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 31 毫秒
1.
The mechanism of assembly/dissociation of a recombinant water-soluble class I major histocompatibility complex (MHC) H-2Kb molecule was studied by a real-time fluorescence resonance energy transfer method. Like the H-2Kd ternary complex [Gakamsky et al. (1996) Biochemistry 35, 14841-14848], the interactions among the heavy chain, beta2-microglobulin (beta2m), and antigenic peptides were found to be controlled by an allosteric mechanism. Association of the heavy chain with beta2m increased peptide binding rate constants by more than 2 orders of magnitude and enhanced affinity of the heavy-chain molecule for peptides. Interaction of peptides with the heavy-chain binding site, in turn, increased markedly the affinity of the heavy chain for beta2m. Binding of peptide variants of the ovalbumin sequence (257-264) to the heavy chain/beta2m heterodimer was found to be a biphasic reaction. The fast phase was a second-order process with nearly the same rate constants as those of binding of peptides derived from the influenza virus nucleoprotein 147-155 to the H-2Kd heavy chain/beta2m heterodimer [(3.0 +/- 1.0) x 10(-6) M-1 s-1 at 37 degrees C]. The slow phase was a result of both the ternary complex assembly from the "free" heavy chain, beta2m, and peptide as well as an intramolecular conformational transition within the heavy chain/beta2m heterodimer to a peptide binding conformation. Biexponential kinetics of peptide or beta2m dissociation from the ternary complex were observed. They suggest that it can exist in two conformations. The rate constants of beta2m dissociation from the H-2Kb ternary complex were, in the limits of experimental accuracy, independent of the structure of the bound peptide, though their affinities differed by an order of magnitude. Dissociation of peptides from the Kb heavy chain was always faster than from the ternary complexes, yet the heavy chain/peptide complexes were considerably more stable compared with their Kd/nucleoprotein peptide counterparts.  相似文献   

2.
Engagement of TCR by its ligand, the MHC/peptide complex, causes T cell activation. T cells respond positively to stimulation with agonists, and are inhibited by antagonist MHC/peptide ligands. Failure to induce proper conformational changes in the TCR or fast TCR/MHC dissociation are the leading models proposed to explain anergy induction by antagonist ligands. In this study, we demonstrate that presentation of between 1 and 10 complexes of agonist/MHC II by unfixed APC induces T cell anergy that persists up to 7 days and has characteristics similar to anergy induced by antagonist ligand or TCR occupancy without costimulation. Furthermore, anergy-inducing doses of hemagglutinin 306-318 peptide led to the engagement of less than 1000 TCR/CD3 complexes. Thus, engagement of a subthreshold number of TCR by either a low density of agonist/MHC or a 2-3 orders of magnitude higher density of antagonist/MHC causes anergy. Moreover, we show that anergy induced by low agonist concentrations is inhibited in the presence of IL-2 or cyclosporin A, suggesting involvement of the calcineurin signaling pathway.  相似文献   

3.
Dimerization of class II major histocompatibility complex (MHC) plays an important role in the MHC biological function. Mycoplasma arthritidis-derived mitogen (MAM) is a superantigen that can activate large fractions of T cells bearing specific T cell receptor Vbeta elements. Here we have used structural, sedimentation, and surface plasmon resonance detection approaches to investigate the molecular interactions between MAM and the class II MHC molecule HLA-DR1 in the context of a hemagglutinin peptide-(306-318) (HA). Our results revealed that zinc ion can efficiently induce the dimerization of the HLA-DR1/HA complex. Because the crystal structure of the MAM/HLA-DR1/hemagglutinin complex in the presence of EDTA is nearly identical to the structure of the complex crystallized in the presence of zinc ion, Zn(2+) is evidently not directly involved in the binding between MAM and HLA-DR1. Sedimentation and surface plasmon resonance studies further revealed that MAM binds the HLA-DR1/HA complex with high affinity in a 1:1 stoichiometry, in the absence of Zn(2+). However, in the presence of Zn(2+), a dimerized MAM/HLA-DR1/HA complex can arise through the Zn(2+)-induced DR1 dimer. In the presence of Zn(2+), cooperative binding of MAM to the DR1 dimer was also observed.  相似文献   

