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 829 毫秒
1.
The structures of concanavalin A (ConA) in complex with two carbohydrate-mimicking peptides, 10-mer (MYWYPYASGS) and 15-mer (RVWYPYGSYLTASGS) have been determined at 2.75 A resolution. In both crystal structures four independent peptide molecules bind to each of the crystallographically independent subunits of ConA tetramer. The peptides exhibit small but significant variability in conformations and interactions while binding to ConA. The crystal structure of another similar peptide, 12-mer (DVFYPYPYASGS), in complex with ConA has been determined (Jain, D., K. J. Kaur, B. Sundaravadivel, and D. M. Salunke. 2000. Structural and functional consequences of peptide-carbohydrate mimicry. J. Biol. Chem. 275:16098-16102). Comparison of the three complexes shows that the peptides bind to ConA at a common binding site, using different contacting residues and interactions depending on their sequence and the local environment at the binding site. The binding is also optimized by corresponding plasticity of the peptide binding site on ConA. The diversity in conformation and interactions observed here are in agreement with the structural leeway concerning plasticity of specific molecular recognition in biological processes. The adaptability of peptide-ConA interactions may also be correlated with the carbohydrate-mimicking property of these peptides.  相似文献   

2.
The preceding paper [Dam, T. K., Roy, R., Pagé, D., and Brewer, C. F. (2002) Biochemistry 41, 1351-1358] demonstrated that Hill plots of isothermal titration microcalorimetry (ITC) data for the binding of di-, tri-, and tetravalent carbohydrate analogues possessing terminal 3,6-di-O-(alpha-D-mannopyranosyl)-alpha-D-mannopyranoside residues to the lectin concanavalin A (ConA) show increasing negative cooperativity upon binding of the analogues to the lectin. The present study demonstrates "reverse" ITC experiments in which the lectin is titrated into solutions of di- and trivalent analogues. The results provide direct determinations of the thermodynamics of binding of ConA to the individual epitopes of the two multivalent analogues. The n values (number of binding sites per carbohydrate molecule) derived from reverse ITC demonstrate two functional binding epitopes on both the di- and trivalent analogues, confirming previous "normal" ITC results with the two carbohydrates [Dam, T. K., Roy, R., Das, S. K., Oscarson, S., and Brewer, C. F. (2000) J. Biol. Chem. 275, 14223-14230]. The reverse ITC measurements show an 18-fold greater microscopic affinity constant of ConA for the first epitope of the divalent analogue versus its second epitope and a 53-fold greater microscopic affinity constant of ConA binding to the first epitope of the trivalent analogue versus its second epitope. The data also demonstrate that the microscopic enthalpies of binding of the two epitopes on the di- and trivalent analogues are essentially the same and that differences in the microscopic K(a) values of the epitopes are due to their different microscopic entropies of binding values. These findings are consistent with the increasing negative Hill coefficients of these analogues binding to ConA in the previous paper.  相似文献   

3.
Hepatocyte growth factor (HGF) is a heparin-binding, multipotent growth factor that transduces a wide range of biological signals, including mitogenesis, motogenesis, and morphogenesis. Heparin or closely related heparan sulfate has profound effects on HGF signaling. A heparin-binding site in the N-terminal (N) domain of HGF was proposed on the basis of the clustering of surface positive charges [Zhou, H., Mazzulla, M. J., Kaufman, J. D., Stahl, S. J., Wingfield, P. T., Rubin, J. S., Bottaro, D. P., and Byrd, R. A. (1998) Structure 6, 109-116]. In the present study, we confirmed this binding site in a heparin titration experiment monitored by nuclear magnetic resonance spectroscopy, and we estimated the apparent dissociation constant (K(d)) of the heparin-protein complex by NMR and fluorescence techniques. The primary heparin-binding site is composed of Lys60, Lys62, and Arg73, with additional contributions from the adjacent Arg76, Lys78, and N-terminal basic residues. The K(d) of binding is in the micromolar range. A heparin disaccharide analogue, sucrose octasulfate, binds with similar affinity to the N domain and to a naturally occurring HGF isoform, NK1, at nearly the same region as in heparin binding. (15)N relaxation data indicate structural flexibility on a microsecond-to-millisecond time scale around the primary binding site in the N domain. This flexibility appears to be dramatically reduced by ligand binding. On the basis of the NK1 crystal structure, we propose a model in which heparin binds to the two primary binding sites and the N-terminal regions of the N domains and stabilizes an NK1 dimer.  相似文献   

