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 656 毫秒
1.
An efficient peptide purification strategy is established, comprising the selective reaction of an N‐terminal N‐(methoxy)glycine residue of the peptide and isothiocyanato‐functionalized resins, and subsequent Edman degradation. These reactions take place in acidic media; in particular, the Edman degradation proceeds smoothly in media containing more than 50% trifluoroacetic acid (v/v). These acidic conditions offer increased solubility, making them advantageous for the purification of hydrophobic and aggregation‐prone peptides. The effectiveness of this method, together with scope and limitations, is demonstrated using model peptides and the practical purification of the loop region of the human dopamine D2 receptor long isoform (residues 240–272). Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Concentrated formic acid is among the most effective solvents for protein solubilization. Unfortunately, this acid also presents a risk of inducing chemical modifications thereby limiting its use in proteomics. Previous reports have supported the esterification of serine and threonine residues (O‐formylation) for peptides incubated in formic acid. However as shown here, exposure of histone H4 to 80% formic (1 h, 20oC) induces N‐formylation of two independent lysine residues. Furthermore, incubating a mixture of Escherichia coli proteins in formic acid demonstrates a clear preference toward lysine modification over reactions at serine/threonine. N‐formylation accounts for 84% of the 225 uniquely identified formylation sites. To prevent formylation, we provide a detailed investigation of reaction conditions (temperature, time, acid concentration) that define the parameters permitting the use of concentrated formic acid in a proteomics workflow for MS characterization. Proteins can be maintained in 80% formic acid for extended periods (24 h) without inducing modification, so long as the temperature is maintained at or below –20oC.  相似文献   

3.
Biophysical studies on amyloidogenic and aggregation‐prone peptides often require large quantities of material. However, solid‐phase synthesis, handling, and purification of peptides often present challenges on these scales. Recombinant expression is an attractive alternative because of its low cost, the ability to isotopically label the peptides, and access to sequences exceeding ~50 residues. However, expression systems that seek to solubilize amyloidogenic peptides suffer from low yields, difficult optimizations, and isolation challenges. We present a general strategy for expressing and isolating amyloidogenic peptides in Escherichia coli by fusion to a polypeptide that drives the expression of attached peptides into bacterial inclusion bodies. This scheme minimizes toxicity during bacterial growth and enables the processing and handling of the peptides in denaturing solutions. Immobilized metal affinity chromatography, reverse phase HPLC, and cyanogen bromide cleavage are used to isolate the peptide, followed by further reverse phase HPLC, which yields milligram quantities of the purified peptide. We demonstrate that driving the peptides into inclusion bodies using fusion to BCL‐XL‐1/2 is a general strategy for their expression and isolation, as exemplified by the production of 11 peptides species.  相似文献   

4.
Antimicrobial peptides are promising alternative to traditional antibiotics and antitumor drugs for the battle against new antibiotic resistant bacteria strains and cancer maladies. The study of their structural and dynamics properties at physiological conditions can help to understand their stability, delivery mechanisms, and activity in the human body. In this article, we have used molecular dynamics simulations to study the effects of solvent environment, temperature, ions concentration, and peptide concentration on the structural properties of the antimicrobial hybrid peptide Cecropin A–Magainin 2. In TFE/water mixtures, the structure of the peptide retained α‐helix contents and an average hinge angle in close agreement with the experimental NMR and CD measurements reported in literature. Compared to the TFE/water mixture, the peptide simulated at the same ionic concentration lost most of its α‐helix structure. The increase of peptide concentration at both 300 and 310 K resulted in the peptide aggregation. The peptides in the complex retained the initial N‐ter α‐helix segment during all the simulation. The α‐helix stabilization is further enhanced in the high salt concentration simulations. The peptide aggregation was not observed in TFE/water mixture simulations and, the peptide aggregate, obtained from the water simulation, simulated in the same conditions did dissolve within few tens of nanoseconds. The results of this study provide insights at molecular level on the structural and dynamics properties of the CA‐MA peptide at physiological and membrane mimic conditions that can help to better understand its delivery and interaction with biological interfaces. © 2014 Wiley Periodicals, Inc. Biopolymers 103: 1–14, 2015.  相似文献   

