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 ubiquitous calpains, mu- and m-calpain, have been implicated in essential physiological processes and various pathologies. Cell-permeable specific inhibitors are important tools to elucidate the roles of calpains in cultivated cells and animal models. The synthetic N-acetylated 27-mer peptide derived from exon B of the inhibitory domain 1 of human calpastatin (CP1B) is unique as a potent and highly selective reversible calpain inhibitor, but is poorly cell-permeant. By addition of N-terminal cysteine residues we have generated a disulfide-conjugated CP1B with the cell-penetrating 16-mer peptide penetratin derived from the third helix of the Antennapedia homeodomain protein. The inhibitory potency and selectivity of CP1B for calpain versus cathepsin B and L, caspase 3 and the proteasome was not affected by the conjugation with penetratin. The conjugate was shown to efficiently penetrate into living LCLC 103H cells, since it prevents ionomycin-induced calpain activation at 200-fold lower concentration than the non-conjugated inhibitor and is able to reduce calpain-triggered apoptosis of these cells. Penetratin-conjugated CP1B seems to be a promising alternative to the widely used cell-permeable peptide aldehydes (e.g. calpain inhibitor 1) which inhibit the lysosomal cathepsins and partially the proteasome as well or even better than the calpains.  相似文献   

2.
A library of 121 pseudopeptides was designed to develop reversible inhibitors of trypanosomal enzymes (cruzain from Trypanosoma cruzi and congopain from Trypanosoma congolense). The peptides share the framework: Cha-X1-X2-Pro (Cha=cyclohexyl-alanine, X1 and X2 were phenylalanyl analogs), based on a previous report [Lecaille, F., Authié, E., Moreau, T., Serveau, C., Gauthier, F. and Lalmanach, G. (2001) Eur. J. Biochem. 268, 2733-2741]. Five peptides containing a nitro-substituted aromatic residue (Tyr/Phe) and one a 4-chloro-phenylalanine at the X1 position, and 3-(2-naphthyl)-alanine, homocyclohexylalanine or 3-nitro-tyrosine (3-NO(2)-Tyr) at the X2 position, were selected. They inhibited congopain more effectively than cruzain, except Cha-4-NO(2)-Phe-3-NO(2)-Tyr-Pro which bound the two parasitic enzymes similarly. Among this series, Cha-3-NO(2)-Tyr-HoCha-Pro and Cha-4-NO(2)-Phe-3-NO(2)-Tyr-Pro are the most selective for congopain relative to host cathepsins. No hydrolysis occurred upon prolonged incubation time with purified enzymes. In addition introduction of non-proteogenic residues in the peptidyl backbone greatly enhanced resistance to proteolysis by mammalian sera.  相似文献   

3.
Evidence that a 27-residue sequence is the actin-binding site of ABP-120   总被引:12,自引:0,他引:12  
Proteolysis experiments of ABP-120 from Dictyostelium discoideum have previously demonstrated that removal of residues 89-115 from a tryptic peptide which retains actin binding activity, abolishes actin binding (Bresnick, A. R., Warren, V., and Condeelis, J. (1990) J. Biol. Chem. 265, 9236-9240). Antibodies made against a synthetic peptide of this 27-amino acid sequence (27-mer) specifically immunoprecipitate native ABP-120 from Dictyostelium high speed supernatants, demonstrating that the 27-mer sequence is on the surface of the molecule as expected for an active site. ABP-120 is inhibited in its binding to F-actin by Fab' fragments of the anti-27-mer IgG. Half-maximal inhibition occurs at an approximate molar ratio of 7 Fab' fragments/ABP-120 monomer. Viscoelastic measurements indicate that ABP-120 forms fewer cross-links with F-actin in the presence of the 27-mer synthetic peptide than in its absence. In F-actin cosedimentation assays, the binding of ABP-120 to actin is inhibited by the 27-mer synthetic peptide. Furthermore, the 27-mer synthetic peptide cosediments with F-actin, whereas a control hydrophobic peptide and a synthetic peptide of residues 69-88 of ABP-120 do not cosediment with F-actin. These observations suggest a direct involvement of the 27-mer sequence in the actin binding activity of ABP-120.  相似文献   

