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.
Hong J  Oren Z  Shai Y 《Biochemistry》1999,38(51):16963-16973
Recently, we reported on a new group of diastereomers of short-model peptides (12 amino acids long) composed of leucine and lysine with varying ratios, possessing several properties that make them potentially better than native or de novo-designed all L-amino acid antimicrobial peptides. Preliminary studies have revealed that modulating the hydrophobicity and positive charges of these diastereomers is sufficient to confer antibacterial activity and cell selectivity. However, the relationship between their biological function, structure, and mode of action was not investigated. Here we synthesized and investigated three types of linear model diastereomers (12 amino acids long) with varying lysine:leucine (or tryptophan) ratios (i.e., K(3)L(8)W, K(5)L(6)W, and K(7)L(4)W), which confer different levels of lytic activities. For each K:L ratio, tryptophan was introduced in the middle or the N- or C-terminus of the peptides, as an intrinsic fluorescent probe. Only the hemolytic peptide K(3)L(8)W binds to both negatively charged and zwitterionic phospholipid membranes. K(5)L(6)W and K(7)L(4)W bind similarly, but only to negatively charged membranes, despite the fact that K(5)L(6)W is substantially more lytic to bacteria than K(7)L(4)W. Interestingly, although K(3)L(8)W contains 33% D-amino acids, ATR-FTIR spectroscopy revealed a structure of approximately 90% alpha-helix in both types of membranes. In addition, K(5)L(6)W contains approximately 40% 3(10)-helix and K(7)L(4)W is predominantly a random coil in membranes. Polarized ATR-FTIR and tryptophan-quenching experiments, using brominated phospholipids, revealed a similar depth of penetration and an orientation that was parallel to the membrane surface for all the peptides, but with K(3)L(8)W affecting the lipid order more than the others. The results provide insight into the mode of action of this group of diastereomeric peptides, and the effect of hydrophobicity and positive charges on their membrane structure, function, and cell selectivity. Moreover, this research should assist in the development of suitable diastereomeric peptide antibiotics for therapeutic use that would overcome the problem the increasing resistance of bacteria to conventional antibiotics.  相似文献   

2.
A 24-amino acid peptide, Humanin (HN), is a novel peptide that protects neuronal cells in vitro and in vivo from Alzheimer's disease-related toxicities. We have shown before that the structures of HN and a 1000-fold more active analog, HNG, with a Ser14Gly mutation are largely disordered. During additional mutational analysis, a shorter 17-amino acid form, AGA-(C8R)HNG17, was accidentally discovered to have a 100-fold higher activity than HNG. Here we have characterized the structural properties of the AGA-(C8R)HNG17 analog by circular dichroism (CD) and sedimentation equilibrium analysis. First, the structure in water was characterized, since these peptides have been dissolved in water prior to biological analysis. The AGA-(C8R)HNG17 peptide exhibited extensive beta-sheet structure in water, completely different from the aqueous HN and HNG structures. The beta-sheet structure was converted to a disordered structure upon dilution into phosphate-buffered saline (PBS) at low peptide concentration (e.g., below 0.2mg/ml), which was similar to the structure of HN and HNG, observed under similar conditions. Sedimentation equilibrium analysis showed that the AGA-(C8R)HNG17 analog was essentially monomeric in PBS, while HNG showed extensive aggregation. Such aggregation of HNG was observed when the peptide was added to the serum-containing cell culture media. Thus, the mutations introduced into the AGA-(C8R)HNG17 analog generated a peptide different from the parent HNG and HN peptides in the self-association properties and hence the solubility, which most likely contributed to the increased biological activity of the AGA-(C8R)HNG17 analog.  相似文献   