5.
6.
The N‐terminal 1–34 segments of both parathyroid hormone (PTH) and parathyroid hormone‐related protein (PTHrP) bind and activate the same membrane receptor in spite of major differences between the two hormones in their amino acid sequence. Recently, it was shown that in (1–34)PTH/PTHrP segmental hybrid peptides, the N‐terminal 1–14 segment of PTHrP is incompatible with the C‐terminal 15–34 region of PTH leading to substantial reduction in potency. The sites of incompatibility were identified as positions 5 in PTH and 19 in PTHrP. In the present paper we describe the synthesis, biological evaluation, and conformational characterization of two point‐mutated PTH/PTHrP 1–34 hybrids in which the arginine residues at positions 19 and 21 of the native sequence of PTHrP have been replaced by valine (hybrid V21) and glutamic acid (hybrid E19), respectively, taken from the PTH sequence. Hybrid V21 exhibits both high receptor affinity and biological potency, while hybrid E19 binds weakly and is poorly active. The conformational properties of the two hybrids were studied in aqueous solution containing dodecylphosphocholine (DPC) micelles and in water/2,2,2‐trifluoroethanol (TFE) mixtures. Upon addition of TFE or DPC micelles to the aqueous solution, both hybrids undergo a coil‐helix transition. The maximum helix content in 1 : 1 water/TFE, obtained by CD data for both hybrids, is ∼ 80%. In the presence of DPC micelles, the maximum helix content is ∼ 40%. The conformational properties of the two hybrids in the micellar system were further investigated by combined 2D‐nmr, distance geometry (DG), and molecular dynamics (MD) calculations. The common structural motif, consisting of two helical segments located at N‐ and C‐termini, was observed in both hybrids. However, the biologically potent hybrid V21 exhibits two flexible sites, centered at residues 12 and 19 and connecting helical segments, while the flexibility sites in the weakly active hybrid E19 are located at position 11 and in the sequence 20–26. Our findings support the hypothesis that the presence and location of flexibility points between helical segments are essential for enabling the active analogs to fold into the bioactive conformation upon interaction with the receptor. © 1999 John Wiley & Sons, Inc. Biopoly 50: 525–535, 1999  相似文献   

7.
Murraya koenigii miraculin‐like protein (MKMLP) gradually precipitates below pH 7.5. Here, we explore the basis for this aggregation by identifying the aggregation‐prone regions via comparative analysis of crystal structures acquired at several pH values. The prediction of aggregation‐prone regions showed the presence of four short peptides either in beta sheets or loops on surface of the protein. These peptides were distributed in two patches far apart on the surface. Comparison of crystal structures of MKMLP, determined at 2.2 Å resolution in pH 7.0 and 4.6 in the present study and determined at 2.9 Å in pH 8.0 in an earlier reported study, reveal subtle conformational differences resulting in gradual exposure of aggregation‐prone regions. As the pH is lowered, there are alterations in ionic interactions within the protein interactions of the chain with water molecules and exposure of hydrophobic residues. The analysis of symmetry‐related molecular interfaces involving one patch revealed shortening of nonpolar intermolecular contacts as the pH decreased. In particular, a decrease in the intermolecular distance between Trp103 of the aggregation‐prone peptide WFITTG (103–108) unique to MLPs was observed. These results demonstrated that aggregation occurs due to the cumulative effect of the changes in interactions in two aggregation‐prone defined regions. Proteins 2014; 82:830–840. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
The 173–195 segment corresponding to the helix 2 of the globular PrP domain is a good candidate to be one of the several ‘spots’ of intrinsic structural flexibility, which might induce local destabilization and concur to protein transformation, leading to aggregation‐prone conformations. Here, we report CD and NMR studies on the α2‐helix‐derived peptide of maximal length (hPrP[180–195]) that is able to exhibit a regular structure different from the prevalently random arrangement of other α2‐helix‐derived peptides. This peptide, which has previously been shown to be affected by buffer composition via the ion charge density dependence typical of Hofmeister effects, corresponds to the C‐terminal sequence of the PrPC full‐length α2‐helix and includes the highly conserved threonine‐rich 188–195 segment. At neutral pH, its conformation is dominated by β‐type contributions, which only very strong environmental modifications are able to modify. On TFE addition, an increase of α‐helical content can be observed, but a fully helical conformation is only obtained in neat TFE. However, linking of the 173–179 segment, as occurring in wild‐type and mutant peptides corresponding to the full‐length α2‐helix, perturbs these intrinsic structural propensities in a manner that depends on whether the environment is water or TFE. Overall, these results confirm that the 180–195 parental region in hPrPC makes a strong contribution to the chameleon conformational behavior of the segment corresponding to the full‐length α2‐helix, and could play a role in determining structural rearrangements of the entire globular domain. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Glutamic acid–rich peptides are crucial to a variety of biological processes, including glutamatergic neurotransmission and immunological defense. Glutamic acid sequences often exhibit unusual organization into β2‐type sheets, where bifurcated H bonds formed between glutamic acid side chains and NH in amide bonds on adjacent βstrands play a paramount role for stabilizing the molecular assembly. Herein, we investigate the self‐assembly and supramolecular structure of simplified models consisting of alternating glutamic acid/phenylalanine residues. Small‐angle X‐ray scattering and atomic force microscopy show that the aggregation pathway is characterized by the formation of small oligomers, followed by coalescence into nanofibrils and nanotapes. Amyloidogenic features are further demonstrated through fiber X‐ray diffraction, which reveal molecular packing according to cross‐β patterns, where βstrands appear perpendicularly oriented to the long axis of nanofibrils and nanotapes. Nanoscale infrared spectroscopy from individual nanoparticles on dried samples shows a remarkable decrease of β2‐sheet content, accompanied by growth of standard β‐sheet fractions, indicating a β2‐to‐β1 transition as a consequence of the release of solvent from the interstices of peptide assemblies. Our findings highlight the key role played by water molecules in mediating H‐bond formation in β2‐sheets commonly found in amyloidogenic glutamic acid–rich aggregates.  相似文献   

