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 109 毫秒
1.
A peptide hydropathically complementary to Big Endothelin [Big ET] residues 16-29 has been synthesized in a multimeric form starting from an octadentate polylysine core, essentially in a way similar to the procedure used for the production of multiple antigenic peptides [MAP's]. Interaction between the multimeric complementary peptide [8 delta ET] and the Big ET fragment 16-32 containing the target complementary region, also synthesized in a multimeric form [8ET], was evaluated by analytical high performance affinity chromatography and solid phase binding assays. While the binding interaction between the monomerics peptide pair was in the micromolar range, the recognition between the corresponding multimeric form was characterized by enhanced binding affinity of at least two orders of magnitude. In solution, complex formation between multimeric complementary peptide and target Big ET sequence in the monomeric and multimeric form was accompanied by precipitation at concentrations higher than 0.5 mg/mL and 0.1 mg/mL, respectively. Polyclonal antibodies raised against the multimeric target sequence recognized multimeric and monomeric ET target sequences with binding affinities similar to binding affinities exhibited by the multimeric complementary peptide. Multimerization of hydropathically complementary peptides could provide an improved opportunity to measure and thus probe quantitative binding properties of complementary peptides.  相似文献   

2.
Different beta(1) integrins bind Arg-Gly-Asp (RGD) peptides with differing specificities, suggesting a role for residues in the alpha subunit in determining ligand specificity. Integrin alpha(5)beta(1) has been shown to bind with high affinity to peptides containing an Arg-Gly-Asp-Gly-Trp (RGDGW) sequence but with relatively low affinity to other RGD peptides. The residues within the ligand-binding pocket that determine this specificity are currently unknown. A cyclic peptide containing the RGDGW sequence was found to strongly perturb the binding of the anti-alpha(5) monoclonal antibody (mAb) 16 to alpha(5)beta(1). In contrast, RGD peptides lacking the tryptophan residue acted as weak inhibitors of mAb 16 binding. The epitope of mAb 16 has previously been localized to a region of the alpha(5) subunit that contains Ser(156)-Trp(157). Mutation of Trp(157) (but not of Ser(156) or surrounding residues) to alanine blocked recognition of mAb 16 and perturbed the high affinity binding of RGDGW-containing peptides to alpha(5)beta(1). The same mutation also abrogated recognition of the alpha(5)beta(1)-specific ligand peptide Arg-Arg-Glu-Thr-Ala-Trp-Ala (RRETAWA). Based on these findings, we propose that Trp(157) of alpha(5) participates in a hydrophobic interaction with the tryptophan residue in RGDGW, and that this interaction determines the specificity of alpha(5)beta(1) for RGDGW-containing peptides. Since the RGD sequence is recognized predominantly by amino acid residues on the beta(1) subunit, our results suggest that Trp(157) of alpha(5) must lie very close to these residues. Our findings therefore provide new insights into the structure of the ligand-binding pocket of alpha(5)beta(1).  相似文献   

3.
The extracellular domain of the 55-kDa TNF receptor (rsTNFR beta) has been expressed as a secreted protein in baculovirus-infected insect cells and Chinese hamster ovary (CHO)/dhfr- cells. A chimeric fusion protein (rsTNFR beta-h gamma 3) constructed by inserting the extracellular part of the receptor in front of the hinge region of the human IgG C gamma 3 chain has been expressed in mouse myeloma cells. The recombinant receptor proteins were purified from transfected cell culture supernatants by TNF alpha- or protein G affinity chromatography and gel filtration. In a solid phase binding assay rsTNFR beta was found to bind TNF alpha with high affinity comparable with the membrane-bound full-length receptor. The affinity for TNF beta was slightly impaired. However, the bivalent rsTNFR beta-h gamma 3 fusion protein bound both ligands with a significantly higher affinity than monovalent rsTNFR beta reflecting most likely an increased avidity of the bivalent construct. A molecular mass of about 140 kDa for both rsTNFR beta.TNF alpha and rsTNFR beta.TNF beta complexes was determined in analytical ultracentrifugation studies strongly suggesting a stoichiometry of three rsTNFR beta molecules bound to one TNF alpha or TNF beta trimer. Sedimentation velocity and quasielastic light scattering measurements indicated an extended structure for rsTNFR beta and its TNF alpha and TNF beta complexes. Multiple receptor binding sites on TNF alpha trimers could also be demonstrated by a TNF alpha-induced agglutination of Latex beads coated with the rsTNFR beta-h gamma 3 fusion protein. Both rsTNFR beta and rsTNFR beta-h gamma 3 were found to inhibit binding of TNF alpha and TNF beta to native 55- and 75-kDa TNF receptors and to prevent TNF alpha and TNF beta bioactivity in a cellular cytotoxicity assay. Concentrations of rsTNFR beta-h gamma 3 equimolar to TNF alpha were sufficient to neutralize TNF activity almost completely, whereas a 10-100-fold excess of rsTNFR beta was needed for similar inhibitory effects. In view of their potent TNF antagonizing activity, recombinant soluble TNF receptor fragments might be useful as therapeutic agents in TNF-mediated disorders.  相似文献   