3.
A series of square-planar Pd(II) complexes of the composition cis-[Pd(L(n))(2)Cl(2)] {L(1)=2-chloro-6-benzylamino-9-isopropylpurine (1), L(2)=2-chloro-6-[(4-methoxybenzyl)amino]-9-isopropylpurine (2), L(3)=2-chloro-6-[(2-methoxybenzyl)amino]-9-isopropylpurine (3) and 2-[(chloropropyl)amino]-6-benzylamino-9-isopropylpurine (6)} has been synthesized by the reaction of PdCl(2) with L(n) in a 1:2 molar ratio. In contrast, the same reaction followed by recrystallization of the product from N,N'-dimethylformamide (DMF) leads to trans-[Pd(L(n))(2)Cl(2)] x nDMF {L(3), n=0 (4), n=1(4( *)DMF); L(4)=2-chloro-6-[(2,3-dimethoxybenzyl)-amino]-9-isopropylpurine, n=0 (5), n=1.5 (5( *)DMF). The compounds have been characterized by elemental analyses, conductivity measurements, electrospray mass spectra in the positive ion mode (ES+MS), FTIR, (1)H and (13)C NMR spectra, thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). Moreover, the complexes 2 and 6 have been also investigated by (15)N NMR spectroscopy. The molecular structures of L(5), {(H(2+)L(5))(Cl(-))(2)} x H(2)O, i.e. the protonated form of L(5), trans-[Pd(L(3))(2)Cl(2)] (4) and trans-[Pd(L(4))(2)Cl(2)] (5) have been determined by single crystal X-ray analysis. NMR data and X-ray structures revealed that the organic molecules are coordinated to Pd via N7 atom of a purine moiety. All the complexes and the corresponding ligands have been tested in vitro for their cytotoxicity against four human cancer cell lines: breast adenocarcinoma (MCF7), malignant melanoma (G361), chronic myelogenous leukaemia (K562) and osteogenic sarcoma (HOS). Promising in vitro cytotoxic effect has been found for cis-[Pd(L(2))(2)Cl(2)] (2), having the IC(50) values of 12, 10, 25, and 14 microM against MCF7, G361, K562, and HOS, respectively, and for trans-[Pd(L(3))(2)Cl(2)].DMF (4) with the IC(50) value of 15 microM against G361.  相似文献   

4.
C-Terminal binding protein (CtBP) interacts with a highly conserved amino acid motif (PXDLS) at the C terminus of adenovirus early region 1A (AdE1A) protein. This amino acid sequence has recently been demonstrated in the mammalian protein C-terminal interacting protein (CtIP) and a number of Drosophila repressors including Snail, Knirps and Hairy. In the study described here we have examined the structures of synthetic peptides identical to the CtBP binding sites on these proteins using NMR spectroscopy. It has been shown that peptides identical to the CtBP binding site in CtIP and at the N terminus of Snail form a series of beta-turns similar to those seen in AdE1A. The PXDLS motif towards the C terminus of Snail forms an alpha-helix. However, the motifs in Knirps and Hairy did not adopt well-defined structures in TFE/water mixtures as shown by the absence of medium range NOEs and a high proportion of signal overlap. The affinities of peptides for Drosophila and mammalian CtBP were compared using enzyme-linked immunosorbent assay. CtIP, Snail (N-terminal peptide) and Knirps peptides all bind to mammalian CtBP with high affinity (K(i) of 1.04, 1.34 and 0.52 microM, respectively). However, different effects were observed with dCtBP, most notably the affinity for the Snail (N-terminal peptide) and Knirps peptides were markedly reduced (K(i) of 332 and 56 microM, respectively) whilst the Hairy peptide bound much more strongly (K(i) for dCtBP of 6.22 compared to 133 microM for hCtBP). In addition we have shown that peptides containing identical PXDLS motifs but with different N and C terminal sequences have appreciably different affinities for mammalian CtBP and different structures in solution. We conclude that the factors governing the interactions of CtBPs with partner proteins are more complex than simple possession of the PXDLS motif. In particular the overall secondary structures and amino acid side chains in the binding sites of partner proteins are of importance as well as possible global structural effects in both members of the complex. These data are considered evidence for a multiplicity of CtBPs and partner proteins in the cell.  相似文献   