10.
Adrenergic receptors are integral membrane proteins involved in cellular signalling that belong to the G protein-coupled receptors. Synthetic peptides resembling the putative transmembrane (TM) segments TM4, TM6 and TM7, of the human α2-adrenergic receptor subtype C10 (P08913) and defined lipid vesicles were used to assess protein-lipid interactions that might be relevant to receptor structure/function. P6 peptide contains the hydrophobic core of TM6 plus the N-terminal hydrophilic motif REKR, while peptides P4 and P7 contained just the hydrophobic stretches of TM4 and TM7, respectively. All the peptides increase their helical tendency at moderate concentrations of TFE (30–50%) and in presence of 1,2-dielaidoyl-sn-glycero-3-phosphatidylethanolamine (DEPE) lipids. However, only P6 displays up to 19% of α-helix in the presence of just the DEPE lipids, evidences a transmembrane orientation and stabilizes the Lα lipid phase. Conversely, P4 and P7 peptides form only stable β-sheet structures in DEPE and favour the non-lamellar, inverted hexagonal (HII) phase of DEPE by lowering its phase transition temperature. This study highlights the potential of using synthetic peptides derived from the amino acid sequence in the native proteins as templates to understand the behaviour of the transmembrane segments and underline the importance of interfacial anchoring interactions to meet hydrophobic matching requirements and define membrane organization.  相似文献   

11.
Introduction – Flavonoids, the primary constituents of the petals of Nelumbo nucifera, are known to have antioxidant properties and antibacterial bioactivities. However, efficient methods for the preparative isolation and purification of flavonoids from this plant are not currently available. Objective – To develop an efficient method for the preparative isolation and purification of flavonoids from the petals of N. nucifera by high‐speed counter‐current chromatography (HSCCC). Methodology – Following an initial clean‐up step on a polyamide column, HSCCC was utilised to separate and purify flavonoids. Purities and identities of the isolated compounds were established by HPLC‐PAD, ESI‐MS, 1H‐NMR and 13C‐NMR. Results – The separation was performed using a two‐phase solvent system composed of ethyl acetate–methanol–water–acetic acid (4 : 1 : 5 : 0.1, by volume), in which the upper phase was used as the stationary phase and the lower phase was used as the mobile phase at a flow‐rate of 1.0 mL/min in the head‐to‐tail elution mode. Ultimately, 5.0 mg syringetin‐3‐O‐β‐d‐glucoside, 6.5 mg quercetin‐3‐O‐β‐d‐glucoside, 12.8 mg isorhamnetin‐3‐O‐β‐d‐glucoside and 32.5 mg kaempferol‐3‐O‐β‐d‐glucoside were obtained from 125 mg crude sample. Conclusion – The combination of HSCCC with a polyamide column is an efficient method for the preparative separation and purification of flavonoids from the petals of N. nucifera. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