4.
Concanavalin A, which binds to specific carbohydrate determinants on the cell surface, was used to investigate the binding of prolactin to its receptors in liver membranes from female rats. The binding of 125I-labeled ovine prolactin to receptors was sharply inhibited by concanavalin A. This effect was reversed by the competitive sugar α-methyl-D-mannopyranoside and thus required the presence of specifically bound lectin. Concentrations of concanavalin A of up to 50 μg/ml caused a progressive decrease in the apparent affinity of the prolactin receptor for hormone. When higher concentrations were used, the number of available binding sites decreased. Concanavalin A-resistant receptors, about 30% of the total, had the same dissociation constant (Kd) as the controls. The binding of 125I-labeled concanavalin A in the same membrane preparations showed the presence of two distinct types of concanavalin A binding. At low concentrations, the lectin bound with high affinity (Kd ≈ 6.6 · 10?8 M). At high lectin concentrations, low affinity (Kd ≈ 6.7 · 10?5 M) binding predominated. Since high affinity concanavalin A binding was saturated at 50 μg/ml, this class of binding most likely alters the affinity of the prolactin receptor for hormone; low affinity concanavalin A binding may mask prolactin receptors, making them inaccessible to the hormone.Binding sites for concanavalin A and prolactin appear to be independent but closely related since (i) concanavalin A did not displace bound prolactin from its receptor, and (ii) detergent-solubilized 125I-labeled prolactin-receptor complexes bound to concanavalin A-Sepharose and were eluted by α-methyl-D-mannopyranoside.  相似文献   

5.
Surfactant proteins A and D bind CD14 by different mechanisms   总被引:13,自引:0,他引:13  
Surfactant proteins A (SP-A) and D (SP-D) are lung collectins that are constituents of the innate immune system of the lung. Recent evidence (Sano, H., Sohma, H., Muta, T., Nomura, S., Voelker, D. R., and Kuroki, Y. (1999) J. Immunol. 163, 387-395) demonstrates that SP-A modulates lipopolysaccharide (LPS)-induced cellular responses by direct interaction with CD14. In this report we examined the structural elements of the lung collectins involved in CD14 recognition and the consequences for CD14/LPS interaction. Rat SP-A and SP-D bound CD14 in a concentration-dependent manner. Mannose and EDTA inhibited SP-D binding to CD14 but did not decrease SP-A binding. The SP-A binding to CD14 was completely blocked by a monoclonal antibody that binds to the SP-A neck domain but only partially blocked by an antibody that binds to the SP-A lectin domain. SP-A but not SP-D bound to deglycosylated CD14. SP-D decreased CD14 binding to both smooth and rough LPS, whereas SP-A enhanced CD14 binding to rough LPS and inhibited binding to smooth LPS. SP-A also altered the migration profile of LPS on a sucrose density gradient in the presence of CD14. From these results, we conclude that 1) lung collectins bind CD14, 2) the SP-A neck domain and SP-D lectin domain participate in CD14 binding, 3) SP-A recognizes a peptide component and SP-D recognizes a carbohydrate moiety of CD14, and 4) lung collectins alter LPS/CD14 interactions.  相似文献   