4.
The major antifreeze polypeptide (AFP) from winter flounder (37 amino acid residues) is a single alpha-helix. Aspartic acid and arginine are found, respectively, at the amino and carboxyl-termini. These charged amino acids are ideally located for stabilizing the alpha-helical conformation of this AFP by means of charge-dipole interaction (Shoemaker, K. R., Kim, P.S., York, E.J., Stewart, J. M., and Baldwin, R. L. (1987) Nature 326, 563-567). In order to understand these and other molecular interactions that maintain the AFP structure, we have carried out the chemical synthesis of AFP analogs and evaluated their conformations by circular dichroism spectroscopy. We synthesized the entire AFP molecule (37-mer) and six COOH-terminal peptide fragments (36-, 33-, 27-, 26-, 16-, and 15-mers). Peptides containing acidic NH2-terminal residues displayed greater helix formation and thermal stability compared to those peptides of similar size, but with neutral NH2-terminal residues. Helix formation was maximum above pH 9.2. The peptide conformations also displayed a pH-dependent sensitivity to changes in ionic strength. Helix formation was reduced in the presence of acetonitrile. We conclude that the AFP helix is most likely stabilized by: charge-dipole interactions between charged terminal amino acids and the helix dipole, a charge interaction between Lys18 and Glu22 (either a salt bridge or a hydrogen bond), and hydrophobic interactions.  相似文献   

5.
Proton and phosphorus NMR studies are reported for two complementary nonanucleotide duplexes containing acyclic abasic sites. The first duplex, d(C-A-T-G-A-G-T-A-C).d(G-T-A-C-P-C-A-T-G), contains an acyclic propanyl moiety, P, located opposite a deoxyadenosine at the center of the helix (designated APP 9-mer duplex). The second duplex, d(C-A-T-G-A-G-T-A-C).d(G-T-A-C-E-C-A-T-G), contains a similarly located acyclic ethanyl moiety, E (designated APE 9-mer duplex). The ethanyl moiety is one carbon shorter than the natural carbon-phosphodiester backbone of a single nucleotide unit of DNA. The majority of the exchangeable and nonexchangeable base and sugar protons in both the APP 9-mer and APE 9-mer duplexes, including those at the abasic site, have been assigned by recording and analyzing two-dimensional phase-sensitive NOESY data sets in H2O and D2O solution between -5 and 5 degrees C. These spectroscopic observations establish that A5 inserts into the helix opposite the abasic site (P14 and E14) and stacks between the flanking G4.C15 and G6.C13 Watson-Crick base pairs in both the APP 9-mer and APE 9-mer duplexes. The helix is right-handed at and adjacent to the abasic site, and all glycosidic torsion angles are anti in both 9-mer duplexes. Proton NMR parameters for the APP 9-mer and APE 9-mer duplexes are similar to those reported previously for the APF 9-mer duplex (F = furan) in which a cyclic analogue of deoxyribose was embedded in an otherwise identical DNA sequence [Kalnik, M. W., Chang, C. N., Grollman, A. P., & Patel, D. J. (1988) Biochemistry 27, 924-931]. These proton NMR experiments demonstrate that the structures at abasic sites are very similar whether the five-membered ring is open or closed or whether the phosphodiester backbone is shortened by one carbon atom. Phosphorus spectra of the APP 9-mer and APE 9-mer duplexes (5 degrees C) indicate that the backbone conformation is similarly perturbed at three phosphodiester backbone torsion angles. These same torsion angles are also distorted in the APF 9-mer but assume a different conformation than those in the APP 9-mer and APE 9-mer duplexes.  相似文献   

