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 0 毫秒
1.
The concept of antibody specificity is analyzed and shown to reside in the ability of an antibody to discriminate between two antigens. Initially, antibody specificity was attributed to sequence differences in complementarity determining regions (CDRs), but as increasing numbers of crystallographic antibody‐antigen complexes were elucidated, specificity was analyzed in terms of six antigen‐binding regions (ABRs) that only roughly correspond to CDRs. It was found that each ABR differs significantly in its amino acid composition and tends to bind different types of amino acids at the surface of proteins. In spite of these differences, the combined preference of the six ABRs does not allow epitopes to be distinguished from the rest of the protein surface. These findings explain the poor success of past and newly proposed methods for predicting protein epitopes. Antibody polyspecificity refers to the ability of one antibody to bind a large variety of epitopes in different antigens, and this property explains how the immune system develops an antibody repertoire that is able to recognize every antigen the system is likely to encounter. Antibody heterospecificity arises when an antibody reacts better with another antigen than with the one used to raise the antibody. As a result, an antibody may sometimes appear to have been elicited by an antigen with which it is unable to react. The implications of antibody polyspecificity and heterospecificity in vaccine development are pointed out. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

2.
Allosteric HIV‐1 integrase (IN) inhibitors (ALLINIs) bind at the dimer interface of the IN catalytic core domain (CCD), and potently inhibit HIV‐1 by promoting aberrant, higher‐order IN multimerization. Little is known about the structural organization of the inhibitor‐induced IN multimers and important questions regarding how ALLINIs promote aberrant IN multimerization remain to be answered. On the basis of physical chemistry principles and from our analysis of experimental information, we propose that inhibitor‐induced multimerization is mediated by ALLINIs directly promoting inter‐subunit interactions between the CCD dimer and a C‐terminal domain (CTD) of another IN dimer. Guided by this hypothesis, we have built atomic models of inter‐subunit interfaces in IN multimers by incorporating information from hydrogen‐deuterium exchange (HDX) measurements to drive protein‐protein docking. We have also developed a novel free energy simulation method to estimate the effects of ALLINI binding on the association of the CCD and CTD. Using this structural and thermodynamic modeling approach, we show that multimer inter‐subunit interface models can account for several experimental observations about ALLINI‐induced multimerization, including large differences in the potencies of various ALLINIs, the mechanisms of resistance mutations, and the crucial role of solvent exposed R‐groups in the high potency of certain ALLINIs. Our study predicts that CTD residues Tyr226, Trp235 and Lys266 are involved in the aberrant multimer interfaces. The key finding of the study is that it suggests the possibility of ALLINIs facilitating inter‐subunit interactions between an external CTD and the CCD‐CCD dimer interface.  相似文献   

3.
The selection of suitable mammalian cell lines with high specific productivities is a crucial aspect of large‐scale recombinant protein production. This study utilizes a metabolomics approach to elucidate the key characteristics of Chinese hamster ovary (CHO) cells with high monoclonal antibody productivities (qmAb). Liquid chromatography‐mass spectrometry (LC‐MS)‐based intracellular metabolite profiles of eight single cell clones with high and low qmAb were obtained at the mid‐exponential phase during shake flask batch cultures. Orthogonal projection to latent structures discriminant analysis (OPLS‐DA) subsequently revealed key differences between the high and low qmAb clones, as indicated by the variable importance for projection (VIP) scores. The mass peaks were further examined for their potential association with qmAb across all clones using Pearson's correlation analysis. Lastly, the identities of metabolites with high VIP and correlation scores were confirmed by comparison with standards through LC‐MS‐MS. A total of seven metabolites were identified—NADH, FAD, reduced and oxidized glutathione, and three activated sugar precursors. These metabolites are involved in key cellular pathways of citric acid cycle, oxidative phosphorylation, glutathione metabolism, and protein glycosylation. To our knowledge, this is the first study to identify metabolites that are associated closely with qmAb. The results suggest that the high producers had elevated levels of specific metabolites to better regulate their redox status. This is likely to facilitate the generation of energy and activated sugar precursors to meet the demands of producing more glycosylated recombinant monoclonal antibodies. Biotechnol. Bioeng. 2012; 109: 3103–3111. © 2012 Wiley Periodicals, Inc.  相似文献   

