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.
Most antigenic sites of proteins, known as discontinuous epitopes, are made up of residues on different loops that are brought together by the folding of the polypeptide chain. The individual loops are sometimes able, on their own, to bind to the antibody and they are then known as continuous epitopes. The binding sites of antibodies, known as paratopes, are built up from residues on six hypervariable loops known as complementarity determining regions (CDRs). Peptides corresponding to individual CDR loops are often able to bind the antigen and such peptides may be viewed as continuous paratopes. Using random combinatorial peptide libraries, it is possible to obtain peptides that bind to an antiprotein antibody without showing any sequence similarity with any part of the protein. Such epitope mimics are called mimotopes provided they are able also to elicit antibodies that react with the original antigen. The binding activity of mimotopes may partly be due to the phenomenon of hydropathic complementarity between epitope and paratope peptides. Although these concepts are vague in their structural connotation, they are useful for describing the immunological activity of linear peptides.  相似文献   

2.
A subset of the neutralizing anti-HIV antibodies recognize epitopes on the envelope protein gp120 of the human immunodeficiency virus. These epitopes are exposed during conformational changes when gp120 binds to its primary receptor CD4. Based on chemical modification of lysine and arginine residues followed by mass spectrometric analysis, we determined the epitope on gp120 recognized by the human monoclonal antibody 559/64-D, which was previously found to be specific for the CD4 binding domain. Twenty-four lysine and arginine residues in recombinant full-length glycosylated gp120 were characterized; the relative reactivities of two lysine residues and five arginine residues were affected by the binding of 559/64-D. The data show that the epitope is discontinuous and is located in the proximity of the CD4-binding site. Additionally, the reactivities of a residue that is located in the secondary receptor binding region and several residues distant from the CD4 binding site were also altered by Ab binding. These data suggest that binding of 559/64-D induced conformational changes which result in altered surface exposure of specific amino acids distant from the CD4-binding site. Consequently, binding of 559/64-D to gp120 affects not only the CD4-binding site, which is recognized as the epitope, but appears to have a global effect on surface exposed residues of the full-length glycosylated gp120.  相似文献   

3.
Affinity selection of peptides displayed on phage particles was used as the basis for mapping molecular contacts between small molecule ligands and their protein targets. Analysis of the crystal structures of complexes between proteins and small molecule ligands revealed that virtually all ligands of molecular weight 300 Da or greater have a continuous binding epitope of 5 residues or more. This observation led to the development of a technique for binding site identification which involves statistical analysis of an affinity-selected set of peptides obtained by screening of libraries of random, phage-displayed peptides against small molecules attached to solid surfaces. A random sample of the selected peptides is sequenced and used as input for a similarity scanning program which calculates cumulative similarity scores along the length of the putative receptor. Regions of the protein sequence exhibiting the highest similarity with the selected peptides proved to have a high probability of being involved in ligand binding. This technique has been employed successfully to map the contact residues in multiple known targets of the anticancer drugs paclitaxel (Taxol), docetaxel (Taxotere) and 2-methoxyestradiol and the glycosaminoglycan hyaluronan, and to identify a novel paclitaxel receptor [1]. These data corroborate the observation that the binding properties of peptides displayed on the surface of phage particles can mimic the binding properties of peptides in naturally occurring proteins. It follows directly that structural context is relatively unimportant for determining the binding properties of these disordered peptides. This technique represents a novel, rapid, high resolution method for identifying potential ligand binding sites in the absence of three-dimensional information and has the potential to greatly enhance the speed of development of novel small molecule pharmaceuticals.  相似文献   