6.
The structure of mitochondria is highly dynamic and depends on the balance of fusion and fission processes. Deletion of the mitochondrial dynamin-like protein Mgm1 in yeast leads to extensive fragmentation of mitochondria and loss of mitochondrial DNA. Mgm1 and its human ortholog OPA1, associated with optic atrophy type I in humans, were proposed to be involved in fission or fusion of mitochondria or, alternatively, in remodeling of the mitochondrial inner membrane and cristae formation (Wong, E. D., Wagner, J. A., Gorsich, S. W., McCaffery, J. M., Shaw, J. M., and Nunnari, J. (2000) J. Cell Biol. 151, 341-352; Wong, E. D., Wagner, J. A., Scott, S. V., Okreglak, V., Holewinske, T. J., Cassidy-Stone, A., and Nunnari, J. (2003) J. Cell Biol. 160, 303-311; Sesaki, H., Southard, S. M., Yaffe, M. P., and Jensen, R. E. (2003) Mol. Biol. Cell, in press). Mgm1 and its orthologs exist in two forms of different lengths. To obtain new insights into their biogenesis and function, we have characterized these isoforms. The large isoform (l-Mgm1) contains an N-terminal putative transmembrane segment that is absent in the short isoform (s-Mgm1). The large isoform is an integral inner membrane protein facing the intermembrane space. Furthermore, the conversion of l-Mgm1 into s-Mgm1 was found to be dependent on Pcp1 (Mdm37/YGR101w) a recently identified component essential for wild type mitochondrial morphology. Pcp1 is a homolog of Rhomboid, a serine protease known to be involved in intercellular signaling in Drosophila melanogaster, suggesting a function of Pcp1 in the proteolytic maturation process of Mgm1. Expression of s-Mgm1 can partially complement the Deltapcp1 phenotype. Expression of both isoforms but not of either isoform alone was able to partially complement the Deltamgm1 phenotype. Therefore, processing of l-Mgm1 by Pcp1 and the presence of both isoforms of Mgm1 appear crucial for wild type mitochondrial morphology and maintenance of mitochondrial DNA.  相似文献   

7.
Yano Y  Matsuzaki K 《Biochemistry》2002,41(41):12407-12413
An important subject for elucidating membrane protein (MP) folding is how transmembrane helices (TMHs) insert into and dissociate from membranes. We investigated helix dissociation kinetics and insertion topology by means of intervesicular transfer of the fluorophore-labeled completely hydrophobic model transmembrane helix NBD-(LALAAAA)(3)-NH(2) (NBD = 7-nitro-2-1,3-benzoxadiazol-4-yl). The peptide forms a topologically stable transmembrane helix, which is in a monomer-antiparallel dimer equilibrium [Yano, Y., Takemoto, T., Kobayashi, S., Yasui, H., Sakurai, H., Ohashi, W., Niwa, M., Futaki, S., Sugiura, Y., and Matsuzaki, K. (2002) Biochemistry 41, 3073-3080]. The helix transfer kinetics, representing the helix dissociation process, was monitored by fluorescence recovery of the quenched peptide in donor vesicles containing a quencher upon its transfer to acceptor vesicles without the quencher. The transfer kinetics and vesicle concentration dependence demonstrated that the transfer was mediated by monomer in the aqueous phase. Furthermore, the activation enthalpy was estimated to be +17.7 +/- 1.3 kcal mol(-1). Helix insertion topology, detected by chemical quenching of the NBD group in the outer leaflet by dithionite ions, was found to be controlled by transmembrane electric potential-helix macro dipole interaction. On the basis of these observations, a model for the helix insertion/dissociation processes was discussed.  相似文献   

8.
Y G Chu  I Tinoco 《Biopolymers》1983,22(4):1235-1246
The kinetics of helix formation were investigated using the temperature-jump technique for the following two molecules: dC-G-T-G-A-A-T-T-C-G-C-G, which forms a double helix containing a G·T base pair(the G·T 12-mer), and dC-G-C-A-G-A-A-T-T-C-G-C-G, which forms a double helix containing an extra adenine (the 13-mer). When data were analyzed in an all-or-none model, the activation energy for the helix association process was 22 ± 4 kcal/mol for the G·T 12-mer and 16 ± 7 kcal/mol for the 13-mer. The activation energy for the helix-dissociation process was 68 ± 2 kcal/mol for the G·T 12-mer and 74 ± 3 kcal/mol for the 13-mer. Rate constants for recombination were near 105s?1M?1 in the temperature range from 32 to 47°C; for the dissociation process, the rate constants varied from 1s?1 near 32°C to 130s?1 near 47°C. Possible effects of hairpin loops and fraying ends on the above data are discussed.  相似文献   

