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.
DM catalyzes the exchange of peptides bound to Class II major histocompatibility complex (MHC) molecules. Because the dissociation and association components of the overall reaction are difficult to separate, a detailed mechanism of DM catalysis has long resisted elucidation. UV irradiation of DR molecules loaded with a photocleavable peptide (caged Class II MHC molecules) enabled synchronous and verifiable evacuation of the peptide-binding groove and tracking of early binding events in real time by fluorescence polarization. Empty DR molecules generated by photocleavage rapidly bound peptide but quickly resolved into species with substantially slower binding kinetics. DM formed a complex with empty DR molecules that bound peptide with even faster kinetics than empty DR molecules just having lost their peptide cargo. Mathematical models demonstrate that the peptide association rate of DR molecules is substantially higher in the presence of DM. We therefore unequivocally establish that DM contributes directly to peptide association through formation of a peptide-loading complex between DM and empty Class II MHC. This complex rapidly acquires a peptide analogous to the MHC class I peptide-loading complex.  相似文献   

2.
Solution studies have demonstrated the existence of two functionally distinct isomers of empty class II MHC: an active isomer that binds peptide and an inactive isomer that does not. Empty MHC molecules on the surface of APCs can load antigenic peptides directly from the extracellular medium, facilitating the generation of a diverse peptide repertoire for T cell presentation. In this report, we examine I-Ek on the surface of Chinese hamster ovary cells with respect to the active and inactive isomers. As in the case of purified soluble active I-Ek, active I-Ek on the cell surface is unstable, decaying to the inactive form in approximately 14 min. Evidence is presented suggesting that at steady state <1% of the total cell surface I-Ek is active and that a significant fraction of these active molecules originates from intracellular pools as well as reactivation of inactive cell surface I-EK.  相似文献   

3.
We have analyzed the interaction of the hen egg-white lysozyme (HEL) peptide 107-116 with the MHC class II molecule I-Ek, using truncated and single residue substitution analogues to measure activation of I-Ek-restricted, 107-116-specific T cell hybridomas and competition for Ag presentation by I-Ek molecules. These results have been compared with previous findings on the interaction of the same peptide with the I-Ed molecule. Stimulation of T cell hybridomas by truncated peptides defines the sequence 108-116 as the minimum epitope necessary for activation of both I-Ek- and I-Ed-restricted T cell hybridomas. Substitution analysis pinpoints three residues (V109, A110, and K116) in the sequence 108-116 as being critical for binding to I-Ek molecules and demonstrates the involvement of most other residues in recognition by T cells. Results previously obtained for binding of HEL 107-116 to I-Ed molecules indicated that peptide residues R112, R114, and K116 were critical for interaction with I-Ed. Comparison of these results indicates a difference in the likely MHC contact residues between the HEL sequence 108-116 and I-Ed or I-Ek molecules, suggesting that the same HEL peptide assumes a different conformation in the binding site of these two MHC molecules. This in turn affects residues interacting with the specific T cell receptor. According to the hypothetical tridimensional structure predicted for class II molecules, the difference in MHC contact residues observed within the sequence 108-116 can be related to polymorphic amino acids in the binding site of I-Ek and I-Ed molecules. A search through published binding data for a common pattern in this and other I-Ek-binding peptides has permitted us to derive a possible motif for predicting peptide binding to I-Ek molecules. This putative motif was tested by determining binding to I-Ek of an unbiased panel of about 150 synthetic peptides. Binding data indeed demonstrate the presence of this motif in the majority of good binders to I-Ek molecules.  相似文献   