6.
The neonatal Fc receptor, FcRn, is responsible for the long half-life of IgG molecules in vivo and is a potential therapeutic target for the treatment of autoimmune diseases. A family of peptides comprising the consensus motif GHFGGXY, where X is preferably a hydrophobic amino acid, was shown previously to inhibit the human IgG:human FcRn protein-protein interaction (Mezo, A. R., McDonnell, K. A., Tan Hehir, C. A., Low, S. C., Palombella, V. J., Stattel, J. M., Kamphaus, G. D., Fraley, C., Zhang, Y., Dumont, J. A., and Bitonti, A. J. (2008) Proc. Natl. Acad. Sci. U.S.A., 105, 2337–2342). Herein, the x-ray crystal structure of a representative monomeric peptide in complex with human FcRn was solved to 2.6 Å resolution. The structure shows that the peptide binds to human FcRn at the same general binding site as does the Fc domain of IgG. The data correlate well with structure-activity relationship data relating to how the peptide family binds to human FcRn. In addition, the x-ray crystal structure of a representative dimeric peptide in complex with human FcRn shows how the bivalent ligand can bridge two FcRn molecules, which may be relevant to the mechanism by which the dimeric peptides inhibit FcRn and increase IgG catabolism in vivo. Modeling of the peptide:FcRn structure as compared with available structural data on Fc and FcRn suggest that the His-6 and Phe-7 (peptide) partially mimic the interaction of His-310 and Ile-253 (Fc) in binding to FcRn, but using a different backbone topology.  相似文献   

7.
G-proteins cycle between an inactive GDP-bound state and an active GTP-bound state, serving as molecular switches that coordinate cellular signaling. We recently used phage display to identify a series of peptides that bind G alpha subunits in a nucleotide-dependent manner [Johnston, C. A., Willard, F. S., Jezyk, M. R., Fredericks, Z., Bodor, E. T., Jones, M. B., Blaesius, R., Watts, V. J., Harden, T. K., Sondek, J., Ramer, J. K., and Siderovski, D. P. (2005) Structure 13, 1069-1080]. Here we describe the structural features and functions of KB-1753, a peptide that binds selectively to GDP x AlF4(-)- and GTPgammaS-bound states of G alpha(i) subunits. KB-1753 blocks interaction of G alpha(transducin) with its effector, cGMP phosphodiesterase, and inhibits transducin-mediated activation of cGMP degradation. Additionally, KB-1753 interferes with RGS protein binding and resultant GAP activity. A fluorescent KB-1753 variant was found to act as a sensor for activated G alpha in vitro. The crystal structure of KB-1753 bound to G alpha(i1) x GDP x AlF4(-) reveals binding to a conserved hydrophobic groove between switch II and alpha3 helices and, along with supporting biochemical data and previous structural analyses, supports the notion that this is the site of effector interactions for G alpha(i) subunits.  相似文献   

8.
Gómez I  Dean DH  Bravo A  Soberón M 《Biochemistry》2003,42(35):10482-10489
The identification of epitopes involved in Cry toxin-receptor interaction could provide insights into the molecular basis of insect specificity and for designing new toxins to overcome the potential problem of insect resistance. In previous works, we determined that the Manduca sexta Cry1A cadherin-like receptor (Bt-R(1)) interacts with Cry1A toxins through epitope (865)NITIHITDTNN(875) and by loop 2 of domain II in the toxin (Gomez, I., Miranda-Rios, J., Rudi?o-Pi?era, E., Oltean, D. I., Gill, S. S., Bravo, A., and Soberón, M. (2002) J. Biol. Chem. 277, 30137-30143.). In this work, we narrowed to 12 amino acids a previously identified Bt-R(1) 66 amino acids epitope (Dorsch, J. A., Candas, M., Griko, N. B., Maaty, W. S. A., Midbo, E. G., Vadlamudi, R. K., and Bulla, L. A., Jr. (2002) Insect Biochem. Mol. Biol. 32, 1025-1036) and identified loop alpha-8 of Cry1Ab domain II as its cognate binding epitope. Two amino acid Bt-R(1) toxin binding regions of 70 residues, one comprised of residues 831-900 containing the (865)NITIHITDTNN(875) epitope (TBR1) and the other comprised of residues 1291-1360 (TBR2) were cloned by RT-PCR and produced in Escherichia coli. Cry1A toxins bind with the two TBR regions in contrast with the nontoxic Cry3A toxin. The loop 2 synthetic peptide competed with the binding of Cry1Ab toxin to both TBR regions in contrast to the alpha-8 synthetic peptide that only competed with Cry1Ab binding to TBR2. Western blots and competition ELISA analysis showed that the Cry1Ab loop 2 RR368-9EE mutant did not show observable binding to TBR1 but still bound the TBR2 peptide. This result suggests that loop alpha-8 interacts with the TBR2 region. Competition ELISA analysis of Cry1Ab binding to the two TBR peptides revealed that the toxin binds the TBR1 region with 6-fold higher affinity than the TBR2 region. The amino acid sequence of TBR2 involved on Cry1Ab interaction was narrowed to 12 amino acids, (1331)IPLPASILTVTV(1342), by using synthetic peptides as competitors for Cry1Ab binding to Bt-R(1). Our results show that the specificity of Cry1A involves at least two structural determinants on both molecules.  相似文献   

