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 15 毫秒
1.
0.15 M inorganic phosphate dramatically increased the alpha-helix content of melittin in aqueous solution. When melittin interacted with egg yolk phosphatidylcholine liposomes in the absence of inorganic phosphate, it was converted to an alpha-helix rich form, as postulated by Dawson et al. (Dawson, C.R., Drake, A.F. Helliwell, J. and Hider, R.C. (1978) Biochim. Biophys. Acta 510, 75--86).  相似文献   

2.
Temperature dependence of the α-helix conformation of bee venom melittin in methanol-water mixed solvents has been examined by NMR, in order to elucidate conformation stability and a phase diagram. At high methanol concentration of 100 - ca. 80 wt.%, melittin forms a full α-helix conformation in the temperature range from 25 °C to 60 °C. At intermediate methanol concentration of ca. 80 - ca. 25 wt.%, it undergoes a thermal transformation from a full α-helix to a partial α-helix. In solutions of low methanol concentrations of ca. 25 - 0 wt.%, partial α-helix monomers and their self-aggregated conformers coexist at low temperatures, and the relative number of the monomers increases with increase in temperature. The monomers turn to a random coil state at high temperatures only below ca. 10 wt. % methanol concentrations. The thermal transitions are discussed from the viewpoint of stability of intra-molecular hydrogen bonds and inter-molecular hydrophobic interactions.  相似文献   

3.
S Kubota  J T Yang 《Biopolymers》1986,25(8):1493-1504
The conformation of melittin, a surface-active polypeptide, in solution was studied by CD spectra between 190 and 240 nm. The molecule was essentially unordered (possibly with a trace of helix) in water without salt at neutral pH. Upon deprotonation of four of the six cationic groups at pH 12 the polypeptide became partially helical (about 35%). The addition of NaDodSO4 to an aqueous melittin solution first caused the solution to become turbid but it became clear again in excess surfactant solution. The conformational changes depended on the molar NaDodSO4/melittin ratio, R. With R from 2.34 to 23.4, the melittin solution was turbid and the polypeptide conformation was probably a mixture of α-helix and β-sheets. This was supported by the ir spectrum of the turbid solution, which indicated the presence of both conformations. With R = 46.8 or 468 (1 or 10 mM NaDodSO4) the polypeptide conformation was characteristic of an α-helix, about 70–80% of the molecule, regardless of whether the surfactant was above or below its critical micelle concentration. This compared well with the x-ray results of 92% helix in crystals. The lower helicity of melittin in NaDodSO4 solution might be attributed to the end effects that destabilize the first and last turn of an helix at its N- and C-terminus, respectively.  相似文献   

4.
We have examined the kinetics of the adsorption of melittin, a secondary amphipathic peptide extracted from bee venom, on lipid membranes using three independent and complementary approaches. We probed (i) the change in the polarity of the 19Trp of the peptide upon binding, (ii) the insertion of this residue in the apolar core of the membrane, measuring the 19Trp-fluorescence quenching by bromine atoms attached on lipid acyl chains, and (iii) the folding of the peptide, by circular dichroism (CD). We report a tight coupling of the insertion of the peptide with its folding as an α-helix. For all the investigated membrane systems (cholesterol-containing, phosphoglycerol-containing, and pure phosphocholine bilayers), the decrease in the polarity of 19Trp was found to be significantly faster than the increase in the helical content of melittin. Therefore, from a kinetics point of view, the formation of the α-helix is a consequence of the insertion of melittin. The rate of melittin folding was found to be influenced by the lipid composition of the bilayer and we propose that this was achieved by the modulation of the kinetics of insertion. The study reports a clear example of the coupling existing between protein penetration and folding, an interconnection that must be considered in the general scheme of membrane protein folding.  相似文献   

5.
The effect of the bee toxin melittin on DMPC dynamics in fast-tumbling bicelles has been investigated. The 13C R1 and 13C-1H NOE relaxation parameters for DMPC were used to monitor the effect of melittin and cholesterol on lipid dynamics. It was found that melittin has the largest effect on the DMPC mobility in DMPC/DHPC bicelles, while less effect was observed in cholesterol-doped bicelles, or in bicelles made with CHAPS, indicating that the rigidity of the membrane affects the melittin-membrane interaction. CD spectra were analysed in terms of cooperativity of the α-helix to random coil transition in melittin, and these results also indicated similar differences between the bicelles. The study shows that bicelles can be used to investigate lipid dynamics by spin relaxation, and in particular of peptide-induced changes in membrane fluidity.  相似文献   

6.
The formation of α-helical assembly by complexing biologically active peptides with de novo designed protein is described. The de novo designed protein described here is a cystinelinked 4-helix bundle protein constructed with 80 amino acid residues and forms a hydrophobic core region surrounded by 4 helices in an aqueous solution. The biologically active peptides, such as melittin and human growth hormone releasing factor, contain the sequences that are able to form amphiphilic helices. These peptides alone do not form the α-helix structure in a diluted solution with low ion strength. But on mixing with the designed helix bundle protein, the peptides are strongly bound to the protein with the induction of α-helical structure in the biologically active peptides. The content of induced α-helix is in accord with that estimated from the amphiphilic sequence. The results mean that a novel architecture composed of α-helices is formed. Fluorescent and temperature-scanning measurement revealed that the α-helical assembly is constructed with hydrophobic interaction. Also, it is shown by means of fluorescence depolarization that the assembly has a compact globular form corresponding to 1 : 1 complex. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
The interferon-gamma inducible large GTPase human guanylate binding protein-1 (hGBP-1) plays a key role in anti-pathogenic and anti-proliferative functions. This protein hydrolyzes GTP to both GDP and GMP (predominant product) through sequential phosphate cleavages, which makes it functionally distinct from other GTPases. Previous study on truncated variants of hGBP-1 suggested that the α-helix present in the intermediate region is essential for dimerization and thus for GMP formation. However, the role of this helix in the full-length protein in GMP formation is not clearly understood. Here, we present that substitution of the helix with a Gly-rich flexible (GGS)3 sequence in the full-length hGBP-1 (termed as linker protein) showed a drastic decrease in GMP formation. Unlike wild-type, the linker protein is not capable of undergoing substrate-induced dimerization and thereby transition state-induced tetramerization, suggesting the importance of the helix in oligomerization. Furthermore, we examined the effect of interactions between this helix and the α2-helix of the globular domain in GMP formation through mutational studies. The L118G mutation in the α2-helix showed a significantly reduced GMP formation. These results indicate that the interactions of the α-helix with the α2-helix are essential for enhanced GMP production. We propose that these interactions help in the oligomerization-assisted proper positioning of the catalytic machinery for efficient second phosphate cleavage. These findings thus provide a better understanding into the regulation of GMP formation in a large GTPase hGBP-1.  相似文献   