5.
To develop novel Pro-rich model AMPs with shorter length and higher bacterial selectivity/therapeutic index (TI) than natural AMP, indolicidin, we synthesized a series of undodecapeptides derived from the sequence XXPXXPWXPXX-NH2 (X indicates Leu or Lys) with different ratios of Lys and Leu residues. Several Pro-rich model peptides (K7 WP3, K6 WL1 P3, K5 WL2 P3-1, K5 WL2 P3-2, and K4 WL3 P3) had approximate 8- to 11-fold higher bacterial selectivity/TI compared to indolicidin. These peptides selectively bind to negatively charged liposomes (EYPG/EYPG; 7:3, w/w) mimicking bacterial membranes. Their high selectivity to negatively charged phospholipids corresponds well with their high bacterial selectivity. Indolicidin showed almost complete depolarization of the cytoplasmic membrane of Staphylococcus aureus and dye-leakage from negatively charged liposomes at 10 microM, whereas all of Pro-rich model peptides had very little activity in these assays even at 80 microM, as observed in buforin 2. These results suggest that the ultimate target of our designed Pro-rich model peptides is probably the intracellular components (e.g. protein, DNA or RNA) rather than the cytoplasmic membranes. Collectively, our designed Pro-rich short model peptides appear to be excellent candidates for future development as a novel antimicrobial agent.  相似文献   

6.
Activity-dependent neurotrophic factor 9 (ADNF9) and NAP are nine and eight amino acid peptides, which exhibit neuroprotective activity at femtomolar concentrations against cell toxic agents. We have here characterized their structures and interactions with dodecylphosphocholine (DPC) in phosphate-buffered saline (PBS). Circular dichroism analysis showed that ADNF9 and NAP are structurally disordered in PBS independent of peptide concentration and temperature, but appear to assume different secondary structure at increasing temperature. Sedimentation equilibrium analysis showed that both ADNF9 and NAP are monomeric at 37 °C, suggesting no self-association under physiological conditions. No secondary structure changes were observed in the presence of DPC, suggesting that ADNF9 and NAP do not interact with lipids.  相似文献   

7.
Spatial structures of proteolytic segment A (sA) of bacterioopsin of H. halobium (residues 1-36) solubilized in a mixture of methanol-chloroform (1:1), 0.1 M LiClO4 organic mixture, or in perdeuterated sodium dodecyl sulfate (SDS) micelles, were determined by 2D 1H-NMR techniques. 324 and 400 NOESY cross-peak volumes were measured in NOESY spectra of sA in organic mixture and SDS micelles, respectively. The sA spatial structures were determined by local structure analysis, distance geometry calculation with program DIANA and systematic search for energetically allowed side chain rotamers consistent with NOESY cross-peak volumes. The structures of sA are similar in both milieus and have the right-handed alpha-helical region from Pro8 to Met32 with root mean square deviation (RMSD) of 0.25 A between backbone heavy atoms and fit well with Pro8 to Met32 alpha-helical region in electron cryo-microscopy model of bacteriorhodopsin. The N-terminal region Ala2-Gly6 of sA in organic mixture has a fixed structure of two consecutive gamma-turns as 2 * 2(7)-helix (RMSD of 0.25 A) stabilized by the Thr5 NH...O = C Gln3 and Ile4 NH...O = C Ala2 hydrogen bonds while this region in SDS micelles has disordered structure with RMSD of 1.44 A for backbone heavy atoms. The C-terminal region Gly33-Asp36 of sA is disordered in both milieus. Torsion angles chi 1 of sA were unequivocally determined for 13 (SDS) and 11 (organic mixture) of alpha-helical residues and are identical in both milieus.  相似文献   