4.
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   

5.
Class II major histocompatibility complex (MHC) proteins bind peptides and present them at the cell surface for interaction with CD4+ T cells as part of the system by which the immune system surveys the body for signs of infection. Peptide binding is known to induce conformational changes in class II MHC proteins on the basis of a variety of hydrodynamic and spectroscopic approaches, but the changes have not been clearly localized within the overall class II MHC structure. To map the peptide-induced conformational change for HLA-DR1, a common human class II MHC variant, we generated a series of monoclonal antibodies recognizing the beta subunit that are specific for the empty conformation. Each antibody reacted with the empty but not the peptide-loaded form, for both soluble recombinant protein and native protein expressed at the cell surface. Antibody binding epitopes were characterized using overlapping peptides and alanine scanning substitutions and were localized to two distinct regions of the protein. The pattern of key residues within the epitopes suggested that the two epitope regions undergo substantial conformational alteration during peptide binding. These results illuminate aspects of the structure of the empty forms and the nature of the peptide-induced conformational change.  相似文献   

6.
The main functions of the T-cell receptor (TCR) involve its specific interaction with short and linear antigenic peptides bound to the major histocompatibility complex (MHC) molecules. In the absence of a 3D structure for TCR and for the TCR/peptide/MHC complex, several attempts to characterize the structural components of the TCR/peptide/MHC interaction have been made. However, this subject is still troublesome. In this paper a computer-based 3D model for a TCR/peptide/MHC complex (5C.C7/moth cytochrome c [MCC] peptide 93-103/I-Ek) was obtained. The complex surface shows a high complementarity between the 5C.C7 structure and the peptide/I-Ek molecule. The mapping of residues involved in the TCR/peptide/MHC interaction shows close agreement with mutational experiments (Jorgensen JL, Reay PA, Ehrich EW, Davis MM, 1992b, Annu Rev Immunol 10:835-873). Moreover, the results are consistent with a recent variability analysis of TCR sequences using three variability indexes (Almagro JC, Zenteno-Cuevas R, Vargas-Madrazo E, Lara-Ochoa F, 1995b, Int J Pept Protein Res 45:180-186). Accordingly, the 3D model of the 5C.C7/MCC peptide 93-103/I-Ek complex provides a framework to generate testable hypotheses about TCR recognition. Thus, starting from this model, the role played by each loop that forms the peptide/MHC binding site of the TCR is discussed.  相似文献   

7.
Saito K  Oda M  Sarai A  Azuma T  Kozono H 《Biochemistry》2004,43(31):10186-10191
We used differential scanning calorimetry to study the thermal denaturation of murine major histocompatibility complex class II, I-E(k), accommodating hemoglobin (Hb) peptide mutants possessing a single amino acid substitution of the chemically conserved amino acids buried in the I-Ek pocket (positions 71 and 73) and exposed to the solvent (position 72). All of the I-Ek-Hb(mut) molecules exhibited greater thermal stability at pH 5.5 than at pH 7.4, as for the I-Ek-Hb(wt) molecule, which can explain the peptide exchange function of MHC II. The thermal stability was strongly dependent on the bound peptide sequences; the I-Ek-Hb(mut) molecules were less stable than the I-Ek-Hb(wt) molecules, in good correlation with the relative affinity of each peptide for I-Ek. This supports the notion that the bound peptide is part of the completely folded MHC II molecule. The thermodynamic parameters for I-Ek-Hb(mut) folding can explain the thermodynamic origin of the stability difference, in correlation with the crystal structural analysis, and the limited contributions of the residues to the overall conformation of the I-Ek-peptide complex. We found a linear relationship between the denaturation temperature and the calorimetric enthalpy change. Thus, although the MHC II-peptide complex could have a diverse thermal stability spectrum, depending on the amino acid sequences of the bound peptides, the conformational perturbations are limited. The variations in the MHC II-peptide complex stability would function in antigen recognition by the T cell receptor by affecting the stability of the MHC II-peptide-T cell receptor ternary complex.  相似文献   