4.
The Escherichia coli nucleotide exchange factor GrpE accelerates the rate of ADP dissociation from high affinity ADP-DnaK, thus enabling ATP binding and transition to the low affinity state. We show here that GrpE, in the absence of ATP, accelerates the rates of the forward and reverse reaction ADP-DnaK-P right harpoon over left harpoon ADP-DnaK + P, where P denotes peptide substrate. Specifically, the binding of GrpE to an ADP-DnaK-P (or DnaK-P) complex increases koff and kon by approximately 200-fold and approximately 60-fold, respectively. The results are consistent with a GrpE- induced conformational change in the C-terminal polypeptide binding domain of an ADP-DnaK molecule, which results in a unique low affinity intermediate from which peptide can dissociate. A simulation of peptide dissociation from DnaK as a function of the [ATP] / [ADP] ratio shows that GrpE induced peptide dissociation from ADP-DnaK is important at elevated cellular concentrations of ADP, which typically occur upon stress.  相似文献   

5.
Joshi RV  Zarutskie JA  Stern LJ 《Biochemistry》2000,39(13):3751-3762
Peptide binding reactions of class II MHC proteins exhibit unusual kinetics, with extremely slow apparent rate constants for the overall association (<100 M(-)(1) s(-)(1)) and dissociation (<10(-)(5) s(-)(1)) processes. Various linear and branched pathways have been proposed to account for these data. Using fluorescence resonance energy transfer between tryptophan residues in the MHC peptide binding site and aminocoumarin-labeled peptides, we measured real-time kinetics of peptide binding to empty class II MHC proteins. Our experiments identified an obligate intermediate in the binding reaction. The observed kinetics were consistent with a binding mechanism that involves an initial bimolecular binding step followed by a slow unimolecular conformational change. The same mechanism is observed for different peptide antigens. In addition, we noted a reversible inactivation of the empty MHC protein that competes with productive binding. The implications of this kinetic mechanism for intracellular antigen presentation pathways are discussed.  相似文献   

6.
Mycoplasma arthritidis-derived mitogen (MAM) is a superantigen that can activate large fractions of T cells bearing particular TCR Vbeta elements. Here we report the crystal structure of MAM complexed with a major histocompatibility complex (MHC) antigen, HLA-DR1, loaded with haemagglutinin peptide 306-318 (HA). The structure reveals that MAM has a novel fold composed of two alpha-helical domains. This fold is entirely different from that of the pyrogenic superantigens, consisting of a beta-grasped motif and a beta barrel. In the complex, the N-terminal domain of MAM binds orthogonally to the MHC alpha1 domain and the bound HA peptide, and to a lesser extent to the MHC beta1 domain. Two MAM molecules form an asymmetric dimer and cross-link two MHC antigens to form a plausible, dimerized MAM-MHC complex. These data provide the first crystallographic evidence that superantigens can dimerize MHC molecules. Based on our structure, a model of the TCR2MAM2MHC2 complex is proposed.  相似文献   