8.
Histatin 5 (Asp1-Ser-His-Ala4-Lys-Arg-His-His8-Gly-Tyr-Lys-Arg12-Lys-Ph e-His-Glu16-Lys-His - His-Ser20-His-Arg-Gly-Tyr24), one of the basic histidine-rich peptides present in human parotid saliva and several of its fragments, 1-16 (N16), 9-24 (C16), 11-24 (C14), 13-24 (C12), 15-24 (C10), and 7-16 (M10), were synthesized by solid-phase procedures. Native histatin 5 from human parotid saliva was also purified. Their antifungal activities on two strains of Candida albicans have been studied and their conformational preferences both in aqueous and non-aqueous solutions examined by circular dichroism. The synthetic histatin 5, C16, and C14 peptides were highly active and inhibited the growth of C. albicans. The candidacidal activity data of synthetic histatin 5 were comparable to the values of the native histatin 5 isolated from parotid saliva and those reported previously, although the assay system used and the strains examined were different. The C16 fragment was as active as the whole peptide itself, whereas the N16 fragment was far less active than C14, suggesting that the sequence at the C-terminal is important for its fungicidal activity. An increase in the chain length of the C-terminal sequence from 12 to 16 residues increased the candidacidal activity, thereby indicating that a peptide chain length of at least 12 residues is necessary to elicit optimum biological activity. The CD spectra of these linear peptides showed that they are structurally more flexible, and they adopt different conformations depending on the solvent environment. CD studies provided evidence that histatin 5 and the longer fragments, C16, N16, and C14 preferred alpha-helical conformations in non-aqueous solvents such as trifluoroethanol and methanol, while in water and pH 7.4 phosphate buffers, they favored random coil structures. The shorter sequences seemed to adopt either turn structures or unordered structures both in aqueous and non-aqueous solutions. It appears that the sequence at the C-terminal of histatin 5 with a minimum chain length of 14 residues and alpha-helical conformation are the important structural requirements for appreciable candidacidal activity.  相似文献   

9.
Structural complexes of the eukaryotic translation initiation factor 4E (eIF4E) with a series of N(7)-alkylated guanosine derivative mRNA cap analogue structures have been characterised. Mass spectrometry was used to determine apparent gas-phase equilibrium dissociation constants (K(d)) values of 0.15 microM, 13.6 microM, and 55.7 microM for eIF4E with 7-methyl-GTP (m(7)GTP), GTP, and GMP, respectively. For tight and specific binding to the eIF4E mononucleotide binding site, there seems to be a clear requirement for guanosine derivatives to possess both the delocalised positive charge of the N(7)-methylated guanine system and at least one phosphate group. We show that the N(7)-benzylated monophosphates 7-benzyl-GMP (Bn(7)GMP) and 7-(p-fluorobenzyl)-GMP (FBn(7)GMP) bind eIF4E substantially more tightly than non-N(7)-alkylated guanosine derivatives (K(d) values of 7.0 microM and 2.0 microM, respectively). The eIF4E complex crystal structures with Bn(7)GMP and FBn(7)GMP show that additional favourable contacts of the benzyl groups with eIF4E contribute binding energy that compensates for loss of the beta and gamma-phosphates. The N(7)-benzyl groups pack into a hydrophobic pocket behind the two tryptophan side-chains that are involved in the cation-pi stacking interaction between the cap and the eIF4E mononucleotide binding site. This pocket is formed by an induced fit in which one of the tryptophan residues involved in cap binding flips through 180 degrees relative to structures with N(7)-methylated cap derivatives. This and other observations made here will be useful in the design of new families of eIF4E inhibitors, which may have potential therapeutic applications in cancer.  相似文献   

10.
The single mutation L30 K in the Hu-Yap65 WW domain increased the stability of the complex with the peptide GTPPPPYTVG (K(d)=40(+/-5) microM). Here we report the refined solution structure of this complex by NMR spectroscopy and further derived structure-activity relationships by using ligand peptide libraries with truncated sequences and a substitution analysis that yielded acetyl-PPPPY as the smallest high-affinity binding peptide (K(d)=60 microM). The structures of two new complexes with weaker binding ligands chosen based on these results (N-(n-octyl)-GPPPYNH(2) and Ac-PLPPY) comprising the wild-type WW domain of Hu-Yap65 were determined. Comparison of the structures of the three complexes were useful for identifying the molecular basis of high-affinity: hydrophobic and specific interactions between the side-chains of Y28 and W39 and P5' and P4', respectively, and hydrogen bonds between T37 (donnor) and P5' (acceptor) and between W39 (donnor) and T2' (acceptor) stabilize the complex.The structure of the complex L30 K Hu-Yap65 WW domain/GTPPPPYTVG is compared to the published crystal structure of the dystrophin WW domain bound to a segment of the beta-dystroglycan protein and to the solution structure of the first Nedd4 WW domain and its prolin-rich ligand, suggesting that WW sequences bind proline-rich peptides in an evolutionary conserved fashion. The position equivalent to T22 in the Hu-Yap65 WW domain sequence is seen as responsible for differentiation in the binding mode among the WW domains of group I.  相似文献   