4.
Cadherins are a family of calcium-dependent cell-surface proteins that are fundamental in controlling the development and maintenance of tissues. Motif B of E-cadherin seems to be a crucial calcium-binding site as single point mutations (D134A and D134K) completely inactivate its adhesion activity. We analyzed peptide models corresponding to motif B (amino acids 128-144) as well as selected mutations of this motif. Our NMR studies showed that this motif B sequence is actually an active calcium-binding region, even in the absence of the rest of the cadherin molecule. We found that the binding affinity of this motif is very sensitive to mutations. For example, our peptide P128-144 with the native calcium-binding sequence has an affinity of Kd 0.4 mM, whereas the mutants P128-144/ D134A and P128-144/D134K containing the replacement of Asp134 by Ala and Lys, have Kd values of only 1.5 and 11 mM, respectively. Removing Asp at position 134, which correlates with the loss of adhesion activity, decreases calcium-binding affinity 20-fold. Ala132, along with residues Asp134, Asp136 and Asn143, is involved in calcium binding in solution. We also demonstrated that the calcium-binding affinity can be increased 3-fold when an additional Asp is introduced at position 132. In 50% organic solvent, this binding affinity of peptide P128-144/A132D (17-mer) from E-cadherin is similar to that of peptide P72-100/C73-77-91A (29-mer) from alpha-lactalbumin.  相似文献   

5.
The segment corresponding to residues 121 to 164 of human plasma apolipoprotein A-I (apo-A-I) has been synthesized by the Merrifield solid phase method. The peptide binds to unilamellar phospholipid vesicles and to phospholipid-cholesterol mixed vesicles. The surface affinity of the peptide measured in this way indicated that the mechanism of binding is the same as that of apo A-I (144-165) and apo A-I itself. The peptide appears to be a globular monomer in a aqueous solution, with 17% alpha helix content. The peptide bound to vesicles activates lecithin:cholesterol acyltransferase: compared to apo A-I, the peptide is about 30% as efficient in the activation of cholesterol esterification and of phospholipid hydrolysis when the surface is saturated by the activator. For a variety of amphiphilic peptides and for apo A-I, the lecithin: cholesterol acyltransferase-activating ability correlates well with their alpha helix contents in 50% trifluoroethanol.  相似文献   

6.
An amphiphilic, cationic peptide composed of eight leucines and six lysines was synthesized by solid phase peptide synthesis (SPPS). The synthetic peptide was bactericidal within 10 min at concentrations as low as 3 microg ml - 1 against mid-exponential Escherichia coli O157:H7 suspended in buffer. Concentrations of 25 microg ml - 1 caused up to 7 log10 cfu ml - 1 reductions. When tested against E. coli O157:H7 grown in TSB, the peptide was bactericidal and bacteriostatic at concentrations of 50 and 25 microg ml - 1, respectively. An inhibitory effect was also observed against stationary phase cells. The synthetic peptide caused the release of u.v.-absorbing materials from the E. coli O157:H7 as well as an increase in its O.D.600 nm. Intracellular K+ and ATP depletion were also observed. These results suggest that the peptide increased the cell membrane permeability but it did not lyse the cells.  相似文献   