8.
Belmares MP  McConnell HM 《Biochemistry》2001,40(34):10284-10292
Major histocompatability complex type II proteins (MHC II) are alphabeta-heterodimeric glycoproteins that present peptides to the T cell receptor (TCR) of CD4(+) T-cells. This presentation may result in activation of these T-cells, depending on the nature of the peptide. Peptides interact specifically with MHC II with nine peptide amino acid positions, and the corresponding MHC II pocket positions are usually labeled P1-P9. However, the length of peptides binding to MHC II may be greater than nine amino acids, and therefore these peptides may potentially bind to the MHC II in more than one registry. To investigate the mechanism by which a long peptide binds to I-E(k), a murine MHC II, a chimeric peptide with two nonoverlapping registries, f-IAYLKQATKQLRMATPLLMR was designed. The IAYLKQATK peptide segment is based on moth cytochrome c 95-103 (MCC 95-103), and the QLRMATPLLMR segment is based on murine Ii CLIP 89-99 M90L (Ii CLIP 89-99 M90L). This chimeric peptide forms two isomeric complexes. The MCC and Ii CLIP registries dissociate from I-E(k) with t(1/2) values of >800 and 4.94 h, respectively. The registry composition of this MHC II/chimeric peptide complex was found to change as a function of time in approaching thermodynamic equilibrium: the results are consistent with a kinetic model that involves no intramolecular isomer interconversion. The model depicts uncorrelated binding to the MHC II determined by relative association rates to the two registries. This is followed by dissociation and subsequent rebinding, leading ultimately to a preponderance of the most stable complex. Similar results were obtained at pH 5.3. The behavior of this chimeric peptide approximates the binding of a 1:1 solution mixture of two peptides to MHC II, where the more stable complex is selected over time. We have also found that a chimeric peptide and a human MHC II, HLA-DR40401, form isomers with relative association rates to DR0401 at pH 5.3 of 15% for one isomer and 85% for the second isomer.  相似文献   

9.
The complexity of the interaction between major histocompatibility complex class II (MHC II) proteins and peptide ligands has been revealed through structural studies and crystallographic characterization. Peptides bind through side-chain "anchor" interactions with MHC II pockets and an extensive array of genetically conserved hydrogen bonds to the peptide backbone. Here we quantitatively investigate the kinetic hierarchy of these interactions. We present results detailing the impact of single side-chain mutations of peptide anchor residues on dissociation rates, utilizing two I-A(d)-restricted peptides, one of which has a known crystal structure, and 24 natural and non-natural amino acid mutant variants of these peptides. We find that the N-terminal P1, P4 and P6 anchor-pocket interactions can make significant contributions to binding stability. We also investigate the interactions of these peptides with four I-A(d) MHC II proteins, each mutated to disrupt conserved hydrogen bonds to the peptide backbone. These complexes exhibit kinetic behavior suggesting that binding energy is disproportionately invested near the peptide N terminus for backbone hydrogen bonds. We then evaluate the effects of simultaneously modifying both anchor and hydrogen bonding interactions. A quantitative analysis of 71 double mutant cycles reveals that there is little apparent cooperativity between anchor residue interactions and hydrogen bonds, even when they are directly adjacent (<5A).  相似文献   

10.
McFarland BJ  Sant AJ  Lybrand TP  Beeson C 《Biochemistry》1999,38(50):16663-16670
Proteins of the class II major histocompatibility complex (MHC) bind antigenic peptides that are subsequently presented to T cells. Previous studies have shown that most of the residues required for binding of the chicken ovalbumin (Ova) 323-339 peptide to the I-A(d) MHC class II protein are contained within the shorter 325-336 peptide. This observation is somewhat inconsistent with the X-ray structure of the Ova peptide covalently attached to I-A(d) ( structure) in which residues 323 and 324 form binding interactions with the protein. A second register for the Ova(325-336) peptide is proposed where residues 326 and 327 occupy positions similar to residues 323 and 324 in the structure. Two Ova peptides that minimally encompass the and alternate registers, Ova(323-335) and Ova(325-336), respectively, were found to dissociate from I-A(d) with distinct kinetics. The dissociation rates for both peptides were enhanced when the His81 residue of the MHC beta-chain was replaced with an asparagine. In the structure the betaH81 residue forms a hydrogen bond to the backbone carbonyl of I323. If the Ova(325-336) peptide were also bound in the register, there would be no comparable hydrogen-bond acceptor for the betaH81 side chain that could explain this peptide's sensitivity to the betaH81 replacement. The Ova(323-335) peptide that binds in the register does not stimulate a T-cell hybridoma that is stimulated by Ova(325-336) bound in the alternate register. These results demonstrate that a single peptide can bind to an MHC peptide in alternate registers producing distinct T-cell responses.  相似文献   