11.
The partition coefficients (K(P)) of a series of single-chain and double-chain fluorescent amphiphiles, between solid ordered (P(beta') and L(beta)) and liquid disordered (L(alpha) of the type l(d)) lipid phases coexisting in the same lipid bilayer, was studied using steady-state fluorescence emission anisotropy. The single-chain amphiphiles were N-(7-nitrobenzoxa-2, 3-diazol-4-yl)-alkylamines, and the double-chain amphiphiles were N-(7-nitrobenzoxa-2, 3-diazol-4-yl)-phosphatidylethanolamines with chain lengths of 12-18 carbon atoms. Saturated 18-carbon alkyl/acyl chain compounds were also compared with Delta(9)-cis unsaturated chains of the same chain length. The fluorescence anisotropy of the probes was examined in lipid bilayers (multilamellar vesicles) prepared from an equimolar mixture of dilauroylphosphatidylcholine and distearoylphosphatidylcholine and studied as a function of temperature through the entire temperature range of coexistence of ordered gel phases and a disordered fluid phase in this system. The unsaturated chain amphiphiles partitioned exclusively into the fluid phase whenever this phase was present, as did the saturated chain amphiphiles with the shortest chains (C(12:0)), while K(P) ranges between 1 and 2, in favor of the L(beta) solid phase, for the amphiphiles with long saturated (C(18:0)) alkyl/acyl chains, with intermediate behavior for the intermediate chain lengths. All probes appeared to be totally excluded from P(beta') solid (gel) phases. The technique was also used to determine partitioning of some of the probes between coexisting liquid ordered (cholesterol-containing) (l(o)) and liquid disordered (l(d)) L(alpha) phases. In this case the ratio of signal amplitude to noise allowed us to obtain a qualitative, but not quantitative, measure of the phase partitioning of the probes. We conclude that the partitioning behavior of the probes examined between coexisting l(o) and l(d) phases is qualitatively similar to that observed between solid ordered and liquid disordered phases.  相似文献   

12.
The N-terminal domain of HIV-1 glycoprotein 41,000 (gp41) participates in viral fusion processes. Here, we use physical and computational methodologies to examine the secondary structure of a peptide based on the N terminus (FP; residues 1-23) in aqueous and detergent environments. (12)C-Fourier transform infrared (FTIR) spectroscopy indicated greater alpha-helix for FP in lipid-detergent sodium dodecyl sulfate (SDS) and aqueous phosphate-buffered saline (PBS) than in only PBS. (12)C-FTIR spectra also showed disordered FP conformations in these two environments, along with substantial beta-structure for FP alone in PBS. In experiments that map conformations to specific residues, isotope-enhanced FTIR spectroscopy was performed using FP peptides labeled with (13)C-carbonyl. (13)C-FTIR results on FP in SDS at low peptide loading indicated alpha-helix (residues 5 to 16) and disordered conformations (residues 1-4). Because earlier (13)C-FTIR analysis of FP in lipid bilayers demonstrated alpha-helix for residues 1-16 at low peptide loading, the FP structure in SDS micelles only approximates that found for FP with membranes. Molecular dynamics simulations of FP in an explicit SDS micelle indicate that the fraying of the first three to four residues may be due to the FP helix moving to one end of the micelle. In PBS alone, however, electron microscopy of FP showed large fibrils, while (13)C-FTIR spectra demonstrated antiparallel beta-sheet for FP (residues 1-12), analogous to that reported for amyloid peptides. Because FP and amyloid peptides each exhibit plaque formation, alpha-helix to beta-sheet interconversion, and membrane fusion activity, amyloid and N-terminal gp41 peptides may belong to the same superfamily of proteins.  相似文献   

