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.
Molecular docking of peptides to proteins can be a useful tool in the exploration of the possible peptide binding sites and poses. CABS‐dock is a method for protein–peptide docking that features significant conformational flexibility of both the peptide and the protein molecules during the peptide search for a binding site. The CABS‐dock has been made available as a web server and a standalone package. The web server is an easy to use tool with a simple web interface. The standalone package is a command‐line program dedicated to professional users. It offers a number of advanced features, analysis tools and support for large‐sized systems. In this article, we outline the current status of the CABS‐dock method, its recent developments, applications, and challenges ahead.  相似文献   

2.
Peptides with both an affinity for ZnO and the ability to generate ZnO nanoparticles have attracted attention for the self‐assembly and templating of nanoscale building blocks under ambient conditions with compositional uniformity. In this study, we have analyzed the specific binding sites of the ZnO‐binding peptide, EAHVMHKVAPRP, which was identified using a phage display peptide library. The peptide binding assay against ZnO nanoparticles was performed using peptides synthesized on a cellulose membrane using the spot method. Using randomized rotation of amino acids in the ZnO‐binding peptide, 125 spot‐synthesized peptides were assayed. The peptide binding activity against ZnO nanoparticles varied greatly. This indicates that ZnO binding does not depend on total hydrophobicity or other physical parameters of these peptides, but rather that ZnO recognizes the specific amino acid alignment of these peptides. In addition, several peptides were found to show higher binding ability compared with that of the original peptides. Identification of important binding sites in the EAHVMHKVAPRP peptide was investigated by shortened, stepwise sequence from both termini. Interestingly, two ZnO‐binding sites were found as 6‐mer peptides: HVMHKV and HKVAPR. The peptides identified by amino acid substitution of HKVAPR were found to show high affinity and specificity for ZnO nanoparticles. Biotechnol. Bioeng. 2010;106: 845–851. © 2010 Wiley Periodicals, Inc.  相似文献   

3.
Substrate binding to Hsp70 chaperones is involved in many biological processes, and the identification of potential substrates is important for a comprehensive understanding of these events. We present a multi‐scale pipeline for an accurate, yet efficient prediction of peptides binding to the Hsp70 chaperone BiP by combining sequence‐based prediction with molecular docking and MMPBSA calculations. First, we measured the binding of 15mer peptides from known substrate proteins of BiP by peptide array (PA) experiments and performed an accuracy assessment of the PA data by fluorescence anisotropy studies. Several sequence‐based prediction models were fitted using this and other peptide binding data. A structure‐based position‐specific scoring matrix (SB‐PSSM) derived solely from structural modeling data forms the core of all models. The matrix elements are based on a combination of binding energy estimations, molecular dynamics simulations, and analysis of the BiP binding site, which led to new insights into the peptide binding specificities of the chaperone. Using this SB‐PSSM, peptide binders could be predicted with high selectivity even without training of the model on experimental data. Additional training further increased the prediction accuracies. Subsequent molecular docking (DynaDock) and MMGBSA/MMPBSA‐based binding affinity estimations for predicted binders allowed the identification of the correct binding mode of the peptides as well as the calculation of nearly quantitative binding affinities. The general concept behind the developed multi‐scale pipeline can readily be applied to other protein‐peptide complexes with linearly bound peptides, for which sufficient experimental binding data for the training of classical sequence‐based prediction models is not available. Proteins 2016; 84:1390–1407. © 2016 Wiley Periodicals, Inc.  相似文献   