4.
An important factor that defines the toxicity of elements such as cadmium(II), mercury(II), and lead(II) with biological macromolecules is metal ion exchange dynamics. Intriguingly, little is known about the fundamental rates and mechanisms of metal ion exchange into proteins, especially helical bundles. Herein, we investigate the exchange kinetics of Cd(II) using de novo designed three-stranded coiled-coil peptides that contain metal complexing cysteine thiolates as a model for the incorporation of this ion into trimeric, parallel coiled coils. Peptides were designed containing both a single Cd(II) binding site, GrandL12AL16C [Grand = AcG-(LKALEEK)(5)-GNH(2)], GrandL26AL30C, and GrandL26AE28QL30C, as well as GrandL12AL16CL26AL30C with two Cd(II) binding sites. The binding of Cd(II) to any of these sites is of high affinity (K(A) > 3 × 10(7) M(-1)). Using (113)Cd NMR spectroscopy, Cd(II) binding to these designed peptides was monitored. While the Cd(II) binding is in extreme slow exchange regime without showing any chemical shift changes, incremental line broadening for the bound (113)Cd(II) signal is observed when excess (113)Cd(II) is titrated into the peptides. Most dramatically, for one site, L26AL30C, all (113)Cd(II) NMR signals disappear once a 1.7:1 ratio of Cd(II)/(peptide)(3) is reached. The observed processes are not compatible with a simple "free-bound" two-site exchange kinetics at any time regime. The experimental results can, however, be simulated in detail with a multisite binding model, which features additional Cd(II) binding site(s) which, once occupied, perturb the primary binding site. This model is expanded into differential equations for five-site NMR chemical exchange. The numerical integration of these equations exhibits progressive loss of the primary site NMR signal without a chemical shift change and with limited line broadening, in good agreement with the observed experimental data. The mathematical model is interpreted in molecular terms as representing binding of excess Cd(II) to surface Glu residues located at the helical interfaces. In the absence of Cd(II), the Glu residues stabilize the three-helical structure though salt bridge interactions with surface Lys residues. We hypothesize that Cd(II) interferes with these surface ion pairs, destabilizing the helical structure, and perturbing the primary Cd(II) binding site. This hypothesis is supported by the observation that the Cd(II)-excess line broadening is attenuated in GrandL26AE28QL30C, where a surface Glu(28), close to the metal binding site, was changed to Gln. The external binding site may function as an entry pathway for Cd(II) to find its internal binding site following a molecular rearrangement which may serve as a basis for our understanding of metal complexation, transport, and exchange in complex native systems containing α-helical bundles.  相似文献   

5.
6.
An increasing number of peptides with specific binding affinity to inorganic materials are being isolated using combinatorial peptide libraries without prior knowledge about the interaction between peptides and target materials. The lack of understanding of the mechanism and the contribution of constituent amino acids to the peptides' inorganic-binding ability poses an obstacle to optimizing and tuning of the binding affinity of peptides to inorganic materials and thus hinders the practical application of these peptides. Using the phage surface display technique, we previously identified a disulfide-bond-constrained peptide (-CHKKPSKSC-, STB1) cognitive of TiO2. In the present study, the interaction of STB1 with TiO2 was probed using a series of point mutants of STB1 displayed on phage surfaces. Their binding affinity was measured using a quartz crystal microbalance with energy dissipation measurement and compared on the basis of the delta f or delta D values. The three K residues of STB1 were found to be essential and sufficient for phage particle binding to TiO2. One mutant with five K residues showed not stronger but weaker binding affinity than STB1 due to its conformational restriction, as illustrated by molecular dynamics simulation, to align five K residues in a way conducive to their simultaneous interaction with the TiO2 surface. The contextual influence of noncharged residues on STB1's binding affinity was also investigated. Our results may provide insight into the electrostatic interaction between peptides and inorganic surfaces.  相似文献   

7.
Noncovalently bound complexes between basic sites of peptides/proteins and sulfonates are studied using Matrix Assisted Laser Desorption/Ionization (MALDI) Mass Spectrometry. Reactive sulfonate dyes such as Cibacron Blue F3G-A are known to bind to protonated amino groups on the exterior of a protein. In this work, we examine a wide range of other sulfonates with distinctly simpler structure and more predictable reactivity. Naphthalene-sulfonic acid derivatives were found to bind to arginine only, as opposed to expected binding to all basic sites (Arg, Lys and His). Detailed control experiments were designed to unambigously confirm this selectivity and to rule out nonspecific adduct formation in the gas phase. The data show that the number of complex adducts found equals the number of accessible arginine sites on the surface of folded peptides and proteins, plus the N-terminus. Lys and His are not complexed nor are buried residues with hindered access. MALDI-MS can therefore provide fast information related to the exposed surface of these biomolecules. Additional titration experiments with 1-anilino-naphthalene-8-sulfonic acid (ANS) revealed that this fluorescent dye, which was often hypothesized to bind to so-called molten globule states of proteins, behaved exactly like all other naphthalene-sulfonic acids. ANS binding thus occurs largely through the sulfonate group.  相似文献   