4.
Given the increasing exploitation of antibodies in different contexts such as molecular diagnostics and therapeutics, it would be beneficial to unravel the atomistic level properties of antibody‐antigen complexes with the help of computational modeling. Thus, here we have studied the feasibility of computational tools to gather atomic scale information regarding the antibody‐antigen complexes solely starting from an amino acid sequence. First, we constructed a homology model for the anti‐testosterone binding antibody based on the knowledge based classification of complementary determining regions (CDRs) and implicit solvent molecular dynamics simulations. To further examine whether the generated homology model is suitable for studying antibody‐antigen interactions, docking calculations were carried out followed by binding free‐energy simulations. Our results indicate that with the antibody modeling approach presented here it is possible to construct accurate homology models for antibodies which correctly describes the antibody‐antigen interactions, and produces absolute binding free‐energies that are comparable with experimental values. In addition, our simulations suggest that the conformations of complementary determining regions (CDRs) may considerably change from the X‐ray configuration upon solvation. In conclusion, here we have introduced an antibody modeling workflow that can be used in studying the interactions between antibody and antigen solely based on an amino acid sequence, which in turn provides novel opportunities to tune the properties of antibodies in different applications. Proteins 2017; 85:322–331. © 2016 Wiley Periodicals, Inc.  相似文献   

5.
The solvation of the antibody–antigen Fv D1.3–lysozyme complex is investigated through a study of the conservation of water molecules in crystal structures of the wild-type Fv fragment of antibody D1.3, 5 free lysozyme, the wild-type Fv D1.3–lysozyme complex, 5 Fv D1.3 mutants complexed with lysozyme and the crystal structure of an idiotope (Fv D1.3)-abti-idiotope (Fv E5.2) complex. In all, there are 99 water molecules common to the wild-type and mutant antibody–lysozyme complexes. The antibody–lysozyme interface includes 25 well-ordered solvent molecules, conserved among the wild-type and mutant Fv D1.3–lysozyme complexes, which are bound directly or through other water molecules to both antibody and antigen. In addition to contributing hydrogen bonds to the antibody–antigen interaction the solvent molecules fill many interface cavities. Comparison with x-ray crystal structures of free Fv D1.3 and free lysozyme shows that 20 of these conserved interface waters in the complex were bound to one of the free proteins. Uo to 23 additional water molecules are also found in the antibody–antigen interface, however these waters do no bridge antibody and antigen and their temperature factors are much higher than those of the 25 well-ordered waters. Fifteen water molecules are displaced to form the complex, some of which are substituted by hydrophilic protein atoms, and 5 water molecules are added at the antibody–antigen interface with the formation of the complex. While the current crystal models of the D1.3–lysozyme complex do not demonstrate the increase in bound waters found in a physico-chemical study of the interaction at decreased water activities, the 25 well-ordered interface water contribute a net gain of 10 hydrogen bonds to complex stability.  相似文献   

6.
An antibody fragment that recognizes the tertiary structure of a target protein with high affinity can be utilized as a crystallization chaperone. Difficulties in establishing conformation‐specific antibodies, however, limit the applicability of antibody fragment‐assisted crystallization. Here, we attempted to establish an alternative method to promote the crystallization of target proteins using an already established anti‐tag antibody. The monoclonal antibody NZ‐1 recognizes the PA tag with an extremely high affinity. It was also established that the PA tag is accommodated in the antigen‐binding pocket in a bent conformation, compatible with an insertion into loop regions on the target. We, therefore, explored the application of NZ‐1 Fab as a crystallization chaperone that complexes with a target protein displaying a PA tag. Specifically, we inserted the PA tag into the β‐hairpins of the PDZ tandem fragment of a bacterial Site‐2 protease. We crystallized the PA‐inserted PDZ tandem mutants with the NZ‐1 Fab and solved the co‐crystal structure to analyze their interaction modes. Although the initial insertion designs produced only moderate‐resolution structures, eliminating the solvent‐accessible space between the NZ‐1 Fab and target PDZ tandem improved the diffraction qualities remarkably. Our results demonstrate that the NZ‐1‐PA system efficiently promotes crystallization of the target protein. The present work also suggests that β‐hairpins are suitable sites for the PA insertion because the PA tag contains a Pro‐Gly sequence with a propensity for a β‐turn conformation.  相似文献   