7.
Peptide presentation by MHC class II is of critical importance to the function of CD4+ T cells. HLA-DM resides in the endosomal pathway and edits the peptide repertoire of newly synthesized MHC class II molecules before they are exported to the cell surface. HLA-DM ensures MHC class II molecules bind high affinity peptides by targeting unstable MHC class II:peptide complexes for peptide exchange. Research over the past decade has implicated the peptide N-terminus in modulating the ability of HLA-DM to target a given MHC class II:peptide combination. In particular, attention has been focused on both the hydrogen bonds between MHC class II and peptide, and the occupancy of the P1 anchor pocket. We sought to solve the crystal structure of a HLA-DR1 molecule containing a truncated hemagglutinin peptide missing three N-terminal residues compared to the full-length sequence (residues 306–318) to determine the nature of the MHC class II:peptide species that binds HLA-DM. Here we present structural evidence that HLA-DR1 that is loaded with a peptide truncated to the P1 anchor residue such that it cannot make select hydrogen bonds with the peptide N-terminus, adopts the same conformation as molecules loaded with full-length peptide. HLA-DR1:peptide combinations that were unable to engage up to four key hydrogen bonds were also unable to bind HLA-DM, while those truncated to the P2 residue bound well. These results indicate that the conformational changes in MHC class II molecules that are recognized by HLA-DM occur after disengagement of the P1 anchor residue.  相似文献   

8.
Class II MHC glycoproteins bind short (7-25 amino acid) peptides in an extended type II polyproline-like conformation and present them for immune recognition. Because empty MHC is unstable, measurement of the rate of the second-order reaction between peptide and MHC is challenging. In this report, we use dissociation of a pre-bound peptide to generate the active, peptide-receptive form of the empty class II MHC molecule I-Ek. This allows us to measure directly the rate of reaction between active, empty I-Ek and a set of peptides that vary in structure. We find that all peptides studied, despite having highly variable dissociation rates, bind with similar association rate constants. Thus, the rate-limiting step in peptide binding is minimally sensitive to peptide side-chain structure. An interesting complication to this simple model is that a single peptide can sometimes bind to I-Ek in two kinetically distinguishable conformations, with the stable peptide-MHC complex isomer forming much more slowly than the less-stable one. This demonstrates that an additional free-energy barrier limits the formation of certain specific MHC-peptide complex conformations.  相似文献   

9.
The connection between experimentally measured values of ED50 (concentration of added peptide required to bind half of the protein), which characterize peptide-protein binding and the equilibrium dissociation constant of peptide-protein complex Kd (affinity) is considered. It is shown and confirmed by experimental studies that in certain cases, as a result of the absence of equilibrium in the system, the value of Kd could be much less than the experimental value of ED50, but not equal to that as commonly assumed. This is especially applicable to the formation of peptide-MHC complexes with low dissociation rates (strong binding), which may require longer time-intervals to reach equilibrium. Thus the search of the good binding peptides based on finding ones with the smallest measured values' of ED50 may result in missing the best binders with the lowest values of dissociation constant (highest affinity). To analyze the problem we considered the formal chemical kinetics of peptide-protein binding. Experimental studies of peptide binding was performed to obtain the parameters of the kinetic model. According to the predictions of the model, it was confirmed that peptide binding occurs through the preceding step, which is either a release of an endogenous peptide or some conformational change of the molecule. The half decay time for this process was determined to be approximately 3 h. Based on the model developed, a new effective method for determination of the dissociation rates of peptide-MHC complexes and the equilibrium dissociation constants Kd was proposed, which implies the comparison of binding levels (ED50) at different instants of time. This method works especially well for the peptide-MHC complexes with relatively slow dissociation rates (stable complexes), for which the direct off-rate measurements as well as obtaining equilibrium binding data to determine Kd are highly time consuming and not very reliable.  相似文献   