12.
Snakin‐1 (SN‐1) is a cysteine‐rich plant antimicrobial peptide and the first purified member of the snakin family. SN‐1 shows potent activity against a wide range of microorganisms, and thus has great biotechnological potential as an antimicrobial agent. Here, we produced recombinant SN‐1 in Escherichia coli by a previously developed coexpression method using an aggregation‐prone partner protein. Our goal was to increase the productivity of SN‐1 via the enhanced formation of insoluble inclusion bodies in E. coli cells. The yield of SN‐1 by the coexpression method was better than that by direct expression in E. coli cells. After refolding and purification, we obtained several milligrams of functionally active SN‐1, the identity of which was verified by MALDI‐TOF MS and NMR studies. The purified recombinant SN‐1 showed effective antimicrobial activity against test organisms. Our studies indicate that the coexpression method using an aggregation‐prone partner protein can serve as a suitable expression system for the efficient production of functionally active SN‐1. © 2017 American Institute of Chemical Engineers Biotechnol. Prog., 33:1520–1528, 2017  相似文献   

13.
Structural modification of the peptide backbone via N‐methylation is a powerful tool to modulate the pharmacokinetic profile and biological activity of peptides. Here we describe a rapid and highly efficient microwave(MW)‐assisted Fmoc/tBu solid‐phase method to prepare short chain N‐methyl‐rich peptides, using Rink amide p‐methylbenzhydrylamine (MBHA) resin as solid‐phase support. This method produces peptides in high yield and purity, and reduces the time required for Fmoc‐N‐methyl amino acid coupling. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Aza‐peptides have been used as tools for studying SARs in programs aimed at drug discovery and chemical biology. Protected aza‐dipeptides were synthesized by a solution‐phase submonomer approach featuring alkylation of N‐terminal benzophenone semicarbazone aza‐Gly‐Xaa dipeptides using different alkyl halides in the presence of potassium tert‐butoxide as base. Benzophenone protected aza‐dipeptide tert‐butyl ester 31c was selectively deprotected at the C‐terminal ester or N‐terminal hydrazone to afford, respectively, aza‐dipeptide acid and amine building blocks 36c and 40c, which were introduced into longer aza‐peptides. Alternatively, removal of the benzophenone semicarbazone protection from aza‐dipeptide methyl esters 29a–c led to intramolecular cyclization to produce aza‐DKPs 39a–c. In light of the importance of aza‐peptides and DKPs as therapeutic agents and probes of biological processes, this diversity‐oriented solution‐phase approach may provide useful tools for studying peptide science. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
Peptide P11‐4 (QQRFEWEFEQQ) was designed to self‐assemble to form β‐sheets and nematic gels in the pH range 5–7 at concentrations ≥12.6 mM in water. This self‐assembly is reversibly controlled by adjusting the pH of the solvent. It can also self‐assemble into gels in biological media. This together with its biocompatibility and biodegradability make P11‐4 an attractive building block for the fabrication of nanoscale materials with uses in, for example, tissue engineering. A limitation to large‐scale production of such peptides is the high cost of solid phase chemical synthesis. We describe expression of peptide P11‐4 in the bacterium Escherichia coli from constructs carrying tandem repeats of the peptide coding sequence. The vector pET31b+ was used to express P11‐4 repeats fused to the ketosteroid isomerase protein which accumulates in easily recoverable inclusion bodies. Importantly, the use of auto‐induction growth medium to enhance cell density and protein expression levels resulted in recovery of 2.5 g fusion protein/L culture in both shake flask and batch fermentation. Whole cell detergent lysis allowed recovery of inclusion bodies largely composed of the fusion protein. Cyanogen bromide cleavage followed by reverse phase HPLC allowed purification of the recombinant peptide with a C‐terminal homoserine lactone (rP11‐4(hsl)). This recombinant peptide formed pH dependent hydrogels, displayed β‐structure measured by circular dichroism and fibril formation observed by transmission electron microscopy. Biotechnol. Bioeng. 2009;103: 241–251. © 2009 Wiley Periodicals, Inc.  相似文献   