8.
应用氨基酸描述子VHSE(Principal component score vector of hydrophobic, steric, and electronic properties)对613个抗原9肽进行结构表征, 在此基础上, 采用支持向量机结合逐步回归变量筛选方法, 成功建立了抗原肽抗原处理相关转运蛋白(Transporter associated with antigen processing, TAP)亲和活性预测模型, 最优线性支持向量机模型的R2, Q2和R2ext分别为0.7386, 0.7270和0.6057. 模型结果分析表明, 影响TAP亲和活性的首要因素是电性, 其次是立体和疏水性质; 底物9肽的P1(N端)及P2, P7和P9(C端)位氨基酸物化性质对TAP亲和活性有重要影响, 而P3, P4, P5和P6位对模型贡献相对较小, P8位则与活性无关. 依据最优模型对模拟点突变9肽的TAP亲和活性的预测结果, 并结合变量载荷分析, 对TAP底物选择特异性进行了分析和总结.  相似文献   

9.
Recent reports show that there is a large increase in heme in the temporal brain of Alzheimer's disease (AD) patients, as heme, biosynthesized in brain cells, binds to amyloid β (Aβ), forming heme-Aβ complexes. This leads to the development of symptoms that are characteristic pathological features of AD, e.g., abnormal iron homeostasis, decay of iron regulatory proteins, dysfunction in mitochondrial complex IV, oxidative stress, etc. However, the active site resulting from heme binding to Aβ is not well characterized. For example, the coordinating residue, relevant second-sphere residues, and spin state of the Fe center are not known. In this study we have used wild-type and mutated Aβ peptides and investigated their interaction with naturally occurring heme. Our results show that, out of several possible binding sites, His(13) and His(14) residues can both bind heme under physiological conditions, resulting in an axial high-spin active site with a trans axial water-derived ligand. Peroxidase assays of these heme-peptide complexes along with pH perturbations indicate that Arg(5) is a key second-sphere residue that H-bonds to the trans axial ligand and is responsible for the peroxidase activity of the heme-Aβ complexes. The His(13) and Arg(5) residues identified in this study are both absent in rodents, which do not show AD, implicating the significance of these residues as well as heme in the pathology of AD.  相似文献   

10.
Electrostatic interactions play an important role in the formation of noncovalent complexes. Our previous work has highlighted the role of certain amino acid residues, such as arginine, glutamate, aspartate, and phosphorylated/sulfated residues, in the formation of salt bridges resulting in noncovalent complexes between peptides. Tandem mass spectrometry (MS) studies of these complexes using collision-induced dissociation (CID) have provided information on their relative stability. However, product-ion spectra produced by CID have been unable to assign specifically the site of interaction for the complex. In this work, tandem MS experiments were conducted on noncovalent complexes using both electron capture dissociation (ECD) and electron-transfer dissociation (ETD). The resulting spectra were dominated by intramolecular fragments of the complex with the electrostatic interaction site intact. Based upon these data, we were able to assign the binding site for the peptides forming the noncovalent complex.  相似文献   