7.
The modulation of biological signal transduction pathways by masking phosphorylated amino acid residues represents a viable route toward pharmacologic protein regulation. Binding of phosphorylated amino acid residues has been achieved with synthetic metal‐chelate receptors. The affinity and selectivity of such receptors can be enhanced if combined with a second binding site. We demonstrate this principle with a series of synthetic ditopic metal‐chelate receptors, which were synthesized and investigated for their binding affinity to phosphorylated short peptides under conditions of physiological pH. The compounds showing highest affinity were subsequently used to inhibit the interaction of the human STAT1 protein to a peptide derived from the interferon‐γ receptor, and between the checkpoint kinase Chk2 and its preferred binding motif. Two of the investigated ditopic synthetic receptors show a significant increase in inhibition activity. The results show that regulation of protein function by binding to phosphorylated amino acids is possible. The introduction of additional binding sites into the synthetic receptors increases their affinity, but the flexibility of the structures investigated so far prohibited stringent amino acid sequence selectivity in peptide binding. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Sudip Kundu 《Proteins》2018,86(8):827-832
Do coding and regulatory segments of a gene co‐evolve with each‐other? Seeking answers to this question, here we analyze the case of Escherichia coli ribosomal protein S15, that represses its own translation by specifically binding its messenger RNA (rpsO mRNA) and stabilizing a pseudoknot structure at the upstream untranslated region, thus trapping the ribosome into an incomplete translation initiation complex. In the absence of S15, ribosomal protein S1 recognizes rpsO and promotes translation by melting this very pseudoknot. We employ a robust statistical method to detect signatures of positive epistasis between residue site pairs and find that biophysical constraints of translational regulation (S15‐rpsO and S1‐rpsO recognition, S15‐mediated rpsO structural rearrangement, and S1‐mediated melting) are strong predictors of positive epistasis. Transforming the epistatic pairs into a network, we find that signatures of two different, but interconnected regulatory cascades are imprinted in the sequence‐space and can be captured in terms of two dense network modules that are sparsely connected to each other. This network topology further reflects a general principle of how functionally coupled components of biological networks are interconnected. These results depict a model case, where translational regulation drives characteristic residue‐level epistasis—not only between a protein and its own mRNA but also between a protein and the mRNA of an entirely different protein.  相似文献   

9.
Computational prediction of RNA‐binding residues is helpful in uncovering the mechanisms underlying protein‐RNA interactions. Traditional algorithms individually applied feature‐ or template‐based prediction strategy to recognize these crucial residues, which could restrict their predictive power. To improve RNA‐binding residue prediction, herein we propose the first integrative algorithm termed RBRDetector (RNA‐Binding Residue Detector) by combining these two strategies. We developed a feature‐based approach that is an ensemble learning predictor comprising multiple structure‐based classifiers, in which well‐defined evolutionary and structural features in conjunction with sequential or structural microenvironment were used as the inputs of support vector machines. Meanwhile, we constructed a template‐based predictor to recognize the putative RNA‐binding regions by structurally aligning the query protein to the RNA‐binding proteins with known structures. The final RBRDetector algorithm is an ingenious fusion of our feature‐ and template‐based approaches based on a piecewise function. By validating our predictors with diverse types of structural data, including bound and unbound structures, native and simulated structures, and protein structures binding to different RNA functional groups, we consistently demonstrated that RBRDetector not only had clear advantages over its component methods, but also significantly outperformed the current state‐of‐the‐art algorithms. Nevertheless, the major limitation of our algorithm is that it performed relatively well on DNA‐binding proteins and thus incorrectly predicted the DNA‐binding regions as RNA‐binding interfaces. Finally, we implemented the RBRDetector algorithm as a user‐friendly web server, which is freely accessible at http://ibi.hzau.edu.cn/rbrdetector . Proteins 2014; 82:2455–2471. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
The antigen‐antibody interaction determines the sensitivity and specificity of competitive immunoassay for hapten detection. In this paper, the specificity of a monoclonal antibody against alternariol‐like compounds was evaluated through indirect competitive ELISA. The results showed that the antibody had cross‐reactivity with 33 compounds with the binding affinity (expressed by IC50) ranging from 9.4 ng/mL to 12.0 μg/mL. All the 33 compounds contained a common moiety and similar substituents. To understand how this common moiety and substituents affected the recognition ability of the antibody, a three‐dimensional quantitative structure‐activity relationship (3D‐QSAR) between the antibody and the 33 alternariol‐like compounds was constructed using comparative molecular field analysis (CoMFA) and comparative molecular similarity indices analysis (CoMSIA) methods. The q2 values of the CoMFA and CoMSIA models were 0.785 and 0.782, respectively, and the r2 values were 0.911 and 0.988, respectively, indicating that the models had good predictive ability. The results of 3D‐QSAR showed that the most important factor affecting antibody recognition was the hydrogen bond mainly formed by the hydroxyl group of alternariol, followed by the hydrophobic force mainly formed by the methyl group. This study provides a reference for the design of new hapten and the mechanisms for antibody recognition.  相似文献   