11.
Major histocompatability class II proteins are transmembrane alphabeta-heterodimers that present peptides to T-cells. MHC II may bind exogenous peptides directly at the cell surface. Alternatively, peptides derived from processing of endosomal protein may bind to MHC II in endosomal compartments. There, HLA-DM catalyzes the formation of peptide/MHC complexes, which are then transported to the cell surface. Here we report evidence that the peptide Ii CLIP 81-104 binds to DR*0404 in two alternate registries, whose dissociation rates, while kinetically indistinguishable at pH 5.3 and 37 degrees C, are kinetically resolved in the presence of HLA-DM. In one registry isomer, CLIP Met 91 is placed in the N-terminal P1 pocket of DR*0404, and peptide dissociation is readily catalyzed by HLA-DM. In a second proposed registry, likely with CLIP Leu 97 in the P1 pocket, the complex is substantially less sensitive to HLA-DM catalysis. Without HLA-DM, or at pH 7, the fraction of each isomer formed in solution is relatively insensitive to the duration of incubation with peptide. However, with HLA-DM, the fraction of the DM-insensitive isomer is dramatically influenced by peptide incubation time. The mechanism of isomer formation appears to be determined by the HLA-DM-modified relative association to the two registries, followed by HLA-DM-catalyzed dissociation of each isomer and rebinding, leading to a final isomer composition determined by these kinetic constants. Intramolecular isomer interconversion does not appear to be involved. The behavior of these complexes may provide a model for peptide editing by DM in endosomes.  相似文献   

12.
Antigenic peptide binding to MHC class II molecules in the endocytic pathway occurs via a multifactorial process that requires the support of a specialized lysosomal chaperone called HLA-DM. DM shows both in primary amino acid sequence and quaternary structure a high homology to both MHC class I and class II molecules. Like the peptide presenting class II molecules, DM is expressed in all professional antigen presenting cells. DM catalyzes the dissociation of peptides that do not bind stably to the class II peptide-binding groove, thereby leading to the preferential presentation of stably binding antigenic peptides. The recently discovered HLA-DO molecule is mainly expressed in B cells and associates with DM, thereby markedly affecting DM function. Like DM, the genes encoding the HLA-DO heterodimer lie within the MHC class II region and exhibit strong homology to classical class II molecules. This review evaluates the unique effects of DO on DM-mediated antigen presentation by MHC class II molecules and discusses the possible physiological relevance for the B cell-specific expression of DO and its function.  相似文献   

13.
TCRs exhibit a high degree of Ag specificity, even though their affinity for the peptide/MHC ligand is in the micromolar range. To explore how Ag specificity is achieved, we studied murine T cells expressing high-affinity TCRs engineered by in vitro evolution for binding to hemoglobin peptide/class II complex (Hb/I-Ek). These TCRs were shown previously to maintain Ag specificity, despite having up to 800-fold higher affinity. We compared the response of the high-affinity TCRs and the low-affinity 3.L2 TCR toward a comprehensive set of peptides containing single substitutions at each TCR contact residue. This specificity analysis revealed that the increase in affinity resulted in a dramatic increase in the number of stimulatory peptides. The apparent discrepancy between observed degeneracy in the recognition of single amino acid-substituted Hb peptides and overall Ag specificity of the high-affinity TCRs was examined by generating chimeric peptides between the stimulatory Hb and nonstimulatory moth cytochrome c peptides. These experiments showed that MHC anchor residues significantly affected TCR recognition of peptide. The high-affinity TCRs allowed us to estimate the affinity, in the millimolar range, of immunologically relevant interactions of the TCR with peptide/MHC ligands that were previously unmeasurable because of their weak nature. Thus, through the study of high-affinity TCRs, we demonstrated that a TCR is more tolerant of single TCR contact residue substitutions than other peptide changes, revealing that recognition of Ag by T cells can exhibit both specificity and degeneracy.  相似文献   