10.
《Current biology : CB》1999,9(18):999-S1
Background: Newly synthesised peptide-receptive major histocompatibility complex (MHC) class I molecules form a transient loading complex in the endoplasmic reticulum with the transporter associated with antigen processing (TAP) and a set of accessory proteins. Binding of peptide to the MHC class I molecule is necessary for dissociation of the MHC class I molecule from the complex with TAP, but other components of the complex might also be involved. To investigate the role of TAP in this process, mutations that block nucleotide binding were introduced into the ATP-binding site of TAP.Results: Mutant TAP formed apparently normal loading complexes with MHC class I molecules and accessory components, but had no nucleotide-binding or peptide-transport activity. Nevertheless, whereas wild-type loading complexes in detergent lysates could be dissociated by addition of peptides that bind MHC class I molecules, mutant complexes could not be dissociated in this way. Depletion of nucleotide diphosphates or triphosphates from wild-type lysates blocked peptide-mediated dissociation of MHC class I molecules, which could be reversed by readdition of nucleotide diphosphates or triphosphates. Complexes between mutant TAP and MHC class I molecules remained associated in vivo until they were degraded. Disruption of nucleotide binding also eliminated TAP's peptide-binding activity.Conclusions: Peptide-mediated dissociation of the MHC class I molecule from the loading complex depends on conformational signals arising from TAP. Integrity of the nucleotide-binding site is required not only for transmission of this conformational signal to the loading complex, but also for binding of peptide to TAP. Thus, the dynamic activity of the loading complex is synchronised with the nucleotide-mediated peptide-binding and transport cycle of TAP.  相似文献   

11.
Ligand binding is a thermodynamically cooperative process in many biochemical systems characterized by the conformational flexibility of the reactants. However, the contribution of conformational entropy to cooperativity of ligation needs to be elucidated. Here, we perform kinetic and thermodynamic analyses on a panel of cycle-mutated peptides, derived from influenza H3 HA(306-319), interacting with wild type and a mutant HLA-DR. We observe that, within a certain range of peptide affinity, this system shows isothermal entropy-enthalpy compensation (iEEC). The incremental increases in conformational entropy measured as disruptive mutations are added in the ligand or receptor are more than sufficient in magnitude to account for the experimentally observed lack of free-energy decrease cooperativity. Beyond this affinity range, compensation is not observed, and therefore, the ability of the residual interactions to form a stable complex decreases in an exponential fashion. Taken together, our results indicate that cooperativity and iEEC constitute the thermodynamic epiphenomena of the structural fluctuation that accompanies ligand-receptor complex formation in flexible systems. Therefore, ligand binding affinity prediction needs to consider how each source of binding energy contributes synergistically to the folding and kinetic stability of the complex in a process based on the trade-off between structural tightening and restraint of conformational mobility.  相似文献   

12.
 A comprehensive analysis was carried out of the tri-molecular complex of peptide, major histocompatibility class II molecule, and T-cell receptor (TcR) involved in the recognition of the promiscuous HA (306–318) peptide, restricted by one of two closely related HLA-DR alleles, HLA-DRB1*0101 and HLA-DRB1*0103. These two DR molecules differ by only three amino acids at positions 67, 70, and 71, in the third variable region of the DRB1 chain. None of the HA (306–318)-specific T-cell clones restricted by these two DR molecules tolerated amino acid substitution at the peptide-binding position 71, despite the fact that the substitution did not interfere with peptide binding. The majority of the DRB1*0103-restricted clones tolerated substitution of the amino acid at the TcR-contacting position 70, while the DRB1*0101-restricted T cells did not. Biased usage of TRVA and TRVB segments was observed for the DRB1*0103-restricted clones; in contrast, apparently random usage was seen in the DRB1*0101-restricted T cells. Finally, limiting dilution analysis revealed a lower frequency of T cells reactive with the HA peptide in a DRB1*0103 compared with a DRB1*0101 individual. Taken together these data suggest that biased TcR gene usage may reflect a relatively low precursor frequency of T cells, and the need for clonal expansion of a limited set of high avidity T cells. Received: 7 August 1998 / Revised: 19 November 1998  相似文献   