13.
With the aim to investigate the mechanism of Cu(I) transport by Wilson ATPase (ATP7B), we have studied the interaction of the peptides 2K10p (CH(3)CO-Lys-Gly-Met-Thr-Cys-Ala-Ser-Cys-Val-His-Asn-Lys-CONH(2)), and 2K8p (CH(3)CO-Lys-Leu-Cys-Ile-Ala-Cys-Pro-Cys-Ser-Lys-CONH(2)), part of the sixth metal binding domain (WD6) and the sixth transmembrane segment (TM6) of Wilson ATPase, respectively, by means of CD, NMR spectroscopy and homology modeling. In addition, the interaction of Cu(I) with the 2K8p mutants 1s (CH(3)CO-Lys-Leu-Ser-Ile-Ala-Cys-Pro-Cys-Ser-Lys-CONH(2)), 2s (CH(3)CO-Lys-Leu-Cys-Ile-Ala-Ser-Pro-Cys-Ser-Lys-CONH(2)) and 3s (CH(3)CO-Lys-Leu-Cys-Ile-Ala-Cys-Pro-Ser-Ser-Lys-CONH(2)), containing two cysteines in various positions, have been studied with the same methods, in order to understand the role of each cysteine in copper binding. Our studies show that the three cysteine thiolates present in the 2K8p peptide sequence act mainly as bridging ligands for Cu(I) binding, and dithiothreitol acts as an important ligand in Cu(I) ligation by 2K10p and the 2K8p mutants. Formation of oligomeric species has been evidenced for all peptides except 2s. Shift of the equilibrium between the various oligomeric species has been accomplished by reducing the Cu(I):peptide ratio. Significant shifts of proline protons upon interaction with Cu(I) have been observed for all proline containing peptides implying a possible role of proline in facilitating Cu(I) binding. These findings have been further discussed with respect to the molecular basis of copper trafficking and intermolecular interactions.  相似文献   

14.
Tendamistat is a strong inhibitory protein of porcine pancreatic alpha-amylase (PPA) with a K(i) value of 0.2 nM. To develop potent alpha-amylase inhibitors, we synthesized six odd-length cyclic peptides (5-15 residues) and four even-length cyclic peptides (10 and 12 residues) having the inhibitory sequence of tendamistat. Their PPA inhibitory activities were evaluated, and, among them, the 11-residue cyclic peptide Ten(15-23) (K(i) = 0.27 microM) exhibited the strongest inhibitory activity (K(i) = 0.27-1.41 microM). To examine the effect of cyclic structure on PPA inhibition, ten linear peptides corresponding to the cyclic peptides were also synthesized, and their PPA inhibitory activities were evaluated (K(i) = 0.28-1.00 microM). Interestingly, the 11-residue linear peptide Ten(15-23) exhibited almost the same inhibitory activity (K(i) = 0.28 microM) as that of cyclic Ten(15-23). The results of a circular dichroism study indicated that stabilization of the beta-hairpin structure occurred only for cyclic Ten(15-23). Also, the results of proteolytic digestion experiments of the cyclic and linear Ten(15-23) peptides by trypsin and chymotrypsin suggested no differences in protease resistance between the cyclic and linear structures. Therefore, we demonstrated that both cyclic and linear peptides containing the inhibitory sequence of tendamistat exhibit potent PPA inhibitory activity.  相似文献   