9.
10.
11.
The binding of a peptide to a biological membrane is often accompanied by a transition from a random coil structure to an amphipathic alpha-helix. Recently, we have presented a new approach which allows the determination of the thermodynamic parameters of membrane-induced helix formation [Wieprecht et al. (1999) J. Mol Biol. 294, 785]. It involves a systematic variation of the helix content of a given peptide by double D-substitution and a correlation of the binding parameters with the helicity. Here we have used this method to study membrane-induced helix formation for the presequence of rat mitochondrial rhodanese (RHD). The thermodynamic parameters of binding of the peptide RHD and of four of its double D-isomers were determined for 30 nm (SUVs) and 100 nm (LUVs) unilamellar vesicles composed of phosphatidylcholine/phosphatidylglycerol (3:1) using circular dichroism spectroscopy, fluorescence spectroscopy, and isothermal titration calorimetry. The incremental changes of the thermodynamic parameters of helix formation were found to be very similar for SUVs and LUVs. Membrane-induced helix formation of RHD entailed a negative enthalpy of Delta H(helix) = -0.5 to -0.6 kcal/mol/residue and was opposed by an entropy of about Delta S(helix) = -1 to -1.4 cal/mol K/residue. The free energy of helix formation, Delta G(helix), was about -0.2 kcal/mol, and helix formation accounted for 50-60% of the total free energy of membrane binding. Dye-release experiments were used to assess the role of helix formation for the membrane perturbation potential of the peptides. While helix formation plays a major role for membrane binding, it appears to have little importance for inducing membrane leakiness.  相似文献   

12.
The 27-mer peptide CP1B-[1-27] derived from exon 1B of calpastatin stands out among the known inhibitors for mu- and m-calpain due to its high potency and selectivity. By systematical truncation, a 20-mer peptide, CP1B-[4-23], was identified as the core sequence required to maintain the affinity/selectivity profile of CP1B-[1-27]. Starting with this peptide, the turn-like region Glu(10)(i)-Leu(11)(i+1)-Gly(12)(i+2)-Lys(13)(i+3) was investigated. Sequence alignment of subdomains 1B, 2B, 3B and 4B from different mammalians revealed that the amino acid residues in position i+1 and i+2 are almost invariably flanked by oppositely charged residues, pointing towards a turn-like conformation stabilized by salt bridge/H-bond interaction. Accordingly, using different combinations of acidic and basic residues in position i and i+3, a series of conformationally constrained variants of CP1B-[4-23] were synthesized by macrolactamization utilizing the side chain functionalities of these residues. With the combination of Glu(i)/Dab(i+3), the maximum of conformational rigidity without substantial loss in affinity/selectivity was reached. These results clearly demonstrate that the linear peptide chain corresponding to subdomain 1B reverses its direction in the region Glu(10)-Lys(13) upon binding to mu-calpain, and thereby adopts a loop-like rather than a tight turn conformation at this site.  相似文献   

13.
The heterotrimeric G protein Galphao is ubiquitously expressed throughout the central nervous system, but many of its functions remain to be defined. To search for novel proteins that interact with Galphao, a mouse brain library was screened using the yeast two-hybrid interaction system. Pcp2 (Purkinje cell protein-2) was identified as a partner for Galphao in this system. Pcp2 is expressed in cerebellar Purkinje cells and retinal bipolar neurons, two locations where Galphao is also expressed. Pcp2 was first identified as a candidate gene to explain Purkinje cell degeneration in pcd mice (Nordquist, D. T., Kozak, C. A., and Orr, H. T. (1988) J. Neurosci. 8, 4780-4789), but its function remains unknown as Pcp2 knockout mice are normal (Mohn, A. R., Feddersen, R. M., Nguyen, M. S., and Koller, B. H. (1997) Mol. Cell. Neurosci. 9, 63-76). Galphao and Pcp2 binding was confirmed in vitro using glutathione S-transferase-Pcp2 fusion proteins and in vitro translated [35S]methionine-labeled Galphao. In addition, when Galphao and Pcp2 were cotransfected into COS cells, Galphao was detected in immunoprecipitates of Pcp2. To determine whether Pcp2 could modulate Galphao function, kinetic constants kcat and koff of bovine brain Galphao were determined in the presence and absence of Pcp2. Pcp2 stimulates GDP release from Galphao more than 5-fold without affecting kcat. These findings define a novel nucleotide exchange function for Pcp2 and suggest that the interaction between Pcp2 and Galphao is important to Purkinje cell function.  相似文献   