13.
The interaction of vanadate ions with the Ca-ATPase from sarcoplasmic reticulum vesicles was studied in a native and a fluorescein-labeled ATPase preparation (Pick, U., and Karlish, S. J. D. (1980) Biochim. Biophys. Acta 626, 255-261). Vanadate induced a fluorescence enhancement in a fluorescein-labeled enzyme, indicating that it shifts the equilibrium between the two conformational states of the enzyme by forming a stable E2-Mg-vanadate complex (E2 is the low affinity Ca2+ binding conformational state of the sarcoplasmic reticulum Ca-ATPase). Indications for tight binding of vanadate to the enzyme (K1/2 = 10 microM) in the absence of Ca2+ and for a slow dissociation of vanadate from the enzyme in the presence of Ca2+ are presented. The enzyme-vanadate complex was identified by the appearance of a time lag in the onset of Ca2+ uptake and by a slowing of the fluorescence quenching response to Ca2+. Ca2+ prevented the binding of vanadate to the enzyme. Pyrophosphate (Kd = 2 mM) and ATP (Kd = 25 microM) competitively inhibited the binding of vanadate, indicating that vanadate binds to the low affinity ATP binding site. Binding of vanadate inhibited the high affinity Ca2+ binding to the enzyme at 4 degrees C. Vanadate also inhibited the phosphorylation reaction by inorganic phosphate (Ki = 10 microM) but had no effect on the phosphorylation by ATP. It is suggested that vanadate binds to a special region in the low affinity ATP binding site which is exposed only in the E2 conformation of the enzyme in the absence of Ca2+ and which controls the rate of the conformation transition in the dephosphorylated enzyme. The implications of these results to the role of the low affinity ATP binding sites are discussed.  相似文献   

14.
Hennecke J  Carfi A  Wiley DC 《The EMBO journal》2000,19(21):5611-5624
An alphabeta T-cell receptor (alphabetaTCR)/hemagglutinin (HA) peptide/human leukocyte antigen (HLA)-DR1 complex was stabilized by flexibly linking the HA peptide with the human HA1.7 alphabetaTCR, to increase the local concentration of the interacting proteins once the peptide has been loaded onto the major histocompatibility complex (MHC) molecule. The structure of the complex, determined by X-ray crystallography, has a binding mode similar to that of the human B7 alphabetaTCR on a pMHCI molecule. Twelve of the 15 MHC residues contacted are at the same positions observed earlier in class I MHC/peptide/TCR complexes. One contact, to an MHC loop outside the peptide-binding site, is conserved and specific to pMHCII complexes. TCR gene usage in the response to HA/HLA-DR appears to conserve charged interactions between three lysines of the peptide and acidic residues on the TCR.  相似文献   

15.
Gakamsky DM  Davis DM  Strominger JL  Pecht I 《Biochemistry》2000,39(36):11163-11169
Class I major histocompatibility complex (MHC) heterodimer, composed of human leukocyte antigen (HLA)-A2 heavy chain and human beta(2)-microglobulin (beta(2)m), was produced by denaturation and gel filtration of the recombinant water-soluble HLA-A2/beta(2)m/peptide ternary complex in 8 M urea Tris-HCl buffer, followed by refolding of the separated chains without peptide. Peptide affinity and kinetics of the ternary complex formation and dissociation were investigated in real time by monitoring the fluorescence resonance energy transfer (FRET) from intrinsic HLA-A2 heavy-chain tryptophans to a dansyl fluorophore conjugated to the bound peptide. Peptide binding to the heterodimer was a second order process with rate constants linearly dependent upon temperature in Arrhenius coordinates over 0-20 degrees C. The binding rate constant of pRT6C-dansyl [ILKEPC(dansyl)HGV] at 37 degrees C evaluated by extrapolation of the Arrhenius plot was (2.0 +/- 0.5) x 10(6) M(-1) s(-1). Association of the heavy chain with beta(2)m was a first order process, apparently controlled by a conformational transition in the heavy chain. One of these conformations bound to beta(2)m to form the heavy chain/beta(2)m heterodimer whereas the second conformer oligomerized. Peptide dissociation from the ternary complex was a first-order reaction over the temperature range 20-37 degrees C, suggesting that the ternary complex also exists in two conformations. Taken together, the present data suggest that association of beta(2)m changes the HLA-A2 heavy-chain conformation thereby promoting peptide binding. Peptide dissociation from the ternary complex induces dissociation of the heavy-chain/beta(2)m heterodimer thereby causing oligomerization of the heavy chain. The lability of the HLA-A2/beta(2)m heterodimer and the strong tendency of the "free" heavy chain to oligomerize may provide an efficient mechanism for control of antigen presentation under physiological conditions by reducing the direct loading of HLA with exogenous peptide at the cell surface.  相似文献   