4.
A procedure for obtaining isotopically labeled peptides, by combining affinity chromatography, urea‐equilibrated gel filtration, and hydrophobic chromatography procedures, is presented using the Disabled‐2 (Dab2) sulfatide‐binding motif (SBM) as a proof of concept. The protocol is designed to isolate unstructured, membrane‐binding, recombinant peptides that co‐purify with bacterial proteins (e.g., chaperones). Dab2 SBM is overexpressed in bacteria as an isotopically labeled glutathione S‐transferase (GST) fusion protein using minimal media containing [15N] ammonium chloride as the nitrogen source. The fusion protein is purified using glutathione beads, and Dab2 SBM is released from GST using a specific protease. It is then dried, resuspended in urea to release the bound bacterial protein, and subjected to urea‐equilibrated gel filtration. Urea and buffer reagents are removed using an octadecyl column. The peptide is eluted with acetonitrile, dried, and stored at ?80 °C. Purification of Dab2 SBM can be accomplished in 6 days with a yield of ~2 mg/l of culture. The properties of Dab2 SBM can be studied in the presence of detergents using NMR spectroscopy. Although this method also allows for the purification of unlabeled peptides that co‐purify with bacterial proteins, the procedure is more relevant to isotopically labeled peptides, thus alleviating the cost of peptide production. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
To perform their various functions, protein surfaces often have to interact with each other in a specific way. Usually, only parts of a protein are accessible and can act as binding sites. Because proteins consist of polypeptide chains that fold into complex three‐dimensional shapes, binding sites can be divided into two different types: linear sites that follow the primary amino acid sequence and discontinuous binding sites, which are made up of short peptide fragments that are adjacent in spatial proximity. Such discontinuous binding sites dominate protein–protein interactions, but are difficult to identify. To meet this challenge, we combined a computational, structure‐based approach and an experimental, high‐throughput method. SUPERFICIAL is a program that uses protein structures as input and generates peptide libraries to represent the protein's surface. A large number of the predicted peptides can be simultaneously synthesised applying the SPOT technology. The results of a binding assay subsequently help to elucidate protein–protein interactions; the approach is applicable to any kind of protein. The crystal structure of the complex of hen egg lysozyme with the well‐characterised murine IgG1 antibody HyHEL‐5 is available, and the complex is known to have a discontinuous binding site. Using SUPERFICIAL, the entire surface of lysozyme was translated into a peptide library that was synthesised on a cellulose membrane using the SPOT technology and tested against the HyHEL‐5 antibody. In this way, it was possible to identify two peptides (longest common sequence and peptide 19) that represented the discontinuous epitope of lysozyme. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

6.
A new type of microtiter plate capable of binding biomolecules covalently in a one step procedure was used to map linear B‐cell epitopes in two different proteins using a peptide‐based solid phase immunoassay. The method was compared with a conventional immobilization method using passive adsorption to microtiter plates. An array of 15‐mer peptides, overlapping by five amino acids, representing the entire sequences of ubiquitin and murine tumor necrosis factor‐α, respectively, was synthesized. The peptides were immobilized covalently using the new, specialized microtiter plates or non‐covalently using conventional ELISA microtiter plates of the high binder type. Subsequently, specific antisera to ubiquitin or murine tumor necrosis factor‐α were added to identify potential linear B‐cell epitopes. All peptides, which were recognized on the conventional microtiter plates, were also recognized on the plates with the covalently bound peptides. In addition, the covalent immobilization method revealed epitopes that were not identified using the method for non‐covalent binding although the peptides were in fact present on the non‐covalent binding surface. The interaction with the hydrophobic surface of the conventional microtiter plate apparently interfered negatively with antibody recognition. The covalently binding microtiter plates described here could be useful for identification of new B‐cell epitopes in protein antigens. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
Oligopeptide‐binding protein A (OppA) from Lactococcus lactis binds peptides of an exceptionally wide range of lengths (4–35 residues), with no apparent sequence preference. Here, we present the crystal structures of OppA in the open‐ and closed‐liganded conformations. The structures directly explain the protein's phenomenal promiscuity. A huge cavity allows binding of very long peptides, and a lack of constraints for the position of the N and C termini of the ligand is compatible with binding of peptides with varying lengths. Unexpectedly, the peptide's amino‐acid composition (but not the exact sequence) appears to have a function in selection, with a preference for proline‐rich peptides containing at least one isoleucine. These properties can be related to the physiology of the organism: L. lactis is auxotrophic for branched chain amino acids and favours proline‐rich caseins as a source of amino acids. We propose a new mechanism for peptide selection based on amino‐acid composition rather than sequence.  相似文献   