11.
Protein engineering has as its goals the design and construction of new peptides and proteins with novel binding and catalytic properties. In one approach to protein engineering, new active sites have been introduced into naturally occurring proteins either by site-directed mutagenesis or by chemical modification. Providing that important changes in the tertiary structures do not result from such alterations, at least a portion of the binding site of the original protein should be available for the formation of complexes between the altered enzyme and its substrates. Many examples of active-site mutations have been described, including the generation by us of a cysteine mutant of alkaline phosphatase. A fundamental limitation of the site-directed mutagenesis methodology is that replacements of residues are restricted to the twenty naturally occurring amino acids. The alternative, chemical modification, is difficult to carry out for the specific replacement of one amino acid by another. However, we have shown that through such modification coenzyme analogues can be introduced covalently into appropriate positions in proteins, allowing us to produce semisynthetic enzymes with catalytic activities radically altered from those of their precursor proteins. In another approach to protein engineering efforts have focused on the construction of systems where, as a first approximation, folding can be neglected and the preparation of secondary structural units is the target. Examples of the successful design of biologically active peptides and proteins along such lines, taken from our own work, include molecules mimicking apolipoproteins, toxins, and many hormones. In recent studies we have progressed to the stage where we are starting to combine the two general approaches to protein engineering we have described and are able to construct small enzymes like ribonuclease T1 and its structural analogues.  相似文献   

12.
Glutaredoxins (Grxs), the oxidoreductase proteins, are involved in several cellular processes, including maintenance of cellular redox potential and iron-sulfur homeostasis. The analysis of 503 amino acid sequences from 167 cyanobacterial species led to the identification of four classes of cyanobacterial Grxs, i.e., class I, II, V, and VI Grxs. Class III and IV Grxs were absent in cyanobacteria. Class I and II Grxs are single module oxidoreductase while class V and VI Grxs are multimodular proteins having additional modules at their C-terminal and N-terminal end, respectively. Furthermore, class VI Grxs were exclusively present in marine cyanobacteria. We also report the identification of class VI Grxs with two novel active site motif compositions. Detailed phylogenetic analysis of all four classes of Grxs revealed the presence of several subgroups within each class of Grx having variable dithiol and/or monothiol catalytic active site motif and putative glutathione binding sites. However, class II Grxs possess CGFS-type highly conserved monothiol catalytic active site motif. Sequence analysis confirmed the highly diverse nature of Grx proteins in terms of their amino acid composition; though, sequence diversity does not affect the overall 3D structure of cyanobacterial Grxs. The active site residues and putative GSH binding residues are uncharged amino acids which are present on the surface of the protein. Additionally, the presence of hydrophilic residues at the surface of Grxs confirms their solubility. Protein-ligand interaction analysis identified novel glutathione binding sites on Grxs. Regulation of Grxs encoding genes expression by light quality and quantity as well as salinity suggests their role in determining the fitness of organisms under abiotic factors.  相似文献   

13.
The structure of peptide p6.7, a mimotope of the nicotinic receptor ligand site that binds alpha-bungarotoxin and neutralizes its toxicity, was compared to that of the acetylcholine binding protein. The central loop of p6.7, when complexed with alpha-bungarotoxin, fits the structure of the acetylcholine binding protein (AChBP) ligand site, whereas peptide terminal residues seem to be less involved in toxin binding. The minimal binding sequence of p6.7 was confirmed experimentally by synthesis of progressively deleted peptides. Affinity maturation was then achieved by random addition of residues flanking the minimal binding sequence and by selection of new alpha-bungarotoxin binding peptides on the basis of their dissociation kinetic rate. The tetra-branched forms of the resulting high-affinity peptides were effective as antidotes in vivo at a significantly lower dose than the tetra-branched lead peptide.  相似文献   