9.
Concanavalin A, which binds to specific carbohydrate determinants on the cell surface, was used to investigate the binding of prolactin to its receptors in liver membranes from female rats. The binding of 125I-labeled ovine prolactin to receptors was sharply inhibited by concanavalin A. This effect was reversed by the competitive sugar alpha-methyl-D-mannopyranoside and thus required the presence of specifically bound lectin. Concentrations of concanavalin A of up to 50 mu/ml caused a progressive decrease in the apparent affinity of the prolactin receptor for hormone. When higher concentrations were used, the number of available binding sites decreased. Concanavalin A-resistant receptors, about 30% of the total, had the same dissociation constant (Kd) as the controls. The binding of 125I-labeled concanavalin A in the same membrane preparations showed the presence of two distinct types of concanavalin A binding. At low concentrations, the lectin bound with high affinity (Kd approximately equal to 6.6 . 10(-8) M. At high lectin concentrations, low affinity (Kd approximately equal to 6.7 . 10(-5) M) binding predominated. Since high affinity concanavalin A binding was saturated at 50 microgram/ml, this class of binding most likely alters the affinity of the prolactin receptor for hormone; low affinity concanavalin A binding may mask prolactin receptors, making them inaccessible to the hormone. Binding sites for concanavalin A and prolactin appear to be independent but closely related since (i) concanavalin A did not displace bound prolactin from its receptor, and (ii) detergent-solubilized 125I-labeled prolactin-receptor complexes bound to concanavalin A-Sepharose and were eluted by alpha-methyl-D-mannopyranoside.  相似文献   

10.
The specificity of SH3 domain complex formation plays an important role in determining signal transduction events. We have previously identified a highly specific interaction between the first CrkSH3 domain [CrkSH3(1)] and proline-rich sequences in the guanine nucleotide exchange factor C3G. A 10 amino acid peptide derived from the first proline-rich sequence (P3P4P5A6L7P8P9K10K11R12) bound with a Kd of 1.89 +/- 0.06 microM and fully retained the high affinity and unique selectivity for the CrkSH3(1) domain. Mutational analysis showed that P5, P8, L7 and K10 are critical for high affinity binding. A conservative mutation, K10R, significantly decreased the affinity for the CrkSH3(1) domain while increasing the affinity for Grb2. Comparative binding studies with the K10R and K10A mutant peptides to c-Crk and v-Crk further suggested that K10 binds via a charge-dependent and a charge-independent interaction to the RT loop of the CrkSH3(1) domain. Besides determining important structural features necessary for high affinity and specificity binding to the CrkSH3(1) domain, our results also demonstrate that a conservative mutation in a single amino acid can significantly alter the specificity of an SH3 binding peptide.  相似文献   