16.
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   

17.
We previously studied a 16‐amino acid‐residue fragment of the C‐terminal β‐hairpin of the B3 domain (residues 46–61), [IG(46–61)] of the immunoglobulin binding protein G from Streptoccocus, and found that hydrophobic interactions and the turn region play an important role in stabilizing the structure. Based on these results, we carried out systematic structural studies of peptides derived from the sequence of IG (46–61) by systematically shortening the peptide by one residue at a time from both the C‐ and the N‐terminus. To determine the structure and stability of two resulting 12‐ and 14‐amino acid‐residue peptides, IG(48–59) and IG(47–60), respectively, we carried out circular dichroism, NMR, and calorimetric studies of these peptides in pure water. Our results show that IG(48–59) possesses organized three‐dimensional structure stabilized by hydrophobic interactions (Tyr50–Phe57 and Trp48–Val59) at T = 283 and 305 K. At T = 313 K, the structure breaks down because of increased chain entropy, but the turn region is preserved in the same position observed for the structure of the whole protein. The breakdown of structure occurs near the melting temperature of this peptide (Tm = 310 K) measured by differential scanning calorimetry (DSC). The melting temperature of IG(47–60) determined by DSC is Tm = 330 K and its structure is similar to that of the native β‐hairpin at all (lower) temperatures examined (283–313 K). Both of these truncated sequences are conserved in all known amino acid sequences of the B domains of the immunoglobulin binding protein G from bacteria. Thus, this study contributes to an understanding of the mechanism of folding of this whole family of proteins, and provides information about the mechanism of formation and stabilization of a β‐hairpin structural element. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

18.
We are developing a rapid, time‐resolved method using laser‐activated cross‐linking to capture protein‐peptide interactions as a means to interrogate the interaction of serum proteins as delivery systems for peptides and other molecules. A model system was established to investigate the interactions between bovine serum albumin (BSA) and 2 peptides, the tridecapeptide budding‐yeast mating pheromone (α‐factor) and the decapeptide human gonadotropin‐releasing hormone (GnRH). Cross‐linking of α‐factor, using a biotinylated, photoactivatable p‐benzoyl‐L‐phenylalanine (Bpa)–modified analog, was energy‐dependent and achieved within seconds of laser irradiation. Protein blotting with an avidin probe was used to detect biotinylated species in the BSA‐peptide complex. The cross‐linked complex was trypsinized and then interrogated with nano‐LC–MS/MS to identify the peptide cross‐links. Cross‐linking was greatly facilitated by Bpa in the peptide, but some cross‐linking occurred at higher laser powers and high concentrations of a non‐Bpa–modified α‐factor. This was supported by experiments using GnRH, a peptide with sequence homology to α‐factor, which was likewise found to be cross‐linked to BSA by laser irradiation. Analysis of peptides in the mass spectra showed that the binding site for both α‐factor and GnRH was in the BSA pocket defined previously as the site for fatty acid binding. This model system validates the use of laser‐activation to facilitate cross‐linking of Bpa‐containing molecules to proteins. The rapid cross‐linking procedure and high performance of MS/MS to identify cross‐links provides a method to interrogate protein‐peptide interactions in a living cell in a time‐resolved manner.  相似文献   

19.
Prosurfactant protein C (proSP‐C) is a 197‐residue integral membrane protein, in which the C‐terminal domain (CTC, positions 59–197) is localized in the endoplasmic reticulum (ER) lumen and contains a Brichos domain (positions 94–197). Mature SP‐C corresponds largely to the transmembrane (TM) region of proSP‐C. CTC binds to SP‐C, provided that it is in nonhelical conformation, and can prevent formation of intracellular amyloid‐like inclusions of proSP‐C that harbor mutations linked to interstitial lung disease (ILD). Herein it is shown that expression of proSP‐C (1–58), that is, the N‐terminal propeptide and the TM region, in HEK293 cells results in virtually no detectable protein, while coexpression of CTC in trans yields SDS‐soluble monomeric proSP‐C (1–58). Recombinant human (rh) CTC binds to cellulose‐bound peptides derived from the nonpolar TM region, but not the polar cytosolic part, of proSP‐C, and requires ≥5‐residues for maximal binding. Binding of rhCTC to a nonhelical peptide derived from SP‐C results in α‐helix formation provided that it contains a long TM segment. Finally, rhCTC and rhCTC Brichos domain shows very similar substrate specificities, but rhCTCL188Q, a mutation linked to ILD is unable to bind all peptides analyzed. These data indicate that the Brichos domain of proSP‐C is a chaperone that induces α‐helix formation of an aggregation‐prone TM region.  相似文献   

20.
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   

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