14.
Continuum solvation methods are frequently used to increase the efficiency of computational methods to estimate free energies. In this paper, we have evaluated how well such methods estimate the nonpolar solvation free-energy change when a ligand binds to a protein. Three different continuum methods at various levels of approximation were considered, viz., the polarized continuum model (PCM), a method based on cavity and dispersion terms (CD), and a method based on a linear relation to the solvent-accessible surface area (SASA). Formally rigorous double-decoupling thermodynamic integration was used as a benchmark for the continuum methods. We have studied four protein-ligand complexes with binding sites of varying solvent exposure, namely the binding of phenol to ferritin, a biotin analogue to avidin, 2-aminobenzimidazole to trypsin, and a substituted galactoside to galectin-3. For ferritin and avidin, which have relatively hidden binding sites, rather accurate nonpolar solvation free energies could be obtained with the continuum methods if the binding site is prohibited to be filled by continuum water in the unbound state, even though the simulations and experiments show that the ligand replaces several water molecules upon binding. For the more solvent exposed binding sites of trypsin and galectin-3, no accurate continuum estimates could be obtained, even if the binding site was allowed or prohibited to be filled by continuum water. This shows that continuum methods fail to give accurate free energies on a wide range of systems with varying solvent exposure because they lack a microscopic picture of binding-site hydration as well as information about the entropy of water molecules that are in the binding site before the ligand binds. Consequently, binding affinity estimates based upon continuum solvation methods will give absolute binding energies that may differ by up to 200 kJ/mol depending on the method used. Moreover, even relative energies between ligands with the same scaffold may differ by up to 75 kJ/mol. We have tried to improve the continuum solvation methods by adding information about the solvent exposure of the binding site or the hydration of the binding site, and the results are promising at least for this small set of complexes.  相似文献   

15.
Protein-RNA interactions perform diverse functions within the cell. Understanding the recognition mechanism of protein-RNA complexes is a challenging task in molecular and computational biology. In this work, we have developed an energy based approach for identifying the binding sites and important residues for binding in protein-RNA complexes. The new approach considers the repulsive interactions as well as the effect of distance between the atoms in protein and RNA in terms of interaction energy, which are not considered in traditional distance based methods to identify the binding sites. We found that the positively charged, polar and aromatic residues are important for binding. These residues influence to form electrostatic, hydrogen bonding and stacking interactions. Our observation has been verified with the experimental binding specificity of protein-RNA complexes and found good agreement with experiments. Further, the propensities of residues/nucleotides in the binding sites of proteins/RNA and their atomic contributions have been derived. Based on these results we have proposed a novel mechanism for the recognition of protein-RNA complexes: the charged and polar residues in proteins initiate recognition with RNA by making electrostatic and hydrogen bonding interactions between them; the aromatic side chains tend to form aromatic-aromatic interactions and the hydrophobic residues aid to stabilize the complex.  相似文献   

16.
Calreticulin (CRT) is localized to and has functions in multiple cellular compartments, including the cell surface, the endoplasmic reticulum, and the extracellular matrix. Mutagenesis studies have identified several residues on a concave β-sheet surface of CRT critical for CRT binding to carbohydrate and other proteins/peptides. How the mutations of these key residues in CRT affect the conformation and dynamics of CRT, further influencing CRT binding to carbohydrates and other proteins to signal the important biological activities remain unknown. In this study, we investigated the effect of three key point mutations (C105A, C137A and W319A) on CRT conformation and dynamics via atomistic molecular dynamics simulations. Results show that these three key residues mutations induced the changes of CRT local backbone flexibility and secondary structure of CRT N-domain, which could further affect CRT’s binding activity. C137A mutation led to dramatic decrease of the overall size of CRT due to the P-domain fold back to the globular domain and formed new inter-domain contacts, which can cause blockage of CRT’s binding with other large substrates. Furthermore, for CRT concave β-strand surface patch containing lectin binding site, CRT C105A, C137A and W319A point mutation resulted in the changes in solvent accessible surface area, key residues’ side chain atom positions and dynamical correlated motions between residues. All these changes could directly affect CRT binding behavior. Results of this study provide molecular and structural insights into understanding the role of key residues of CRT in its binding behavior.  相似文献   

17.
Arginine residues undergo several kinds of post-translational modifications (PTMs). These PTMs are associated with several inflammatory diseases, such as rheumatoid arthritis, atherosclerosis, and diabetes. Mass spectrometric studies of arginine modified proteins and peptides are very important, not only to identify the reactive arginine residues but also to understand the tandem mass spectrometry behavior of these peptides for assigning the sequences unambiguously. Herein, we utilize tandem mass spectrometry to report the performance of two widely used arginine labeling reagents, 1,2-cyclohexanedione (CHD) and phenylglyoxal (PG) with several arginine containing peptides and proteins. Time course labeling studies were performed to demonstrate the selectivity of the reagents in proteins or protein digests. Structural studies on the proteins were also explored to better understand the reaction sites and position of arginine residues. We found CHD showed better labeling efficiencies compared to phenylglyoxal. Reactive arginine profiling on a purified albumin protein clearly pointed out the cellular glycation modification site for this protein with high confidence.  相似文献   