7.
A syngeneic monoclonal idiotypic antibody was prepared by immunizing the sequence peptide of complementary determining region-1 (CDRL-1) of 41S-2-L which is an antibody light chain capable of catalytically decomposing the antigen peptide (gp41 peptide:original antigen) as well as the intact gp41 molecule of HIV-1 envelope. The obtained idiotypic antibody, i41SL1-2, showed a high specificity to the CDRL-1 peptide. The intact i41SL1-2 and its heavy and light chains displayed apparent affinity constants to the CDRL-1 peptide of 3.6 × 109, 2.7 × 107, 1.8 × 106/M, respectively. The i41SL1-2 recognized the artificial molecule CA2, which has a more complex steric conformation than the CDRL-1, while the i41SL1-2 showed very low affinity to the original monoclonal antibody 41S-2 and its light chain 41S-2-L. However, a homologous sequence, EGG-D, with the gp41 peptide was expressed in the complementary determining region-3 (CDRH-3) of the heavy chain of i41SL1-2. Furthermore, the consensus sequence EGG was located at the important position of the CDRH-3 loop of i41SL1-2. Although the sequence of CDRL-1 (16 mer) is quite shorter than that of whole light chain (112 mer), the CDRL-1 could induce the rearrangement of CDRH-3 gene of i41SL1-2 so as to express a homologous sequence with the original antigen.  相似文献   

8.
Two peptides with hydropathic complementarity to residues 356-375 of the c-raf protein were synthesized to determine if they recognize the raf-(356-375) peptide as well as the entire protein. One peptide was deduced from the complementary mRNA for the raf protein corresponding to residues 356-375, whereas the other was deduced solely from the amino acid sequence of the 20-mer segment using a computer program able to generate peptide sequences with hydropathic complementarity to a given sequence. Specific binding of both peptides to the raf 20-mer segment was demonstrated when either the raf 20-mer peptide or the complementary peptides were immobilized on a column. Binding affinities were in the millimolar-micromolar range. Identical binding properties were observed with peptides synthesized with either all D- or all L-amino acids, suggesting a lack of conformational dependence. Binding was also unaffected by the presence of 8 M urea or detergents, was dependent on solvent characteristics of pH and ionic strength, and was abolished by the presence of competing peptides in the eluting buffer. Recognition between raf complementary peptides was accompanied by spectral changes in the far and near UV region, as monitored by circular dichroism. Proteolytic degradation was retarded by the binding of these peptides. Once immobilized on a column, these peptides proved useful for the isolation by affinity chromatography of a recombinant c-raf protein from an Escherichia coli crude cell extract.  相似文献   

9.
Molecular recognition between Big Endothelin (Big ET) and a computer generated peptide hydropathically complementary to Big ET[16-29] sequence has been studied by analytical high performance liquid affinity chromatography (HPLAC), circular dichroism (CD) and nuclear magnetic resonance (NMR) experiments. Specific binding was observed between solid support immobilized complementary peptide and Big ET[1-38], [1-32], and [16-32], but not with Big ET fragments [1-21], [16-21], [22-32], and [22-38], obtained by chymotrypsin proteolytic degradation. Selectivity in the recognition process was clearly demonstrated by the ability of complementary peptide affinity column to purify the Big ET molecule from complex peptide mixtures, even when present in very low concentrations. Similar selectivity was evidenced with the Big ET fragment [16-32], [NH2-HLDIIWVNTPEHIVPYG-COOH] containing the entire hydropathically complementary sequence. Binding was followed by marked spectroscopic changes, as monitored by circular dichroism and one- and two-dimensional nuclear magnetic resonance experiments. The NMR spectra of the complementary peptides 1:1 mixture showed variations in the chemical shifts of proton resonances in several residues, both in the main chain (amide protons) and in the side chains (aliphatic and aromatic protons). These data support the hypothesis of a multilocalized type of interaction between complementary peptides, where many residues along the peptide chains participate in co-operative stabilizing contacts in the forming complex.  相似文献   