11.
Phosphofructokinase‐1 from Saccharomyces cerevisiae is composed of two types of subunits, α and β. Subunit‐specific monoclonal antibodies were raised to elucidate structural and functional properties of both subunits. One monoclonal antibody, α‐F3, binds to an epitope either at the C‐terminal or at the N‐terminal part of the α‐polypeptide chain. By screening a heptapeptide library with this monoclonal antibody, a set of heptapeptides was selected, which contained the consensus sequences D–A–F and D–S–F. Two heptapeptides with these motifs were synthesized in order assess their capacity to inhibit the binding of antibody α‐F3 to native phosphofructokinase‐1. The peptide G–I–K–D–A–F–L inhibited the binding more strongly (IC50 = 1.5 µM) than the peptide A–P–W–H–D–S–F (IC50 = 33.3 µM). Sequence matching revealed the presence of the D–A–F motif in the polypeptide chain of phosphofructokinase‐1 at amino acid position 172–174. As a control, the nonapeptide A–P–T–S–K–D–A–F–L which corresponds to the sequence of the putative epitope was tested in the inhibition assay. In view of the high inhibitory capacity (IC50 = 0.3 µM) it was concluded that this nonapeptide represents the continuous epitope of phosphofructokinase‐1 that is recognized by antibody α‐F3. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

12.
In addition to the well‐established sense‐antisense complementarity abundantly present in the nucleic acid world and serving as a basic principle of the specific double‐helical structure of DNA, production of mRNA, and genetic code‐based biosynthesis of proteins, sense‐antisense complementarity is also present in proteins, where sense and antisense peptides were shown to interact with each other with increased probability. In nucleic acids, sense‐antisense complementarity is achieved via the Watson‐Crick complementarity of the base pairs or nucleotide pairing. In proteins, the complementarity between sense and antisense peptides depends on a specific hydropathic pattern, where codons for hydrophilic and hydrophobic amino acids in a sense peptide are complemented by the codons for hydrophobic and hydrophilic amino acids in its antisense counterpart. We are showing here that in addition to this pattern of the complementary hydrophobicity, sense and antisense peptides are characterized by the complementary order‐disorder patterns and show complementarity in sequence distribution of their disorder‐based interaction sites. We also discuss how this order‐disorder complementarity can be related to protein evolution.  相似文献   