18.
Nephrin is a type-1 transmembrane glycoprotein and the first identified principal component of the glomerular filtration barrier. Ten potential asparagine (N)-linked glycosylation sites have been predicted within the ectodomain of nephrin. However, it is not known which of these potential sites are indeed glycosylated and what type of glycans are involved. In this work, we have identified the terminal sugar residues on the ectodomain of human nephrin and utilized a straightforward and reliable mass spectrometry-based approach to selectively identify which of the ten predicted sites are glycosylated. Purified recombinant nephrin was subjected to peptide-N-glycosidase F (PNGase F) to enzymatically remove all the N-linked glycans. Since PNGase F is an amidase, the asparagine residues from which the glycans have been removed are deaminated to aspartic acid residues, resulting in an increase in the peptide mass with 1 mass unit. Following trypsin digestion, deglycosylated tryptic peptides were selectively identified by MALDI-TOF MS and their sequence was confirmed by tandem TOF/TOF. The 1 Da increase in peptide mass for each asparagine-to-aspartic acid conversion, along with preferential cleavage of the amide bond carboxyl-terminal to aspartic acid residues in peptides where the charge is immobilized by an arginine residue, was used as a diagnostic signature to identify the glycosylated peptides. Thus, nine of ten potential glycosylation sites in nephrin were experimentally proven to be modified by N-linked glycosylation.  相似文献   

19.
Background: Zinc binding proteins make up a significant proportion of the proteomes of most organisms and, within those proteins, zinc performs rôles in catalysis and structure stabilisation. Identifying the ability to bind zinc in a novel protein can offer insights into its functions and the mechanism by which it carries out those functions. Computational means of doing so are faster than spectroscopic means, allowing for searching at much greater speeds and scales, and thereby guiding complimentary experimental approaches. Typically, computational models of zinc binding predict zinc binding for individual residues rather than as a single binding site, and typically do not distinguish between different classes of binding site—missing crucial properties indicative of zinc binding. Methods: Previously, we created ZincBindDB, a continuously updated database of known zinc binding sites, categorised by family (the set of liganding residues). Here, we use this dataset to create ZincBindPredict, a set of machine learning methods to predict the most common zinc binding site families for both structure and sequence. Results: The models all achieve an MCC ≥ 0.88, recall ≥ 0.93 and precision ≥ 0.91 for the structural models (mean MCC = 0.97), while the sequence models have MCC ≥ 0.64, recall ≥ 0.80 and precision ≥ 0.83 (mean MCC = 0.87), with the models for binding sites containing four liganding residues performing much better than this. Conclusions: The predictors outperform competing zinc binding site predictors and are available online via a web interface and a GraphQL API.  相似文献   

20.
《Chemistry & biology》1996,3(3):207-215
Background: Recent work has indicated that dimerization is important in the mode of action of the vancomycin group of glycopeptide antibiotics. NMR studies have shown that one member of this group, ristocetin A1 forms an asymmetric dimer with two physically different binding sites for cell wall peptides. Ligand binding by ristocetin A and dimerization are slightly anti-cooperative. In contrast, for the other glycopeptide antibiotics of the vancomycin group that have been examined so far, binding of cell wall peptides and dimerization are cooperative.Results: Here we show that the two halves of the asymmetric homodimer formed by ristocetin A have different affinities for ligand binding. One of these sites is preferentially filled before the other, and binding to this site is cooperative with dimerization. Ligand binding to the other, less favored half of the dimer, is anti-cooperative with dimerization.Conclusions: In dinner complexes, anti-cooperativity of dimerization upon ligand binding can be a result of asymmetry, in which two binding sites have different affinities for ligands. Such a system, in which one binding site is filled preferentially, may be a mechanism by which the cooperativity between ligand binding and dimerization is fine tuned and may thus have relevance to the control of signal transduction in biological systems.  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