8.
Several accurate prediction systems have been developed for prediction of class I major histocompatibility complex (MHC):peptide binding. Most of these are trained on binding affinity data of primarily 9mer peptides. Here, we show how prediction methods trained on 9mer data can be used for accurate binding affinity prediction of peptides of length 8, 10 and 11. The method gives the opportunity to predict peptides with a different length than nine for MHC alleles where no such peptides have been measured. As validation, the performance of this approach is compared to predictors trained on peptides of the peptide length in question. In this validation, the approximation method has an accuracy that is comparable to or better than methods trained on a peptide length identical to the predicted peptides. AVAILABILITY: The algorithm has been implemented in the web-accessible servers NetMHC-3.0: http://www.cbs.dtu.dk/services/NetMHC-3.0, and NetMHCpan-1.1: http://www.cbs.dtu.dk/services/NetMHCpan-1.1  相似文献   

9.
Protein p16INK4a (p16) is a well‐known biomarker for diagnosis of human papillomavirus (HPV) related cancers. In this work, we identify novel p16 binding peptides by using phage display selection method. A random heptamer phage display library was screened on purified recombinant p16 protein‐coated plates to elute only the bound phages from p16 surfaces. Binding affinity of the bound phages was compared with each other by enzyme‐linked immunosorbent assay (ELISA), fluorescence imaging technique, and bioinformatic computations. Binding specificity and binding selectivity of the best candidate phage‐displayed p16 binding peptide were evaluated by peptide blocking experiment in competition with p16 monoclonal antibody and fluorescence imaging technique, respectively. Five candidate phage‐displayed peptides were isolated from the phage display selection method. All candidate p16 binding phages show better binding affinity than wild‐type phage in ELISA test, but only three of them can discriminate p16‐overexpressing cancer cell, CaSki, from normal uterine fibroblast cell, HUF, with relative fluorescence intensities from 2.6 to 4.2‐fold greater than those of wild‐type phage. Bioinformatic results indicate that peptide ‘Ser‐His‐Ser‐Leu‐Leu‐Ser‐Ser’ binds to p16 molecule with the best binding score and does not interfere with the common protein functions of p16. Peptide blocking experiment shows that the phage‐displayed peptide ‘Ser‐His‐Ser‐Leu‐Leu‐Ser‐Ser’ can conceal p16 from monoclonal antibody interaction. This phage clone also selectively interacts with the p16 positive cell lines, and thus, it can be applied for p16‐overexpressing cell detection. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
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   

11.
Sun L  Rush J  Ghosh I  Maunus JR  Xu MQ 《BioTechniques》2004,37(3):430-6, 438, 440 passim
Peptide arrays are increasingly used to define antibody epitopes and substrate specificities of protein kinases. Their use is hampered, however, by ineffective and variable binding efficiency of peptides, which often results in low sensitivity and inconsistent results. To overcome these limitations, we have developed a novel method for making arrays of synthetic peptides on various membranes after ligating the peptide substrates to an intein-generated carrier protein. We have conducted screening for optimal carrier proteins by immunoreactivity and direct assessment of binding using a peptide derivatized at a lysine sidechain with fluorescein, CDPEK(fluorescein)DS. Ligation of a synthetic peptide antigen to a carrier protein, HhaI methylase, resulted in an improved retention of peptides and an increased sensitivity of up to 10(4)-fold in immunoassay- and epitope-scanning experiments. Denaturing the ligation products with 2% sodium dodecyl sulfate (SDS) or an organic solvent (20% methanol) prior to arraying did not significantly affect the immunoreactivity of the HhaI methylase-peptide product. Because the carrier protein dominates the binding of ligation products and contains one peptide reactive site, the amount of peptide arrayed onto the membranes can be effectively normalized. This technique was utilized in the alanine scanning of hemagglutinin (HA) antigen using two monoclonal antibodies, resulting in distinguishing the different antigen epitope profiles. Furthermore, we show that this method can be used to characterize the antibodies that recognize phosphorylated peptides. This novel approach allows for synthetic peptides to be uniformly arrayed onto membranes, compatible with a variety of applications.  相似文献   