14.
The stability of a specific complex between the peptide Ag representing residue 323-339 of OVA and the MHC class II protein, I-Ad, in a lipid bilayer was investigated as a function of pH and temperature. The complex is much more stable in a lipid bilayer than previously reported for detergent micelles. Measureable dissociation was detectable only after several hours at a pH below 5. The results show that a purified preparation of MHC class II molecules can sequentially bind, release, and rebind peptide, indicating that, in principle, MHC class II molecules could be used more than once for peptide binding. However, the time and pH required for peptide-MHC dissociation suggests that, in an Ag presenting cell, either a prolonged residence in an acidic compartment or other factors will be required for regeneration of the peptide binding site.  相似文献   

15.
The binding of peptides to MHC class II molecules is mediated in part by a conserved array of intermolecular hydrogen bonds. We have evaluated the consequences of disrupting the hydrogen bond between beta-His-81 of the class II molecule and bound peptide. These studies revealed that peptide dissociation rates were accelerated by factors ranging to 200-fold. The sensitivity of a peptide to loss of the hydrogen bond is inversely correlated with the inherent kinetic stability of the peptide-MHC complex. The same relationship has been observed between inherent kinetic stability and the susceptibility to DM. Given that the rate enhancement observed for MHC class II I-Ad protein mutated at position 81 in the beta-chain is comparable with DM-catalyzed rates for other class II molecules, we suggest that DM could function by stabilizing a peptide-MHC intermediate in which one or more hydrogen bonds between the peptide and MHC, such as that contributed by the beta-His-81 hydrogen bond, are disrupted.  相似文献   

16.
Class I major histocompatibility complex (MHC) molecules bind peptides derived from degraded proteins for display to T cells of the immune system. Peptides bind to MHC proteins with varying affinities, depending upon their sequence and length. We demonstrate that the thermal stability of the MHC-peptide complex depends directly on peptide binding affinity. We use this correlation to develop a convenient method to determine peptide dissociation constants by measuring MHC-peptide complex stability using thermal denaturation profiles monitored by circular dichroism.  相似文献   

17.
MHC class II molecules associate with peptides through pocket interactions and the formation of hydrogen bonds. The current paradigm suggests that the interaction of side chains of the peptide with pockets in the class II molecule is responsible for the formation of stable class II-peptide complexes. However, recent evidence has shown that the formation of hydrogen bonds between genetically conserved residues of the class II molecule and the main chain of the peptide contributes profoundly to peptide stability. In this study, we have used I-A(k), a class II molecule known to form strong pocket interactions with bound peptides, to probe the general importance of hydrogen bond integrity in peptide acquisition. Our studies have revealed that abolishing hydrogen bonds contributed by positions 81 or 82 in the beta-chain of I-A(k) results in class II molecules that are internally degraded when trafficked through proteolytic endosomal compartments. The presence of high-affinity peptides derived from either endogenous or exogenous sources protects the hydrogen bond-deficient variant from intracellular degradation. Together, these data indicate that disruption of the potential to form a complete hydrogen bond network between MHC class II molecules and bound peptides greatly diminishes the ability of class II molecules to bind peptides. The subsequent failure to stably acquire peptides leads to protease sensitivity of empty class II molecules, and thus to proteolytic degradation before export to the surface of APCs.  相似文献   