11.
The purpose of this study was to identify the binding site(s) within laminin for the alpha 3 beta 1 integrin receptor. It has been previously shown, using proteolytic fragments and anti-laminin antibodies, that the region in laminin for alpha 3 beta 1 integrin binding is localized to the carboxy-terminal region at the end of the long arm (Gehlsen, K. R., E. Engvall, K. Dickerson, W. S. Argraves, and E. Ruoslahti. 1989. J. Biol. Chem. 264:19034-19038; Tomaselli, K. J., D. E. Hall, L. T. Reichardt, L. A. Flier, K. R. Gehlsen, D. C. Turner, and S. Carbonetto. 1990. Neuron. 5:651-662). Using synthetic peptides, we have identified an amino acid sequence within the carboxy-terminal region of the laminin A chain that is recognized by the alpha 3 beta 1 integrin. The amino acid sequence represented by the synthetic peptide GD-6 (KQNCLSSRASFRGCVRNLRLSR residues numbered 3011 to 3032) of the globular domain of the murine A chain supports cell attachment and inhibits cell adhesion to laminin-coated surfaces. By affinity chromatography, peptide GD-6-Sepharose specifically bound solubilized alpha 3 beta 1 from extracts of surface-iodinated cells in a cation-dependent manner, while it did not bind other integrins. In addition, exogenous peptide GD-6 specifically eluted bound alpha 3 beta 1 from laminin-Sepharose columns but did not elute the alpha 3 beta 1 integrin from a fibronectin-Sepharose column. Using integrin subunit-specific monoclonal antibodies, only those antibodies against the alpha 3 and beta 1 subunits inhibited cell adhesion to peptide GD-6-coated surfaces. Finally, a polyclonal antibody made against peptide GD-6 reacted specifically with both murine and human laminin and significantly inhibited cell adhesion to laminin-coated surfaces but not those coated with other matrix proteins. These results identify the laminin A chain amino acid sequence of peptide GD-6 as representing a binding site in laminin for the alpha 3 beta 1 integrin.  相似文献   

12.
Binding specificity in lactose permease toward galactopyranosides is governed by H-bonding interactions at C-2, C-3, C-4, and C-6 OH groups, while binding affinity can be increased dramatically by nonspecific hydrophobic interactions with the non-galactosyl moiety [Sahin-Tóth, M., Akhoon, K. M., Runner, J., and Kaback, H. R. (2000) Biochemistry 39, 5097-5103]. To characterize the contribution of individual hydroxyls, binding of structural analogues of p-nitrophenyl alpha-D-galactopyranoside (NPG) was examined by site-directed N-[(14)C]ethylmaleimide (NEM) labeling of the substrate-protectable Cys148 in the binding site. NPG blocks NEM alkylation of Cys148 with an apparent affinity of approximately 14 microM. A deoxy derivative at position C-2 binds with 25-fold lower affinity (K(D) 0.35 mM), and the deoxy analogue at C-3 exhibits ca. 70-fold decreased binding (K(D) 1 mM), while binding of 6-deoxy-NPG is at least 130-fold diminished (K(D) 1.9 mM). Remarkably, the C-4 deoxy derivative of NPG binds with almost 1500-fold reduced affinity (K(D) approximately 20 mM). No significant substrate protection is afforded by NPG analogues with methoxy (CH(3)-O-) substitutions at positions C-3, C-4, and C-6. In contrast, the C-2 methoxy analogue binds almost normally (K(D) 26 microM). The results confirm and extend the observations that the C-2, C-3, C-4, and C-6 OH groups of galactopyranosides participate in important H-bonding interactions. Moreover, the C-4 hydroxyl is identified as the major determinant of ligand binding, suggesting that sugar recognition in lactose permease may have evolved to discriminate primarily between gluco- and galactopyranosides.  相似文献   