12.
The growing demand for binding assays to study protein–protein interaction can be addressed by peptide array‐based methods. The SPOT technique is a widespread peptide‐array technology, which is able to distinguish semi‐quantitatively the binding affinities of peptides to defined protein targets within one array. The quality of an assay system used for probing peptide arrays depends on the well‐balanced combination of screening and read‐out methods. The former address the steady‐state of analyte capture, whereas the latter provide the means to detect captured analyte. In all cases, however, false‐positive results can occur when challenging a peptide array with analyte or detecting captured analyte with label conjugates. Little is known about the cross‐reactivity of peptides with the detection agents. Here, we describe at the amino acid level the potential of (i) 5‐(and 6)‐carboxytetramethylrhodamine (5(6)‐TAMRA), (ii) fluoresceinisothiocyanate in form of the peptide‐bound fluorescein‐substituted thiourea derivative (FITC), and (iii) biotin/streptavidin‐POD to cross‐react with individual amino acids in a peptide sequence. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Protein–peptide interactions are a common occurrence and essential for numerous cellular processes, and frequently explored in broad applications within biology, medicine, and proteomics. Therefore, understanding the molecular mechanism(s) of protein–peptide recognition, specificity, and binding interactions will be essential. In this study, we report the first detailed analysis of antibody–peptide interaction characteristics, by combining large‐scale experimental peptide binding data with the structural analysis of eight human recombinant antibodies and numerous peptides, targeting tryptic mammalian and eukaryote proteomes. The results consistently revealed that promiscuous peptide‐binding interactions, that is, both specific and degenerate binding, were exhibited by all antibodies, and the discovery was corroborated by orthogonal data, indicating that this might be a general phenomenon for low‐affinity antibody–peptide interactions. The molecular mechanism for the degenerate peptide‐binding specificity appeared to be executed through the use of 2–3 semi‐conserved anchor residues in the C‐terminal part of the peptides, in analogue to the mechanism utilized by the major histocompatibility complex–peptide complexes. In the long‐term, this knowledge will be instrumental for advancing our fundamental understanding of protein–peptide interactions, as well as for designing, generating, and applying peptide specific antibodies, or peptide‐binding proteins in general, in various biotechnical and medical applications.  相似文献   

14.
We have developed a rapid method for probing the affinity of peptides toward an amphiphilic surface. Hydrophobic polystyrene-divinylbenzene beads of 5.7 +/- 1.5 micron diameter are coated with a monomolecular film of egg lecithin to achieve the equilibrium spreading density of the phospholipid, 6 X 10(-3) molecule/A2. The coated beads are ideally suited for assessing the affinity of peptides for phospholipid surfaces: Large quantities of lipid-coated beads of known surface area can be prepared easily and rapidly. Within the pH range 2.0 to 9.0, the adsorbed phospholipids are relatively resistant to hydrolysis and remain bound indefinitely. Following incubation with peptide ligands, beads can be separated from the reaction mixture by centrifugation. Peptides, such as melittin, which destroy or cause fusion of single bilayer phospholipid vesicles, cannot disrupt lecithin-coated beads in a comparable way, and do not displace lecithin from the surface of beads. After incubating these beads in solutions of peptides and proteins, we have determined the parameters for the binding of several ligands to the phospholipid surface. The binding of many amphiphilic peptides obeys a Langmuir adsorption isotherm, i.e., saturable reversible binding to independent and equivalent sites on the bead. That the binding is a true reversible equilibrium is shown by desorption of the ligand upon dilution. From the isotherm, the surface areas occupied by the ligand molecules were calculated, and were observed to be similar to those observed in monolayers at the air-water interface. In comparing the binding of amphiphilic peptides to that of completely hydrophilic peptides, we observed that only the former bind at levels measurable by our techniques. Thus, this method can serve as a rapid assay for detecting amphiphilicity in peptides of putative amphiphilic character.  相似文献   

15.
Chicken egg yolk immunoglobulin (IgY) is a functional substitute for mammalian IgG for antigen detection. Traditional IgY purification methods involve multi‐step procedures resulting in low purity and recovery of IgY. In this study, we developed a simple IgY purification system using IgY‐specific peptides identified by T7 phage display technology. From disulfide‐constrained random peptide libraries constructed on a T7 phage, we identified three specific binding clones (Y4‐4, Y5‐14, and Y5‐55) through repeated biopanning. The synthetic peptides showed high binding specificity to IgY‐Fc and moderate affinity for IgY‐Fc (Kd: Y4‐4 = 7.3 ± 0.2 μM and Y5‐55 = 4.4 ± 0.1 μM) by surface plasmon resonance analysis. To evaluate the ability to purify IgY, we performed immunoprecipitation and affinity high‐performance liquid chromatography using IgY‐binding peptides; the result indicated that these peptides can be used as affinity ligands for IgY purification. We then used a peptide‐conjugated column to purify IgY from egg yolks pre‐treated using an optimized delipidation technique. Here, we report the construction of a cost‐effective, one‐step IgY purification system, with high purity and recovery. © 2017 The Authors. Journal of Peptide Science published by European Peptide Society and John Wiley & Sons Ltd.  相似文献   