16.
Affinity-purified major histocompatability complex (MHC) class II molecules are known to bind antigenic peptide in vitro. This peptide-bound MHC class II is known to undergo a change in structure upon stable binding of antigenic peptide. Previous results from our, and other laboratories, have suggested a relationship between MHC class II structure and peptide association that enables class II to enter into a stable conformation upon peptide binding. In this report we describe that stable binding of high-affinity antigenic peptide to MHC class II molecule results in transition of aggregated purified MHC class II proteins to a stable heterodimeric state. Such transition was demonstrated by using purified human HLA-DR2 class II molecule and high-affinity myelin basic protein (MBP) 83-102)Y83 peptide. Highly aggregated purified DR2 (high molecular weight; HMW) was first separated from heterodimer (low molecular weight: LMW) in the presence of 50-fold molar excess of MBP(83-102)Y83 peptide. We then show that the aggregated HMW preparation can be successfully converted into a stable dimer by further incubation with MBP(83-102)Y83 and changing various binding parameters such as pH, temperature, reducing agent, and peptide concentrations. Under optimized conditions, the highly aggregated inactive DR2 molecules can be completely loaded with the antigenic peptide. The transformed heterodimers with bound peptide prepared by this method are biologically active, as shown by their ability to induce the production of gamma-interferon by SS8T-transformed human T cells. These results suggest that in solution, MHC class II molecules may be aggregated in the absence of bound peptide. Such aggregated MHC class II molecules can be converted to stable and biologically active heterodimers in the presence of high-affinity antigenic peptide.  相似文献   

17.
Class I major histocompatibility complex (MHC) molecules bind short peptides derived from proteins synthesized within the cell. These complexes of peptide and class I MHC (pMHC) are transported from the endoplasmic reticulum to the cell surface. If a clonotypic T cell receptor expressed on a circulating T cell binds to the pMHC complex, the cell presenting the pMHC is killed. In this manner, some tumor cells expressing aberrant proteins are recognized and removed by the immune system. However, not all tumors are recognized efficiently. One reason hypothesized for poor T cell recognition of tumor-associated peptides is poor binding of those peptides to class I MHC molecules. Many peptides, derived from the proto-oncogene HER-2/neu have been shown to be recognized by cytotoxic T cells derived from HLA-A2(+) patients with breast cancer and other adenocarcinomas. Seven of these peptides were found to bind with intermediate to poor affinity. In particular, GP2 (HER-2/neu residues 654-662) binds very poorly even though it is predicted to bind well based upon the presence of the correct HLA-A2.1 peptide-binding motif. Altering the anchor residues to those most favored by HLA-A2.1 did not significantly improve binding affinity. The crystallographic structure shows that unlike other class I-peptide structures, the center of the peptide does not assume one specific conformation and does not make stabilizing contacts with the peptide-binding cleft.  相似文献   

