献   

15.
Protein C‐termini study is still a challenging task and far behind its counterpart, N‐termini study. MS based C‐terminomics study is often hampered by the low ionization efficiency of C‐terminal peptides and the lack of efficient enrichment methods. We previously optimized the C‐terminal amine‐based isotope labeling of substrates (C‐TAILS) method and identified 369 genuine protein C‐termini in Escherichia coli. A key limitation of C‐TAILS is that the prior protection of amines and carboxylic groups at protein level makes Arg‐C as the only specific enzyme in practice. Herein, we report an approach combining multi‐enzyme digestion and C‐TAILS, which significantly increases the identification rate of C‐terminal peptides and consequently improves the applicability of C‐TAILS in biological studies. We carry out a systematic study and confirm that the omission of the prior amine protection at protein level has a negligible influence and allows the application of multi‐enzyme digestion. We successfully apply five different enzyme digestions to C‐TAILS, including trypsin, Arg‐C, Lys‐C, Lys‐N, and Lysarginase. As a result, we identify a total of 722 protein C‐termini in E. coli, which is at least 66% more than the results using any single enzyme. Moreover, the favored enzyme and enzyme combination are discovered. Data are available via ProteomeXchange with identifier PXD004275.  相似文献   

16.
Antimicrobial peptides are recognized candidates with pharmaceutical potential against epidemic emerging multi‐drug resistant bacteria. In this study, we use nuclear magnetic resonance spectroscopy and molecular dynamics simulations to determine the unknown structure and evaluate the interaction with dodecylphosphatidylcholine (DPC) and sodium dodecylsulphate (SDS) micelles with three W6‐Hylin‐a1 analogs antimicrobial peptides (HyAc, HyK, and HyD). The HyAc, HyK, and HyD bound to DPC micelles are all formed by a unique α‐helix structure. Moreover, all peptides reach the DPC micelles' core, which thus suggests that the N‐terminal modifications do not influence the interaction with zwiterionic surfaces. On the other hand, only HyAc and HyK peptides are able to penetrate the SDS micelle core while HyD remains always at its surface. The stability of the α‐helical structure, after peptide‐membrane interaction, can also be important to the second step of peptide insertion into the membrane hydrophobic core during permeabilization. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
We describe a mass spectrometric method for distinguishing between free and modified forms of the C‐terminal carboxyl group of peptides and proteins, in combination with chemical approaches for the isolation of C‐terminal peptides and site‐specific derivatization of the C‐terminal carboxyl group. This method could most advantageously be exploited to discriminate between peptides having C‐terminal carboxyl groups in the free (COOH) and amide (CONH2) forms by increasing their mass difference from 1 to 14 Da by selectively converting the free carboxyl group into methylamide (CONHCH3). This method has been proven to be applicable to peptides containing aspartic and glutamic acids, because all the carboxyl groups except the C‐terminal one are inert to derivatization, according to oxazolone chemistry. The efficiency of the method is illustrated by a comparison of the peaks of processed peptides obtained from a mixture of adrenomedullin, calcitonin, and BSA. Among these components of the mixture, only the C‐terminal peptide of BSA exhibited the mass shift of 13 Da upon treatment, eventually unambiguously validating the C‐terminal amide structures of adrenomedullin and calcitonin. The possibility of extending this method for the analysis of C‐terminal PTMs is also discussed.  相似文献   

18.
The polyproline‐II helix is the most extended naturally occurring helical structure and is widely present in polar, exposed stretches and “unstructured” denatured regions of polypeptides. Can it be hydrophobic? In this study, we address this question using oligomeric peptides formed by a hydrophobic proline analogue, (2S,3aS,7aS)‐octahydroindole‐2‐carboxylic acid (Oic). Previously, we found the molecular principles underlying the structural stability of the polyproline‐II conformation in these oligomers, whereas the hydrophobicity of the peptide constructs remains to be examined. Therefore, we investigated the octan‐1‐ol/water partitioning and inclusion in detergent micelles of the oligo‐Oic peptides. The results showed that the hydrophobicity is remarkably enhanced in longer oligomeric sequences, and the oligo‐Oic peptides with 3 to 4 residues and higher are specific towards hydrophobic environments. This contrasts significantly to the parent oligoproline peptides, which were moderately hydrophilic. With these findings, we have demonstrated that the polyproline‐II structure is compatible with nonpolar media, whereas additional manipulations of the terminal functionalities feature solubility in extremely nonpolar solvents such as hexane.  相似文献   

19.
Although the N‐terminal region in human apolipoprotein (apo) A‐I is thought to stabilize the lipid‐free structure of the protein, its role in lipid binding is unknown. Using synthetic fragment peptides, we examined the lipid‐binding properties of the first 43 residues (1–43) of apoA‐I in comparison with residues 44–65 and 220–241, which have strong lipid affinity in the molecule. Circular dichroism measurements demonstrated that peptides corresponding to each segment have potential propensity to form α‐helical structure in trifluoroethanol. Spectroscopic and thermodynamic measurements revealed that apoA‐I (1–43) peptide has the strong ability to bind to lipid vesicles and to form α‐helical structure comparable to apoA‐I (220–241) peptide. Substitution of Tyr‐18 located at the center of the most hydrophobic region in residues 1–43 with a helix‐breaking proline resulted in the impaired lipid binding, indicating that the α‐helical structure in this region is required to trigger the lipid binding. In contrast, apoA‐I (44–65) peptide exhibited a lower propensity to form α‐helical structure upon binding to lipid, and apoA‐I (44–65/S55P) peptide exhibited diminished, but not completely impaired, lipid binding, suggesting that the central region of residues 44–65 is not pivotally involved in the formation of the α‐helical structure and lipid binding. These results indicate that the most N‐terminal region of apoA‐I molecule, residues 1–43, contributes to the lipid interaction of apoA‐I through the hydrophobic helical residues. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Globally, death due to cancers is likely to rise to over 20 million by 2030, which has created an urgent need for novel approaches to anticancer therapies such as the development of host defence peptides. Cn‐AMP2 (TESYFVFSVGM), an anionic host defence peptide from green coconut water of the plant Cocos nucifera, showed anti‐proliferative activity against the 1321N1 and U87MG human glioma cell lines with IC50 values of 1.25 and 1.85 mM, respectively. The membrane interactive form of the peptide was found to be an extended conformation, which primarily included β‐type structures (levels > 45%) and random coil architecture (levels > 45%). On the basis of these and other data, it is suggested that the short anionic N‐terminal sequence (TES) of Cn‐AMP2 interacts with positively charged moieties in the cancer cell membrane. Concomitantly, the long hydrophobic C‐terminal sequence (YFVFSVGM) of the peptide penetrates the membrane core region, thereby driving the translocation of Cn‐AMP2 across the cancer cell membrane to attack intracellular targets and induce anti‐proliferative mechanisms. This work is the first to demonstrate that anionic host defence peptides have activity against human glioblastoma, which potentially provides an untapped source of lead compounds for development as novel agents in the treatment of these and other cancers. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  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