16.
Reversed phase microcolumns have been widely used for peptide pretreatment to desalt and remove interferences before tandem LC–MS in proteomics studies. However, few studies have characterized the effects of experimental parameters as well as column characteristics on the composition of identified peptides. In this study, several parameters including the concentration of ACN in washing buffer, the microcolumn's purification effect, the peptide recovery rate, and the dynamic‐binding capacity were characterized in detail, based upon stable isotope labeling by amino acids in a cell culture quantitative approach. The results showed that peptide losses can be reduced with low ACN concentration in washing buffers resulting in a recovery rate of approximately 82%. Furthermore, the effects of ACN concentration and loading amount on the properties of identified peptides were also evaluated. We found that the dynamic‐binding capacity of the column was approximately 26 μg. With increased loading amounts, more hydrophilic peptides were replaced by hydrophobic peptides.  相似文献   

17.
Many important protein–protein interactions are mediated by the binding of a short peptide stretch in one protein to a large globular segment in another. Recent efforts have provided hundreds of examples of new peptides binding to proteins for which a three-dimensional structure is available (either known experimentally or readily modeled) but where no structure of the protein–peptide complex is known. To address this gap, we present an approach that can accurately predict peptide binding sites on protein surfaces. For peptides known to bind a particular protein, the method predicts binding sites with great accuracy, and the specificity of the approach means that it can also be used to predict whether or not a putative or predicted peptide partner will bind. We used known protein–peptide complexes to derive preferences, in the form of spatial position specific scoring matrices, which describe the binding-site environment in globular proteins for each type of amino acid in bound peptides. We then scan the surface of a putative binding protein for sites for each of the amino acids present in a peptide partner and search for combinations of high-scoring amino acid sites that satisfy constraints deduced from the peptide sequence. The method performed well in a benchmark and largely agreed with experimental data mapping binding sites for several recently discovered interactions mediated by peptides, including RG-rich proteins with SMN domains, Epstein-Barr virus LMP1 with TRADD domains, DBC1 with Sir2, and the Ago hook with Argonaute PIWI domain. The method, and associated statistics, is an excellent tool for predicting and studying binding sites for newly discovered peptides mediating critical events in biology.  相似文献   

18.
In order to facilitate a novel means for coupling proteins to metal oxides, peptides were identified from a dodecamer peptide yeast surface display library that bound a model metal oxide material, the C, A, and R crystalline faces of synthetic sapphire (alpha-Al(2)O(3)). Seven rounds of screening yielded peptides enriched in basic amino acids compared to the naive library. While the C-face had a high background of endogenous yeast cell binding, the A- and R faces displayed clear peptide-mediated cell adhesion. Cell detachment assays showed that cell adhesion strength correlated positively with increasing basicity of expressed peptides. Cell adhesion was also shown to be sensitive to buffer ionic strength as well as incubation with soluble peptide (with half maximal inhibition of cell binding at approximately 5 microM peptide). Next, dodecamer peptides cloned into yeast showed that lysine led to stronger interactions than arginine, and that charge distribution affected adhesion strength. We postulate binding to arise from peptide geometries that permit conformation alignment of the basic amino acids towards the surface so that the charged groups can undergo local electrostatic interactions with the surface oxide. Lastly, peptide K1 (-(GK)(6)) was cloned onto the c-terminus of maltose binding protein (MBP) and the resultant mutant protein showed a half-maximal binding at approximately 10(-7)-10(-6) M, which marked a approximately 500- to 1,000-fold binding improvement to sapphire's A-face as compared with wild-type MBP. Targeting proteins to metal oxide surfaces with peptide tags may provide a facile one-step alternative coupling chemistry for the formation of protein bioassays and biosensors.  相似文献   