13.
Carbohydrate mimetic peptides are designable, and they can carry T-cell epitopes and circumvent tolerance. A mimic-based human immunodeficiency virus (HIV) vaccine can be a viable alternative to carbohydrate-based antigens if the diversity of epitopes found on gp120 can be recapitulated. To improve existing mimics, an attempt was made to study the structural correlates of the observed polyspecificity of carbohydrate mimetic peptides based on the Y(P/R)Y motif in more detail. A carbohydrate mimetic peptide, D002 (RGGLCYCRYRYCVCVGR), bound a number of lectins with different specificities. Although this peptide reacted strongly with both lotus and concanavalin A (ConA) lectins, it bound to lotus stronger than ConA. By varying the central motif RYRY, five versions were produced in multiple antigen peptide format, and their avidity for lotus and ConA lectins was tested by surface plasmon resonance. Although the kinetic parameters were similar, the version based on the sequence YPYRY had an optimal affinity for both lectins as well as improved avidity for wheat germ agglutinin and phytohemagglutinin. Thus, as far as lectin specificity is concerned, YPYRY had improved multiple antigenic properties. Both RYRY and YPYRY precipitated antibodies from human IgG for intravenous use that bound to gp120 in vitro and immunoprecipitated gp120 from transfected CHO-PI cells. Thus, Y(P/R)Y motifs mimic multiple carbohydrate epitopes, many of which are found on HIV, and preimmune human IgG antibodies that bind to HIV carbohydrates cross-react to a comparable extent with both RYRY and YPYRY carbohydrate mimetic peptides.  相似文献   

14.
The human immunodeficiency virus type 1 (HIV-1) capsid protein plays a critical role in virus core particle assembly and is an important target for novel therapeutic strategies. In a previous study, we characterized the binding affinity of a hydrocarbon stapled helical peptide, NYAD-1, for the capsid protein (K(d) approximately 1 mum) and demonstrated its ability to penetrate the cell membrane (Zhang, H., Zhao, Q., Bhattacharya, S., Waheed, A. A., Tong, X., Hong, A., Heck, S., Goger, M., Cowburn, D., Freed, E. O., and Debnath, A. K. (2008) J. Mol. Biol. 378, 565-580). In cell-based assays, NYAD-1 colocalized with the Gag polyprotein during traffic to the plasma membrane and disrupted the formation of mature and immature virus particles in vitro systems. Here, we complement the cellular and biochemical data with structural characterization of the interactions between the capsid and a soluble peptide analogue, NYAD-13. Solution NMR methods were used to determine a high resolution structure of the complex between the inhibitor and a monomeric form of the C-terminal domain of the capsid protein (mCA-CTD). The intermolecular interactions are mediated by the packing of hydrophobic side chains at the buried interface and unperturbed by the presence of the olefinic chain on the solvent-exposed surface of the peptide. The results of the structural analysis provide valuable insight into the determinants for high affinity and selective inhibitors for HIV-1 particle assembly.  相似文献   

15.
The fibronectin-binding proteins FnBPA and FnBPB are multifunctional adhesins than can also bind to fibrinogen and elastin. In this study, the N2N3 subdomains of region A of FnBPB were shown to bind fibrinogen with a similar affinity to those of FnBPA (2 μM). The binding site for FnBPB in fibrinogen was localized to the C-terminus of the γ-chain. Like clumping factor A, region A of FnBPB bound to the γ-chain of fibrinogen in a Ca(2+)-inhibitable manner. The deletion of 17 residues from the C-terminus of domain N3 and the substitution of two residues in equivalent positions for crucial residues for fibrinogen binding in clumping factor A and FnBPA eliminated fibrinogen binding by FnBPB. This indicates that FnBPB binds fibrinogen by the dock-lock-latch mechanism. In contrast, the A domain of FnBPB bound fibronectin with K(D) = 2.5 μM despite lacking any of the known fibronectin-binding tandem repeats. A truncate lacking the C-terminal 17 residues (latching peptide) bound fibronectin with the same affinity, suggesting that the FnBPB A domain binds fibronectin by a novel mechanism. The substitution of the two residues required for fibrinogen binding also resulted in a loss of fibronectin binding. This, combined with the observation that purified subdomain N3 bound fibronectin with a measurable, but reduced, K(D) of 20 μM, indicates that the type I modules of fibronectin bind to both the N2 and N3 subdomains. The fibronectin-binding ability of the FnBPB A domain was also functional when the protein was expressed on and anchored to the surface of staphylococcal cells, showing that it is not an artifact of recombinant protein expression.  相似文献   