10.
Structural determinants of L-[(3)H]nicotine binding to synthetic peptides comprising residues 188-207 of nicotinic acetylcholine receptor alpha subunits were invesitigated by equilibrium binding analysis. Two binding components were detected, one of low affinity (K(d) approximately 1.5 microM) that did not differ significantly among peptides and another of high affinity. The high affinity binding component was higher for the neuronal peptides (K(d) = 14-23 nM) than the muscle alpha1 peptides (K(d) = 52 nM). The following nonconservative substitutions in the alpha4 peptide resulted in a significant decrease in nicotine affinity for the peptide: Y190A, Y190D, C192G, E195A, E195-, P199A, P199-, and Y203A. Substitution of alpha4P199 with a leucine which is present in the alpha1 sequence decreased the affinity of the alpha4 peptide for nicotine and substitution of alpha1L199 with a proline (alpha4) or a glutamine (alpha3) increased the affinity of the alpha1 peptide. It is concluded that aromatic residues contribute to the binding site for nicotine on the alpha4 subunit and that the residue present at position 199 partly determines differences in nicotine affinity for different alpha subunits.  相似文献   

11.
To identify the sequence segments of the alpha 3 subunit of the neuronal nicotinic acetylcholine receptor (N-nAChR) forming the binding site for the cholinergic antagonist kappa-bungarotoxin (kappa-BGT), overlapping peptides corresponding to the complete alpha 3 sequence were tested for their ability to bind 125I-labeled kappa-BGT. Two peptides located within the N-terminal extracellular domain specifically bound kappa-BGT in a solid phase assay, i.e. peptide N alpha(3)51-70 with a Kd approximately 300 nM and peptide N alpha(3)1-18 with slightly lower affinity (Kd approximately 500 nM). Preincubation of 125I-kappa-BGT with peptides N alpha(3)51-70 or N alpha(3)1-18 resulted in greater than 90% inhibition of kappa-125I-BGT binding to native N-nAChR expressed on the neuronal cell line PC12. Under the same conditions, two additional peptides, N alpha(3)180-199 and N alpha(3)183-201, were found to inhibit kappa-125I-BGT binding to PC12 by approximately 50%. These latter peptides represent sequences that are homologous to those shown previously to bind alpha-bungarotoxin. Peptide N alpha(3)51-70 (400 microM) also reduced by approximately 4-fold the observed rate of association of kappa-BGT to PC12 cells. The results of these experiments identify sequence segments of the alpha 3 subunit which are likely to interact with kappa-BGT and may indicate the relative contribution that these segments make in the formation of the high affinity kappa-BGT-binding site of this N-nAChR subtype.  相似文献   

12.
Synthetic overlapping peptides of the alpha-subunit of human chorionic gonadotropin (hCG) were made by solid-phase peptide synthesis employing a comprehensive synthetic approach. The entire primary structure of the alpha-subunit was synthesized as a series of nine consecutive peptides, each 15 residues in length, and overlapping with its two adjacent neighbors by 5 residues on each side. Receptor binding activity of each synthetic peptide was measured by the inhibition of binding of 125I-labeled hCG to rat ovarian receptor. Peptides alpha 21-35, alpha 31-45, alpha 71-85, and alpha 81-92 were shown to compete for binding with native hCG, thus demonstrating that at least two regions on the alpha-subunit may be part of the binding site(s) of the hormone. The low affinity of the peptides (10(-5)-10(-6) M) compared to native hormone (10(-10) M) for receptor is not unexpected due to the probability of discontinuous and multiple sites involved in receptor binding. An ultrapure preparation of hCG alpha-subunit also had low affinity (10(-5), suggesting that conformational changes upon combination with beta-subunit to form dimer or changes in conformation after binding are necessary for high affinity interaction. These results correlate with previous predictions of binding sites based on studies employing chemical and enzymatic modifications of intact hormone and show that synthetic peptide strategies are helpful in the elucidation of protein structure and function.  相似文献   