18.
Formation of stable class II MHC/peptide complex involves conformational changes and proceeds via an intermediate. Although this intermediate complex forms and dissociates in minutes, its conversion to a stable complex is a very slow process, taking up to a few days to reach completion. Here, we investigate the different steps of this binding and demonstrate that the conformational changes necessary to generate a receptive molecule is the rate-determining slow step in the process, while formation of the stable MHC/peptide complex is very rapid. With HLA-DR1 as our model class II molecule, we first used low affinity variants of hemagglutinin peptide (HA306-318), which lack the principal anchor, to shape the conformation of the MHC and then studied the kinetics of stable binding of HA306-318 to such an induced conformation. We found that the apparent association rate of HA306-318 is equivalent to the dissociation rate of the low affinity peptide. A 4- to 18-fold enhancement in the binding rates of HA306-318 was observed depending on the dissociation rates of the low affinity peptides. These results establish that 1) formation of stable MHC/peptide complexes is very rapid and 2) prior binding of low affinity peptide induces a receptive conformation in MHC for efficient stable peptide binding. Furthermore, in the absence of any free peptide, this receptive molecule rapidly reverts to slow binding behavior toward the subsequently offered peptide. These results have important implications for the roles of low affinity MHC/peptide complexes in Ag presentation.  相似文献   

19.
Tapasin plays an important role in the quality control of major histocompatibility complex (MHC) class I assembly, but its precise function in this process remains controversial. Whether tapasin participates in the assembly of HLA-G has not been studied. HLA-G, an MHC class Ib molecule that binds a more restricted set of peptides than class Ia molecules, is a particularly interesting molecule, because during assembly, it recycles between the endoplasmic reticulum (ER) and the cis-Golgi until it is loaded with a high affinity peptide. We have taken advantage of this unusual trafficking property of HLA-G and its requirement for high affinity peptides to demonstrate that a critical function of tapasin is to transform class I molecules into a high affinity, peptide-receptive form. In the absence of tapasin, HLA-G molecules cannot bind high affinity peptides, and an abundant supply of peptides cannot overcome the tapasin requirement for high affinity peptide loading. The addition of tapasin renders HLA-G molecules capable of loading high affinity peptides and of transporting to the surface, suggesting that tapasin is a prerequisite for the binding of high-affinity ligands. Interestingly, the "tapasin-dependent" HLA-G molecules are not empty in the absence of tapasin but are in fact associated with suboptimal peptides and continue to recycle between the ER and the cis-Golgi. Together with the finding that empty HLA-G heterodimers are strictly retained in the ER and degraded, our data suggest that MHC class I molecules bind any available peptides to avoid ER-mediated degradation and that the peptides are in turn replaced by higher affinity peptides with the aid of tapasin.  相似文献   

20.
During maturation of MHC II molecules, newly synthesized and assembled complexes of MHC II alphabeta dimers with invariant chain (Ii) are targeted to endosomes, where Ii is proteolyzed, leaving remnant class II-associated Ii peptides (CLIP) in the MHC II peptide binding groove. CLIP must be released, usually with assistance from the endosomal MHC II peptide exchange factor, HLA-DM, before MHC II molecules can bind endosomal peptides. Structural factors that control rates of CLIP release remain poorly understood, although peptide side chain-MHC II specificity pocket interactions and MHC II polymorphism are important. Here we report that mutations betaS11F, betaS13Y, betaQ70R, betaK71E, betaK71N, and betaR74Q, which map to the P4 and P6 pockets of the groove of HLA-DR3 molecules, as well as alphaG20E adjacent to the groove, are associated with elevated CLIP in cells. Most of these mutations increase the resistance of CLIP-DR3 complexes to dissociation by SDS. In vitro, the groove mutations increase the stability of CLIP-DR3 complexes to dissociation. Dissociation rates in the presence of DM, as well as coimmunoprecipitation of some mutant DR3 molecules with DM, are also diminished. The profound phenotypes associated with some of these point mutations suggest that the need to maintain efficient CLIP release represents a constraint on naturally occurring MHC II polymorphism.  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