16.
The complete amino acid sequence of 2-acetamido-2-deoxy-D-galactose-binding Cytisus scoparius seed lectin II (CSII) was determined using a protein sequencer. After digestion of CSII with endoproteinase Lys-C or Asp-N, the resulting peptides were purified by reversed-phase high performance liquid chromatography (HPLC) and then subjected to sequence analysis. Comparison of the complete amino acid sequence of CSII with the sequences of other leguminous seed lectins revealed regions of extensive homology. The amino acid residues of concanavalin A (Con A) involved in the metal binding site are highly conserved among those of CSII. A carbohydrate-binding peptide of CSII was obtained from the endoproteinase Asp-N digest of CSII by affinity chromatography on a column of GalNAc-Gel. This peptide was retained on the GalNAc-Gel column and was presumed to have affinity for the column. The amino acid sequence of the retarded peptide was determined using a protein sequencer. The retarded peptide was found to correspond to the putative metal-binding region of Con A. These results strongly suggest that this peptide represents the carbohydrate-binding and metal ion-binding sites of CSII.  相似文献   

17.
The DNA binding and structural properties of Xfin-31 (Lee, M.S., Gippert, G.P., Soman, K.V., Case, D.A. and Wright, P.E., 1989, Science 245, 635-637), a twenty five amino acid zinc finger peptide, in the reduced, oxidized and zinc complex forms, as well as the fourteen residue helical segment of the zinc finger (residues 12-25) have been compared using affinity coelectrophoresis (ACE) and circular dichroism (CD) spectroscopy. The zinc complex and oxidized peptides bind cooperatively to DNA although the cooperativity factor, omega, is more than 15-fold greater for the zinc complex. The reduced peptide in the absence of zinc and the helical segment do not bind cooperatively (omega = 1). Hence, the binding constant for singly contiguous sites (K omega) ranges over 100-fold for the various peptides even though the intrinsic binding constants (K) are similar. An increase in binding order and affinity for the other forms of Xfin-31 is correlated with an increasing similarity of the CD spectrum to that of the Xfin-31 zinc complex. The surprising DNA binding activity of the oxidized peptide may result from hydrophobic interactions between the amino-terminal loop formed by the Cys3-Cys6 disulfide bond and conserved hydrophobic residues in the carboxyl-terminal segment. Xfin-31 may be a particularly useful model for studying several poorly understood aspects of cooperative, non-specific DNA binding since it is small, has a stable, well-defined structure, and structures of zinc fingers bound to DNA have been determined.  相似文献   

18.
A photoaffinity substrate analogue, 8-azido-ADP-[14C]glucose, reacts specifically and covalently with Escherichia coli ADP-glucose synthetase. The site(s) of reaction of 8-azido-ADP-[14C]glucose with the enzyme was identified by isolation of tryptic peptides containing the labeled analogue by use of high performance liquid chromatography technique and subsequent NH2-terminal sequence analysis of the purified radioactive peptides. One major binding region of the azido analogue is a peptide segment composed of residues 107-114 of the enzyme's polypeptide chain. Lys 108 and Arg 114 become trypsin-resistant sites when the enzyme is photoinactivated by 8-azido-ADP-[14C] glucose, suggesting that the analogue binds at or near the vicinity of these 2 basic amino acid residues. Conformational analysis of this peptide segment (residues 107-114) shows a strong probability of a reverse beta-turn secondary structure, suggesting that this peptide segment is on the enzyme surface. Two minor reaction regions of the enzyme with the analogue were also identified by chemical characterization. One region was composed of residues 162-207. Lys 194 was previously suggested as the activator-binding site by chemical modification studies with pyridoxal phosphate (Parsons, T. F., and Preiss, J. (1978) J. Biol. Chem. 253, 7638-7645). Another minor region where the analogue binds the tryptic peptide composed of residues 380-385 is near the COOH-terminal side of the enzyme. It is postulated that all these peptide segments are juxtaposed in tertiary structure.  相似文献   