13.
Design, synthesis and DNA binding activity of a nonlinear 102 residue peptide are reported. The peptide contains four sequence-specific DNA binding domains of 434 Cro protein. These four domains were linked covalently to a symmetrical carboxyterminal crosslinker that contains four arms each ending with an aliphatic aminogroup. From CD studies we have found that in aqueous buffer in the presence of 20% trifluoroethanol the peptide residues assume alpha helical, beta-sheet and random coiled conformations with an alpha helical content of about 16% at room temperature. The alpha helicity is increased up to 40% in the presence of 40% trifluoroethanol. Upon complex formation between the peptide and DNA a change in the peptide conformation takes place which is consistent with an alpha-beta transition in the DNA binding, helix-turn-helix motif of 434 Cro repressor. Evidently residues present in helices alpha(2) and alpha(3) form a beta hairpin which is inserted in the minor DNA groove. The latter inference is supported by our observations that the peptide can displace minor groove binding antibiotic distamycin A from a complex with poly(dA).poly(dT). As revealed from DNase protection studies the peptide exhibits preferences for binding to operator and pseudooperator sites recognized by 434 Cro repressor. It binds strongly to operator sites OR1, OR2 and OR3 and exhibits a greater affinity for pseudooperator site Op1. From analysis of nucleotide sequences in the strong affinity binding sites for the peptide on DNA a conclusion is drawn that it binds to pseudosymmetrical nucleotide sequences 5'-ACAA(W)nCTGT-3', where W is an arbitrary nucleotide. n is equal to six or seven. In the strongest affinity binding site for the peptide on DNA (Op1) motif 5'-ACAA-3' is replaced by sequence 5'-ACCA-3'. A difference in binding specificity shown by the peptide and 434 Cro protein could be attributed to a flexibility of the connecting chains between DNA-binding domains in the peptide molecule as well as to a replacement of Thr - Ala in the alpha 2 helix. Removal of two residues from the N-terminal end of helix alpha 2 in each of the four DNA binding domains of 434 Cro present in the peptide leads to a loss of binding specificity, although the modified peptide binds to DNA unspecifically.  相似文献   

14.
The effect of bacterial lipopolysaccharide (LPS) on macrophage receptors for tumor necrosis factor/cachectin (TNF-R) was studied. At equilibrium, iodinated recombinant human TNF alpha (rTNF alpha) bound to 1100 +/- 200 sites/cell on macrophage-like RAW 264.7 cells with a Kd of 1.3 +/- 0.1 x 10(-9) M. Preexposure of RAW 264.7 cells to 10 ng/ml LPS for 1 h at 37 degrees C resulted in complete loss of cell surface TNF alpha binding sites. 50% loss ensued after 1 h with 0.6 ng/ml LPS, or after 15 min with 10 ng/ml LPS. Complete loss of TNF alpha binding sites occurred without change in numbers of complement receptor type 3. No decrease in TNF-R followed preexposure to LPS at 4 degrees C, nor could LPS displace 125I-rTNF alpha from its binding sites. Although TNF-R disappeared from the surface of intact macrophages following exposure to LPS, specific TNF alpha binding sites were unchanged in permeabilized macrophages, indicating that TNF-R were rapidly internalized. Conditioned media from LPS-treated RAW 264.7 cells induced 30% down-regulation of TNF-R on macrophages from LPS-hyporesponsive mice (C3H/HeJ), suggesting that a soluble macrophage product may be responsible for a minor portion of the LPS effect. Additional evidence against endogenous TNF alpha being the major cause of TNF-R internalization was the rapid onset of the effect of LPS on TNF-R compared to the reported onset of TNF alpha production, the relatively high concentrations of exogenous rTNF alpha required to mimic the effect of LPS, and the inability of TNF alpha-neutralizing antibody to block the effect of LPS. LPS-induced down-regulation of TNF-R was complete or nearly complete not only in RAW 264.7 cells, but also in primary macrophages of both human and murine origin, was less marked in human endothelial cells, and was absent in human granulocytes and melanoma cells and mouse L929 cells. Thus, in situ, macrophages and some other host cells may be resistant to the actions of TNF alpha produced during endotoxinemia, because such cells may internalize their TNF-R in response to LPS before TNF alpha is produced.  相似文献   