15.
Sequence requirements of the ATP-binding site within the C-terminal nucleotide-binding domain (NBD2) of mouse P-glycoprotein were investigated by using two recombinantly expressed soluble proteins of different lengths and photoactive ATP analogues, 8-azidoadenosine triphosphate (8N(3)-ATP) and 2',3',4'-O-(2,4,6-trinitrophenyl)-8-azidoadenosine triphosphate (TNP-8N(3)-ATP). The two proteins, Thr(1044)-Thr(1224) (NBD2(short)) and Lys(1025)-Ser(1276) (NBD2(long)), both incorporated the four consensus sequences of ABC (ATP-binding cassette) transporters, Walker A and B motifs, the Q-loop, and the ABC signature, while differing in N-terminal and C-terminal extensions. Radioactive photolabeling of both proteins was characterized by hyperbolic dependence on nucleotide concentration and high-affinity binding with K(0.5)(8N(3)-ATP) = 36-37 microM and K(0.5)(TNP-8N(3)-ATP) = 0.8-2.6 microM and was maximal at acidic pH. Photolabeling was strongly inhibited by TNP-ATP (K(D) = 0.1-5 microM) and ATP (K(D) = 0.5-2.7 mM). Since flavonoids display bifunctional interactions at the ATP-binding site and a vicinal steroid-interacting hydrophobic sequence [Conseil, G., Baubichon-Cortay, H., Dayan, G., Jault, J.-M., Barron, D., and Di Pietro, A. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 9831-9836], a series of 30 flavonoids from different classes were investigated for structure-activity relationships toward binding to the ATP site, monitored by protection against photolabeling. The 3-OH and aromaticity of conjugated rings A and C appeared important, whereas opening of ring C abolished the binding in all but one case. It can be concluded that the benzopyrone portion of the flavonoids binds at the adenyl site and the phenyl ring B at the ribosyl site. The Walker A and B motifs, intervening sequences, and small segments on both sides are sufficient to constitute the ATP site.  相似文献   

16.
A novel esterase gene (estI) of Lactobacillus casei CL96 was localized on a 3.3-kb BamHI DNA fragment containing an open reading frame (ORF) of 1,800 bp. The ORF of estI was isolated by PCR and expressed in Escherichia coli, the methylotrophic bacterium Methylobacterium extorquens, and the methylotrophic yeast Pichia pastoris under the control of T7, methanol dehydrogenase (P(mxaF)), and alcohol oxidase (AOX1) promoters, respectively. The amino acid sequence of EstI indicated that the esterase is a novel member of the GHSMG family of lipolytic enzymes and that the enzyme contains a lipase-like catalytic triad, consisting of Ser325, Asp516, and His558. E. coli BL21(DE3)/pLysS containing estI expressed a novel 67.5-kDa protein corresponding to EstI in an N-terminal fusion with the S. tag peptide. The recombinant L. casei CL96 EstI protein was purified to electrophoretic homogeneity in a one-step affinity chromatography procedure on S-protein agarose. The optimum pH and temperature of the purified enzyme were 7.0 and 37 degrees C, respectively. Among the pNP (p-nitrophenyl) esters tested, the most selective substrate was pNP-caprylate (C(8)), with K(m) and k(cat) values of 14 +/- 1.08 microM and 1,245 +/- 42.3 S(-1), respectively.  相似文献   

17.
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   

18.
We previously reported that a small peptide based on amino acids 143-153 of the c-Jun N-terminal kinase (JNK)-binding domain of JIP-1 functioned as an in vitro inhibitor of JNK activity. This peptide (TI-JIP: RP-KRPTTLNLF) resembles the kinase-interaction motif (KIM = (K/R)(2-3)X(1-6)(L/I)X(L/I)), which is common to upstream activators, downstream substrates, phosphatases, and scaffold proteins present in MAPK cascades. In this study, we characterized the mechanism of JNK inhibition by this peptide and further investigated the biochemical features of this peptide resulting in potent JNK inhibition. We also tested various KIM-based peptides for their ability to inhibit JNK activity. TI-JIP was found to be competitive with respect to the phosphoacceptor substrate c-Jun (K(I) = 0.39 +/- 0.08 microm), and exhibit mixed (non-competitive) inhibition with respect to ATP. All seven substitutions of Pro-5 we tested significantly reduced the JNK inhibition, as did altering the Pro-5 to Leu-8 spacing. When we independently tested eight substitutions of either Thr-6 or Thr-7, only one substitution in each position was well tolerated. Furthermore, peptides based on the KIMs from other proteins were significantly less potent JNK inhibitors than TI-JIP, including a peptide from the JNK interactor Sab that contained all critical inhibitory residues present in TI-JIP. Therefore, despite having previously identified Arg-4, Pro-5, Leu-8, and Leu-10 in TI-JIP as independently critical for mediating JNK inhibition, we find their presence in other 11-mer peptides is not sufficient for JNK inhibition. TI-JIP is therefore a unique KIM-based inhibitor of JNK activity.  相似文献   