18.
The ATP-binding cassette (ABC) transporter TAP plays an essential role in antigen processing and immune response to infected or malignant cells. TAP translocates proteasomal degradation products from the cytosol into the endoplasmic reticulum, where MHC class I molecules are loaded with these peptides. Kinetically stable peptide-MHC complexes are transported to the cell surface for inspection by cytotoxic T lymphocytes. The transport cycle of TAP is initiated by peptide binding, which is responsible for peptide selection and for stimulation of ATP-hydrolysis and subsequent translocation. Here we have analysed the driving forces for the formation of the peptide-TAP complex by kinetic and thermodynamic methods. First, the apparent peptide association and dissociation rates were determined at various temperatures. Strikingly, very high activation energies for apparent association (E(a)(ass)=106 kJmol(-1)) and dissociation (E(a)(diss)=80 kJmol(-1)) of the peptide-TAP complex were found. Next, the temperature-dependence of the peptide affinity constants was investigated by equilibrium-binding assays. Along with calculations of free enthalpy deltaG, enthalpy deltaH and entropy deltaS, a large positive change in heat capacity was resolved (deltaC degrees =23 kJmol(-1)K(-1)), indicating a fundamental structural reorganization of the TAP complex upon peptide binding. The inspection of the conformational entropy reveals that approximately one-fourth of all TAP residues is rearranged. These thermodynamic studies indicate that at physiological temperature, peptide binding is endothermic and driven by entropy.  相似文献   

19.
Stable, recombinant, water-soluble complexes of HLA-A2 and HLA-B27 were reconstituted from 125I-labeled beta 2-microglobulin (beta 2m), a synthetic peptide, and HLA H chain fragments expressed as inclusion bodies in the Escherichia coli cytoplasm. Using this system, we were able to show: 1) the t1/2 of beta 2m dissociation from HLA complexes at 37 degrees C varied from approximately 40 h to less than 1 h, depending on the peptide employed for reconstitution. Peptide length and composition were found to be critical factors in determining the beta 2m dissociation rate. Endogenous peptides form complexes that are about as stable as those formed with typical antigenic peptides. 2) Peptide exchange reactions, in which an exogenous peptide replaces the peptide that is already bound by the class I molecule, proceed readily for complexes that have rapid beta 2m dissociation rates. Thus, difficulties in demonstrating peptide binding to complexes that contain endogenous peptides can be attributed to the stability of the endogenous peptide/class I molecule complex. 3) The peptide exchange reaction does not require concomitant beta 2m dissociation. 4) Distal parts of the class I molecule, which are not directly involved in peptide binding or beta 2m binding, have a major impact on the stability of class I molecules. Thus, these studies show that the dissociation rate of beta 2m is an excellent measure of how tightly a given peptide binds to class I MHC molecules, that the ability to bind peptide is tightly coupled to the binding of beta 2m and vice versa, and that regions of the molecule distal from the binding site influence the stability of peptide binding.  相似文献   

20.
Conformational dependence of TCR contact residues of the H-2Kb molecule on the two buried tyrosine side chains of the vesicular stomatitis virus (VSV)-8 peptide was investigated by systematic substitutions of the tyrosines with phenylalanine, p-fluorophenylalanine (pFF), or p-bromophenylalanine (pBrF). The results of peptide competition CTL assays revealed that all of the peptide variants, except for the pBrF analogues, had near-native binding to the H-2Kb molecule. Epitope-mapped anti-H-2Kb mAbs detected conformational differences among H-2Kb molecules stabilized with these VSV-8 variants on RMA-S cells. Selective recognition of the VSV-8 analogues was displayed by a panel of three H-2Kb-restricted, anti-VSV-8 TCRs. Thus, these substitutions result in an antigenically significant conformational change of the MHC molecular surface structure at both C and D pockets, and the effect of this change on cognate T cell recognition is dependent on the TCR structure. Our results confirm that the structure of buried peptide side chains can determine the surface conformation of the MHC molecule and demonstrate that even a very subtle structural nuance of the buried side chain can be incorporated into the surface conformation of the MHC molecule. The ability of buried residues to modulate this molecular surface augments the number of residues on the MHC-peptide complex that can be recognized as "foreign" by the CD8+ T cell repertoire and allows for a higher level of antigenic discrimination. This may be an important mechanism to expand the total number of TCR specificities that can respond to a single peptide determinant.  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