15.
Polyclonal antibodies were raised in rabbits against the GABAA-receptor beta 3 subunit peptide sequence, KQSMPREGHGRHMDR-NH2 coupled to keyhole limpet haemocyanin. These anti-beta 3 379-393 antibodies immunoprecipitated in a dose-dependent manner specific benzodiazepine agonist binding sites from Na+ deoxycholate extracts of bovine cerebral cortex. In immunoblots, anti-beta 3 379-393 antibodies recognised two species with Mr 59,900 and Mr 57,200 in all preparations tested, which included crude detergent-solubilised, benzodiazepine affinity chromatography-purified receptor, anti-alpha 1 324-341 antibody, anti-Cys alpha 2 414-424 antibody and anti-Cys alpha 3 454-467 antibody immunoaffinity-purified GABAA-receptor subpopulations. These results provide evidence for the ubiquity and promiscuity of the GABAA-receptor beta 3 subunit.  相似文献   

16.
Previously identified high affinity integrin-binding motifs in collagens, GFOGER and GLOGER, are not present in type III collagen. Here, we first characterized the binding of recombinant I domains from integrins alpha(1) and alpha(2) (alpha(1)I and alpha(2)I) to fibrillar collagen types I-III and showed that each I domain bound to the three types of collagens with similar affinities. Using rotary shadowing followed by electron microscopy, we identified a high affinity binding region in human type III collagen recognized by alpha(1)I and alpha(2)I. Examination of the region revealed the presence of two sequences that contain the critical GER motif, GROGER and GAOGER. Collagen-like peptides containing these two motifs were synthesized, and their triple helical nature was confirmed by circular dichroism spectroscopy. Experiments show that the GROGER-containing peptide was able to bind both alpha(1)I and alpha(2)I with high affinity and effectively inhibit the binding of alpha(1)I and alpha(2)I to type III and I collagens, whereas the GAOGER-containing peptide was considerably less effective. Furthermore, the GROGER-containing peptide supported adhesion of human lung fibroblast cells when coated on a culture dish. Thus, we have identified a novel high affinity binding sequence for the collagen-binding integrin I domains.  相似文献   

17.
The sequence segment 181-200 of the Torpedo nicotinic acetylcholine receptor (nAChR) alpha subunit forms a binding site for alpha-bungarotoxin (alpha-BTX) [e.g., see Conti-Tronconi, B. M., Tang, F., Diethelm, B. M., Spencer, S. R., Reinhardt-Maelicke, S., & Maelicke, A. (1990) Biochemistry 29, 6221-6230]. Synthetic peptides corresponding to the homologous sequences of human, calf, mouse, chicken, frog, and cobra muscle nAChR alpha 1 subunits were tested for their ability to bind 125I-alpha-BTX, and differences in alpha-BTX affinity were determined by using solution (IC50S) and solid-phase (KdS) assays. Panels of overlapping peptides corresponding to the complete alpha 1 subunit of mouse and human were also tested for alpha-BTX binding, but other sequence segments forming the alpha-BTX site were not consistently detectable. The Torpedo alpha 1(181-200) and the homologous frog and chicken peptides bound alpha-BTX with higher affinity (KdS approximately 1-2 microM, IC50s approximately 1-2 microM) than the human and calf peptides (Kds approximately 3-5 microM, IC50s approximately 15 microM). The mouse peptide bound alpha-BTX weakly when attached to a solid support (Kd approximately 8 microM) but was effective in competing for 125I-alpha-BTX in solution (IC50 approximately 1 microM). The cobra nAChR alpha 1-subunit peptide did not detectably bind alpha-BTX in either assay. Amino acid substitutions were correlated with alpha-BTX binding activity peptides from different species. The role of a putative vicinal disulfide bound between Cys-192 and -193, relative to the Torpedo sequence, was determined by modifying the peptides with sulfhydryl reagents. Reduction and alkylation of the peptides decreased alpha-BTX binding, whereas oxidation of the peptides had little effect. Modifications of the cysteine/cystine residues of the cobra peptide failed to induce alpha-BTX binding activity. These results indicate that while the adjacent cysteines are likely to be involved in forming the toxin/alpha 1-subunit interface a vicinal disulfide bound was not required for alpha-BTX binding.  相似文献   