19.
Human rho‐associated coiled‐coil forming kinases (ROCKs) ROCK‐I and ROCK‐II have been documented as attractive therapeutic targets for cerebrovascular diseases. Although ROCK‐I and ROCK‐II share a high degree of structural conservation and are both present in classic rho/ROCK signaling pathway, their downstream substrates and pathological functions may be quite different. Selective targeting of the two kinase isoforms with traditional small‐molecule inhibitors is a great challenge due to their surprisingly high homology in kinase domain (~90%) and the full identity in kinase active site (100%). Here, instead of developing small‐molecule drugs to selectively target the adenosine triphosphate (ATP) site of two isoforms, we attempt to design peptide agents to selectively disrupt the homo‐dimerization event of ROCK kinases through their dimerization domains which have a relatively low conservation (~60%). Three helical peptides H1, H2, and H3 are split from the kinase dimerization domain, from which the isolated H2 peptide is found to have the best capability to rebind at the dimerization interface. A simulated annealing (SA) iteration method is used to improve the H2 peptide selectivity between ROCK‐I and ROCK‐II. The method accepts moderate degradation in peptide affinity in order to maximize the affinity difference between peptide binding to the two isoforms. Consequently, hundreds of parallel SA runs yielded six promising peptide candidates with ROCK‐I over ROCK‐II (I over II [IoII]) calculated selectivity and four promising peptide candidates with ROCK‐II over ROCK‐I (II over I [IIoI]) calculated selectivity. Subsequent anisotropy assays confirm that the selectivity values range between 13.2‐fold and 83.9‐fold for IoII peptides, and between 5.8‐fold and 21.2‐fold for IIoI peptides, which are considerably increased relative to wild‐type H2 peptide (2.6‐fold for IoII and 2.0‐fold for IIoI). The molecular origin of the designed peptide selectivity is also analyzed at structural level; it is revealed that the peptide residues can be classified into conserved, non‐conserved, and others, in which the non‐conserved residues play a crucial role in defining peptide selectivity, while conserved residues confer stability to kinase‐peptide binding.  相似文献   

20.
Alzheimer disease is a neurodegenerative disease affecting an increasing number of patients worldwide. Current therapeutic strategies are directed to molecules capable to block the aggregation of the β‐amyloid(1‐42) (Aβ) peptide and its shorter naturally occurring peptide fragments into toxic oligomers and amyloid fibrils. Aβ‐specific antibodies have been recently developed as powerful antiaggregation tools. The identification and functional characterization of the epitope structures of Aβ antibodies contributes to the elucidation of their mechanism of action in the human organism. In previous studies, the Aβ(4‐10) peptide has been identified as an epitope for the polyclonal anti‐Aβ(1‐42) antibody that has been shown capable to reduce amyloid deposition in a transgenic Alzheimer disease mouse model. To determine the functional significance of the amino acid residues involved in binding to the antibody, we report here the effects of alanine single‐site mutations within the Aβ‐epitope sequence on the antigen‐antibody interaction. Specific identification of the essential affinity preserving mutant peptides was obtained by exposing a Sepharose‐immobilized antibody column to an equimolar mixture of mutant peptides, followed by analysis of bound peptides using high‐resolution MALDI‐Fourier transform‐Ion Cyclotron Resonance mass spectrometry. For the polyclonal antibody, affinity was preserved in the H6A, D7A, S8A, and G9A mutants but was lost in the F4, R5, and Y10 mutants, indicating these residues as essential amino acids for binding. Enzyme‐linked immunosorbent assays confirmed the binding differences of the mutant peptides to the polyclonal antibody. In contrast, the mass spectrometric analysis of the mutant Aβ(4‐10) peptides upon affinity binding to a monoclonal anti‐Aβ(1‐17) antibody showed complete loss of binding by Ala‐site mutation of any residue of the Aβ(4‐10) epitope. Surface plasmon resonance affinity determination of wild‐type Aβ(1‐17) to the monoclonal Aβ antibody provided a binding constant KD in the low nanomolar range. These results provide valuable information in the elucidation of the binding mechanism and the development of Aβ‐specific antibodies with improved therapeutic efficacy.  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