13.
The interaction of antigen (Ag) and antibody (Ab) with poly diallyldimethylammonium chloride (PDDA) in aqueous solutions has been studied by optical absorption and resonance light‐scattering (RLS) spectroscopies. The formation of the three‐component‐complex is due to aggregates of Ab or Ag with PDDA by electrostatic interaction and aggregates of Ab with Ag by immunoreaction. The influences of some experimental factors, including incubation time, pH value, concentration of PDDA and concentration of Ab, on the aggregation process have also been studied. A linear relationship between the concentration of Ag and the RLS intensity was found. Under the optimal conditions, for a given concentration of Ab (4.6 µg/mL), the enhancement of RLS intensity is in proportion to the concentration of Ag in the range 0.03–0.83 µg/mL. The RLS could, in combination with immunoassay, be a rapid and sensitive detection method for Ag. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
15.
Although green fluorescence protein (GFP) and its antibody are widely used to track a protein or a cell in life sciences, the binding behavior between them remains unclear. In this work, diazo coupling method that synthesized a new stationary GFP was oriented immobilized on the surface of macro‐porous silica gel by a phase. The stationary phase was utilized to confirm the validation of injection amount‐dependent analysis in exploring protein–protein interaction that use GFP antibody as a probe. GFP antibody was proved to have one type of binding site on immobilized GFP. The number of binding site and association constant were calculated to be (6.41 ± 0.76) × 10‐10 M and (1.39 ± 0.12) × 109 M‐1. Further analysis by molecular docking showed that the binding of GFP to its antibody is mainly driven by hydrogen bonds and salt bridges. These results indicated that injection amount‐dependent analysis is capable of exploring the protein–protein interactions with the advantages of ligand and time saving. It is a valuable methodology for the ligands, which are expensive or difficult to obtain. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
17.
18.
19.
Neisseria meningitidis is a globally important cause of bacterial meningitis and septicemia. No comprehensive antimeningococcal vaccine is available, largely as a consequence of the high sequence diversity of those surface proteins that could function as components of a vaccine. One such component is the protein PorA, a major surface porin of this Gram-negative organism that has been used in a number of experimental and licensed vaccines. Here we describe a series of experiments designed to investigate the consequences for antibody recognition of sequence diversity within a PorA antigen. The binding of a 14-residue peptide, corresponding to the P1.2 subtype antigen, to the MN16C13F4 monoclonal antibody was sensitive to mutation of five out of the six residues within the epitope sequence. The crystal structure of the antibody Fab fragment, determined in complex with the peptide antigen, shows a remarkably hydrophobic binding site and interactions between the antigen and antibody are dominated by apolar residues. Nine intrachain hydrogen bonds are formed within the antigen which maintain the beta-hairpin conformation of the peptide. These hydrogen bonds involve residues that are highly conserved amongst different P1.2 sequence variants, suggesting that some positions may be conserved for structural reasons in these highly polymorphic regions. The sensitivity of antibody recognition of the antigen towards mutation provides a structural explanation for the widespread sequence variation seen in different PorA sequences in this region. Single point mutations are sufficient to remove binding capability, providing a rationale for the manner in which different meningococcal PorA escape variants arise.  相似文献   

20.
Heat‐stable toxin type I (STa)‐ovalbumin chemical conjugates are currently used as the only coating antigen in ELISA to titrate anti‐STa antibodies for ETEC vaccine candidates. STa‐ovalbumin chemical conjugation requires STa toxin purification, a process that can be carried out by only a couple of laboratories and often with a low yield. Alternative ELISA coating antigens are needed for anti‐STa antibody titration for ETEC vaccine development. In the present study, we genetically fused STa toxin gene (three copies) to a modified chicken ovalbumin gene for genetic fusion 3×STa‐ovalbumin, and examined application of this fusion protein as an alternative coating antigen of anti‐STa antibody titration ELISA. Data showed fusion protein 3×STa‐ovalbumin was effectively expressed and extracted, and anti‐STa antibody titration ELISA using this recombinant protein (25 ng per well) or STa‐ovalbumin chemical conjugates (10 ng/well) showed the same levels of sensitivity and specificity. Furthermore, mice immunized with this fusion protein developed anti‐STa antibodies; induced antibodies showed in vitro neutralization activity against STa toxin. These results indicate that recombinant fusion protein 3×STa‐ovalbumin is an effective ELISA coating antigen for anti‐STa antibody titration, enabling a reliable reagent supply to make standardization of STa antibody titration assay feasible and to accelerate ETEC vaccine development.
  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