18.
The integrin alpha(v)beta3 has been shown to act as the receptor for internalization of foot-and-mouth disease virus (FMDV) (A12), with attachment being through a highly conserved RGD motif located on the G-H loop of viral capsid protein VP1. In addition, however, we have recently shown that efficient infection of culture-grown cells by FMDV (O1BFS) requires binding to cell surface heparan sulfate. In this study, we have used a solid-phase receptor binding assay to characterize the binding by FMDV to purified alpha(v)beta3 in the absence of heparan sulfate and other cell surface components. In this assay, FMDV (O1BFS) successfully replicated authentic ligand binding by cellular alpha(v)beta3 in terms of its high affinity, dependence on divalent cations, and activation by manganese ions. Virus binding to this preparation of alpha(v)beta3 was exquisitely sensitive to competition by short RGD-containing peptides (50% inhibition at < 10(-8) M peptide), and this inhibition was highly sequence specific, with the equivalent RGE peptide being at least 10(4) fold less effective as a competitor. Representative viruses of the other six serotypes of FMDV bound to alpha(v)beta3 in a similar RGD-specific manner, although significant differences in sensitivity to RGD peptides suggest that the affinity of the different FMDV serotypes for alpha(v)beta3 is influenced, in part, by the variable amino acid residues in the VP1 G-H loop on either side of the RGD.  相似文献   

19.
Talts JF  Timpl R 《FEBS letters》1999,458(3):319-323
A RRKRRQ sequence unique to the LG3 module of the laminin alpha2 chain was previously shown to be sensitive to endogenous proteolysis during the recombinant production of the tandem array alpha2LG1-3. Mutation of RQ surrounding the cleaved peptide bond did not prevent this processing and intracellular degradation. Alanine mutagenesis of three alternate basic residues, however, was shown to prevent the cleavage in alpha2LG1-3, allowing for the alpha2LG3 module to be obtained as a folded, globular fragment. The mutation did not change heparin and sulfatide binding or cell adhesion of alpha2LG1-3 which can be mediated by alpha3beta1 and alpha6beta1 integrins. It did, however, cause a 10-fold reduction in alpha-dystroglycan binding. The data favor the interpretation that binding epitopes for heparin/sulfatides, beta1 integrins and alpha-dystroglycan occupy different parts of the alpha2LG1-3 structure.  相似文献   

20.
The solution NMR structure is reported for Ca(2+)-loaded S100B bound to a 12-residue peptide, TRTK-12, from the actin capping protein CapZ (alpha1 or alpha2 subunit, residues 265-276: TRTKIDWNKILS). This peptide was discovered by Dimlich and co-workers by screening a bacteriophage random peptide display library, and it matches exactly the consensus S100B binding sequence ((K/R)(L/I)XWXXIL). As with other S100B target proteins, a calcium-dependent conformational change in S100B is required for TRTK-12 binding. The TRTK-12 peptide is an amphipathic helix (residues W7 to S12) in the S100B-TRTK complex, and helix 4 of S100B is extended by three or four residues upon peptide binding. However, helical TRTK-12 in the S100B-peptide complex is uniquely oriented when compared to the three-dimensional structures of other S100-peptide complexes. The three-dimensional structure of the S100B-TRTK peptide complex illustrates that residues in the S100B binding consensus sequence (K4, I5, W7, I10, L11) are all involved in the S100B-peptide interface, which can explain its orientation in the S100B binding pocket and its relatively high binding affinity. A comparison of the S100B-TRTK peptide structure to the structures of apo- and Ca(2+)-bound S100B illustrates that the binding site of TRTK-12 is buried in apo-S100B, but is exposed in Ca(2+)-bound S100B as necessary to bind the TRTK-12 peptide.  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