14.
The three-dimensional structure of the anti-apoptotic protein Bcl-xL complexed to a 25-residue peptide from the death promoting region of Bad was determined using NMR spectroscopy. Although the overall structure is similar to Bcl-xL bound to a 16-residue peptide from the Bak protein (Sattler et al., 1997), the Bad peptide forms additional interactions with Bcl-xL. However, based upon site-directed mutagenesis experiments, these additional contacts do not account for the increased affinity of the Bad 25-mer for Bcl-xL compared to the Bad 16-mer. Rather, the increased helix propensity of the Bad 25-mer is primarily responsible for its greater affinity for Bcl-xL. Based on this observation, a pair of 16-residue peptides were designed and synthesized that were predicted to have a high helix propensity while maintaining the interactions important for complexation with Bcl-xL. Both peptides showed an increase in helix propensity compared to the wild-type and exhibited an enhanced affinity for Bcl-xL.  相似文献   

15.
Helix formation of an S-peptide analog, comprising the first 20 residues of Ribonuclease A and two additional N-terminal residues, was studied by measuring hydrogen bond (H-bond) (h3)J(NC') scalar couplings as a function of 2,2,2-trifluoroethanol (TFE) concentration. The (h3)J(NC') couplings give direct evidence for the closing of individual backbone N-H***O = C H-bonds during the TFE-induced formation of secondary structure. Whereas no (h3)J(NC') correlations could be detected without TFE, alpha-helical (i,i +4) H-bond correlations were observed for the amides of residues A5 to M15 in the presence of TFE. The analysis of individual coupling constants indicates that alpha-helix formation starts at the center of the S-peptide around residue E11 and proceeds gradually from there to both peptide ends as the TFE concentration is increased. At 60% to 90% TFE, well-formed alpha-helical H-bonds were observed for the amides hydrogens of residues K9 to Q13, whereas H-bonds of residues T5 to A8, H14, and M15 are affected by fraying. No intramolecular backbone H-bonds are present at and beyond the putative helix stop signal D16. As the (h3)J(NC') constants represent ensemble averages and the dependence of (h3)J(NC') on H-bond lengths is very steep, the size of the individual (h3)J(NC') coupling constants can be used as a measure for the population of a closed H-bond. These individual populations are in agreement with results derived from the Lifson-Roig theory for coil-to-helix transitions. The present work shows that the closing of individual H-bonds during TFE-induced helix formation can be monitored by changes in the size of H-bond scalar couplings.  相似文献   

16.
Eukaryotic topoisomerase II is capable of binding two separate nucleic acid helices prior to its DNA cleavage and strand passage events (Zechiedrich, E. L., and Osheroff, N (1990) EMBO J. 9, 4555-4562). Presumably, one of these helices represents the helix that the enzyme cleaves (i.e. cleavage helix), and the other represents the helix that it passes (i.e. passage helix) through the break in the nucleic acid backbone. To determine whether the passage helix is required for reaction steps that precede the enzyme's DNA strand passage event, interactions between Drosophila melanogaster topoisomerase II and a short double-stranded oligonucleotide were assessed. These studies employed a 40-mer that contained a specific recognition/cleavage site for the enzyme. The sigmoidal DNA concentration dependence that was observed for cleavage of the 40-mer indicated that topoisomerase II had to interact with more than a single oligonucleotide in order for cleavage to take place. Despite this requirement, results of enzyme DNA binding experiments indicated no binding cooperativity for the 40-mer. These findings strongly suggest a two-site model for topoisomerase II action in which the passage and the cleavage helices bind to the enzyme independently, but the passage helix must be present for efficient topoisomerase II-mediated DNA cleavage to occur.  相似文献   