19.
Ward DG  Taylor M  Lilley KS  Cavieres JD 《Biochemistry》2006,45(10):3460-3471
ATP has high- and low-affinity effects on the sodium pump and other P-type ATPases. We have approached this question by using 2',3'-O-(trinitrophenyl)-8-azidoadenosine 5'-diphosphate (TNP-8N(3)-ADP) to photoinactivate and label Na,K-ATPase, both in its native state and after covalent FITC block of its high-affinity ATP site. With the native enzyme, the photoinactivation rate constant increases hyperbolically with a K(D(TNP-8N)3(-)(ADP)) of 0.11 microM; TNP-ATP and ATP protect the site with high affinities. The inactivation does not require Na(+), but K(+) inhibits with a K(K)' of 12 microM; Na(+) reverses this effect, with a K(Na) of 0.17 mM. This pattern suggests that Na(+) and K(+) are binding at sites in their "intracellular" conformation. It was known that FITC did not abolish the reverse phosphorylation by P(i), or the K(+)-phosphatase activity, and that TNP-8N(3)-ADP could subsequently photoinactivate the latter with >100-fold lower affinity; in that case, the cation sites acted as if facing outward [Ward, D. G., and Cavieres, J. D. (1998) J. Biol. Chem. 273, 14277-14284, 33759-33765]. Native and FITC-modified enzymes have now been photolabeled with TNP-8N(3)-[alpha-(32)P]ADP and alpha-chain soluble tryptic peptides separated by reverse-phase HPLC. With native Na,K-ATPase, three labeled peaks lead to the unique sequence alpha-(470)Ile-Val-Glu-Ile-Pro-Phe-Asn-Ser-Thr-Asn-X-Tyr-Gln-Leu-Ser-Ile-His-Lys(487), the dropped residue being alphaLys480. With the FITC enzyme, instead, two independent labeling and purification cycles return the sequence alpha-(721)Ala-Asp-Ile-Gly-Val-Ala-Met-Gly-Ile-Ala-Gly-Ser-Asp-Val-Ser-Lys(736). These results suggest that Na,K-ATPase also has a low-affinity nucleotide binding region, one that is under distinctive allosteric control by Na(+) and K(+). Moreover, the cation effects seem compatible with a slow, passive Na(+)/K(+) carrier behavior of the FITC-modified sodium pump.  相似文献   

20.
Glutathione (GSH) is the major low-molecular weight antioxidant in mammalian cells. Thus, its analogues carrying similar and/or additional positive properties might have clinical perspectives. Here, we report the design and synthesis of a library of tetrapeptidic GSH analogues called UPF peptides. Compared to cellular GSH our designed peptidic analogues showed remarkably higher hydroxyl radical scavenging ability (EC(50) of GSH: 1,231.0 +/- 311.8 microM; EC(50) of UPF peptides: from 0.03 to 35 microM) and improved antiradical efficiency towards a stable alpha,alpha-diphenyl-beta-picrylhydrazyl (DPPH) radical. The best of UPF peptides was 370-fold effective hydroxyl radical scavengers than melatonin (EC(50): 11.4 +/- 1.0 microM). We also found that UPF peptides do not influence the viability and membrane integrity of K562 human erythroleukemia cells even at 200 microM concentration. Dimerization of GSH and UPF peptides was compared in water and in 0.9% saline solutions. The results, together with an earlier finding that UPF1 showed protective effects in global cerebral ischemia model in rats, suggest that UPF peptides might serve both as potent antioxidants as well as leads for design of powerful non-peptidic antioxidants that correct oxidative stress-driven events.  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