19.
8-Azidoadenine was employed as a photoaffinity probe of the adenine binding site of the seed lectin from lima beans and from Phaseolus vulgaris erythroagglutinin. This compound was shown to (a) bind competitively to the adenine binding site of these lectins and (b) exhibit enhanced binding in the presence of 1,8-anilinonaphthalenesulfonic acid in the same manner as adenine. The presence or absence of 1,8-anilinonaphthalenesulfonic acid during labeling caused no change in the peptide maps of either lectin when digested with trypsin. The peptide maps of each lectin showed one major peak of radioactivity. Sequencing of the corresponding tryptic peptide from lima bean lectin indicated the primary structure to be Val-Leu-Ile-Thr-Tyr-Asp-Ser-Ser-Thr-Lys. The sequence of the labeled peptide isolated from P. vulgaris erythroagglutinin was Thr-Thr-Thr-Trp-Asp-Phe-Val-Gly-Glu-Asn-Glu-Val-Leu-Ile-Thr-Tyr, which corresponded to residues 173-190 of the cDNA-derived sequence (Hoffman, L. M., and Donaldson, D. D. (1985) EMBO J. 4, 883-889). Residues 186-190 (italicized) are identical to the first five amino acids in the lima bean lectin peptide. The peptides are located at the COOH-terminal half of the lectin and show extensive homology with other legume lectins.  相似文献   

20.
Recently the pH gradient evoked by a K+ diffusion potential was shown to translocate a synthetic monobasic amphipathic hexapeptide across the bilayer of lipid vesicles (De Kroon, A.I.P.M., Vogt, B., Van 't Hof, R., De Kruijff, B. and De Gier, J. (1991) Biophys. J. 60, in press). Here this observation is extended by studying the effect of a membrane potential on a set of bioactive peptides. The panel of peptides comprises the toxin mastoparan X, a tryptophan-containing analogue of the presequence of the mitochondrial protein cytochrome oxidase subunit IV (preCoxIV(1-25)W18), and the regulatory peptides ACTH(1-24), alpha-MSH, ACTH(1-10), dynorphin A, bombesin, and LHRH. The interaction of these peptides with phospholipid vesicles has been measured using the intrinsic tryptophan residue as fluorescent probe. In the absence of a K+ diffusion potential only mastoparan X and the presequence show considerable binding to vesicles consisting of phosphatidylcholine (PC). In contrast, under these conditions all peptides display affinity for vesicles consisting of the acidic phospholipid cardiolipin (CL), the extent of which depends on the net positive charge of the peptide. Application of a K+ diffusion potential to large unilamellar vesicles (LUV) consisting of PC results in a time dependent tryptophan fluorescence increase for mastoparan X, which is accelerated upon incorporating increasing amounts of CL into the LUV. A similar fluorescence increase in response to a K+ diffusion potential was observed for the above model peptide. Yet the mechanism resulting in the fluorescence increase of mastoparan X is completely different from that of the hexapeptide. Binding experiments indicate that a membrane potential-induced enhanced binding of the peptide to the outer surface of the vesicles contributes to the fluorescence increase. PreCoxIV(1-25)W18, dynorphin A, and ACTH(1-24) show fluorescence responses upon applying a membrane potential that are consistent with that of mastoparan X, whereas the other peptides tested do not respond up to a LUV CL content of 50%. The results tentatively suggest that the membrane potential only affects a peptide when it has the ability to adopt a stable membrane bound conformation.  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