17.
The manner by which peptidic ligands bind and activate their corresponding G-protein-coupled receptors is not well understood. One of the better characterized peptidic ligands is the chemotactic cytokine complement factor 5a (C5a), a 74-amino acid helical bundle. Previous studies showed 6-mer peptide analogs derived from the C terminus of the C5a ligand can bind to C5aR (Kd values approximately 0.1-1 microm) and either agonize or antagonize the receptor (Gerber, B. O., Meng, E. C., Dotsch, V., Baranski, T. J., and Bourne, H. R. (2001) J. Biol. Chem. 276, 3394-3400). Here, we provide direct biochemical data using disulfide trapping to support a model that these peptides bind within a transmembrane helical triad formed by alpha-helices III, VI, and VII. We show that the three amino acids on the C terminus of the peptide analogs bind too weakly to exert a functional effect themselves. However, when a cysteine residue is placed on their N terminus they can be trapped by disulfide interchange to specific cysteines in helix III and VI and not to other cysteines, engineered into the C5aR. The trapped peptides function as agonists or partial antagonists, similar to the non-covalent parents from which they were derived. These data help to further refine the binding mode for C5a to the C5aR and suggest an approach and a binding site that may be applicable to studying other peptide binding receptors.  相似文献   

18.
The formyl peptide receptor (FPR) is a chemotactic G protein-coupled receptor found on the surface of phagocytes. We have previously shown that the formyl peptide binding site maps to the membrane-spanning region (Miettinen, H. M., Mills, J. S., Gripentrog, J. M., Dratz, E. A., Granger, B. L., and Jesaitis, A. J. (1997) J. Immunol. 159, 4045-4054). Recent reports have indicated that non-formylated peptides, such as MMWLL can also activate this receptor (Chen, J., Bernstein, H. S., Chen, M., Wang, L., Ishi, M., Turck, C. W., and Coughlin, S. R. (1995) J. Biol. Chem. 270, 23398-23401.) Here we show that the selectivity for the binding of different NH(2)-terminal analogs of MMWLL or MLF can be markedly altered by mutating Asp-106 to asparagine or Arg-201 to alanine. Both D106N and R201A produced a similar change in ligand specificity, including an enhanced ability to bind the HIV-1 peptide DP178. In contrast, the mutation R205A exhibited altered specificity at the COOH terminus of fMLF, with R205A binding fMLF-O-butyl > fMLF-O-methyl > fMLF, whereas wt FPR bound fMLF > fMLF-O-methyl approximately fMLF-O-butyl. These data, taken together with our previous finding that the leucine side chain of fMLF is probably bound to FPR near FPR (93)VRK(95) (Mills, J. S., Miettinen, H. M., Barnidge, D., Vlases, M. J., Wimer-Mackin, S., Dratz, E. A., and Jesaitis, A. J. (1998) J. Biol. Chem. 273, 10428-10435.), indicate that the most likely positioning of fMLF in the binding pocket of FPR is approximately parallel to the fifth transmembrane helix with the formamide group of fMLF hydrogen-bonded to both Asp-106 and Arg-201, the leucine side chain pointing toward the second transmembrane region, and the COOH-terminal carboxyl group of fMLF ion-paired with Arg-205.  相似文献   

19.
Proton and phosphorus NMR studies are reported for the complementary d(C-A-T-G-A-G-T-A-C).d(G-T-A-C-F-C-A-T-G) nonanucleotide duplex (designated APF 9-mer duplex) which contains a stable abasic site analogue, F, in the center of the helix. This oligodeoxynucleotide contains a modified tetrahydrofuran moiety, isosteric with 2-deoxyribofuranose, which serves as a structural analogue of a natural apurinic/apyrimidinic site [Takeshita, M., Chang, C.N., Johnson, F., Will, S., & Grollman, A.P. (1987) J. Biol. Chem. 262, 10171-10179]. Exchangeable and nonexchangeable base and sugar protons, including those located at the abasic site, have been assigned in the complementary APF 9-mer duplex by recording and analyzing two-dimensional phase-sensitive NOESY data sets in H2O and D2O solution at low temperature (0 degrees C). These studies indicate that A5 inserts into the helix opposite the abasic site F14 and stacks with flanking G4.C15 and G6.C13 Watson-Crick base pairs. Base-sugar proton NOE connectivities were measured through G4-A5-G6 on the unmodified strand and between the base protons of C15 and the sugar protons of the 5'-flanking residue F14 on the modified strand. These studies establish that all glycosidic torsion angles are anti and that the helix is right-handed at and adjacent to the abasic site in the APF 9-mer duplex. Two of the 16 phosphodiester groups exhibit phosphorus resonances outside the normal spectral dispersion indicative of altered torsion angles at two of the phosphate groups in the backbone of the APF 9-mer duplex.  相似文献   

20.
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