8.
9.
The structural properties and dynamic behavior of the antimicrobial peptide melittin in hydrophobic and polar environments have been investigated. The main characteristics of the secondary structure of melittin in different media have been analyzed and compared with the data on an ideal α-helix. It has been shown that melittin is an α-helix bent in the region of Pro14; the N-terminus of the peptide tends to unfold, while the C-terminal segment (residues 14–23) retains a helical structure for 20 ns of the simulation. 2,2,2-Trifluoroethanol molecules stabilize the helical structure of the peptide by lowering the dielectric constant of the environment and preferentially accumulating near particular sites of the polypeptide chain.  相似文献   

10.
Human brain relies on a steady supply of glucose as the source of fuel, and type I hexokinase is the major isozyme governing the introduction of glucose to glycolysis in the brain. One unique regulatory property associated with type I isozyme is the alleviation of product inhibition by inorganic phosphate which binds to the N-terminal half, and the conformational change induced by inorganic phosphate must be propagated to the active site in the C-terminal half. With a single interdomain α-helix as the only covalent connection between the N- and C-terminal halves, the question arises as what role the interdomain α-helix plays at the interdomain signal transduction. Two mutants were constructed in an attempt to answer this question. The first mutant, A464P/E465G, with a helix breaker embedded in the interdomain α-helix had a smaller magnitude of phosphate alleviation than the wild type. The second mutant, with an insertion of seven additional residues between Gln 466 and His 467, had this phosphate relief property further diminished. Neither mutant showed dramatic changes nor the other kinetic properties. It is speculated that the interdomain α-helix is important for keeping the proper non-covalent contact so that transmission of the conformational changes across the N- and C-terminal half boundary can be achieved.  相似文献   

11.
13C-nmr spectra of poly(β-benzyl L-aspartate) containing 13C-enriched [3-13C]L -alanine residues in the solid state were recorded by the cross polarization–magic angle spinning method, in order to elucidate the conformation-dependent 13C chemical shifts of L -alanine residues taking various conformations such as the antiparallel β-sheet, the right-handed α-helix, the left-handed α-helix, and the left-handed ω-helix forms obtained by appropriate treatment. The latter two conformations for L -alanine residues are achieved when L -alanine residues are incorporated into poly(β-benzyl L -aspartate). We found that the alanine Cβ carbon show significant 13C chemical shift displacement depending on conformational change, and gave the 13C chemical shift values at about 17 ppm for the left-handed ω-helix, 14 ppm for the left-handed α-helix, 15.5 ppm for the right-handed α-helix, and 21.0 ppm for the antiparallel β-sheet relative to tetramethylsilane.  相似文献   

12.
The molecular chaperone αA‐crystallin, mainly localized in the human ocular lens, is believed to protect the lens from opacification and cataract, by suppressing the aggregation of the other lens proteins. The present study provides structural and thermodynamic insights into the ability of human αA‐crystallin (HAA) to bind to its partially unfolded clients in the lens, using a small peptide, melittin from bee venom, as a model client. We characterized the thermodynamic parameters of the binding process between melittin and HAA through isothermal titration calorimetry (ITC), and found the binding to be endothermic and entropy‐driven. We identified the amino acids in melittin important for binding to HAA by saturation‐transfer difference (STD) nuclear magnetic resonance (NMR) experiments, and analysis of NMR line broadening upon titration of melittin with HAA. Our results suggest that hydrophobic residues Ile17 and Ile20 on the C‐terminal region of melittin are in close contact with HAA in the melittin‐HAA complex. Information obtained from NMR experiments was used to generate structural models of the melittin‐HAA complex by molecular docking with high‐ambiguity driven docking (HADDOCK). Structural models of the melittin‐HAA complex reveal important principles underlying the interaction of HAA with its clients.  相似文献   

13.
Wang L  Guo H  Reyes N  Lee S  Bortz E  Guo F  Sun R  Tong L  Deng H 《Journal of virology》2012,86(3):1348-1357
Epstein-Barr virus and Kaposi's sarcoma-associated herpesvirus are etiologically associated with several types of human malignancies. However, as these two human gammaherpesviruses do not replicate efficiently in cultured cells, the morphogenesis of gammaherpesvirus virions is poorly understood. Murine gammaherpesvirus 68 (MHV-68) provides a tractable model to define common, conserved features of gammaherpesvirus biology. ORF52 of MHV-68 is conserved among gammaherpesviruses. We have previously shown that this tegument protein is essential for the envelopment and egress of viral particles and solved the crystal structure of ORF52 dimers. To more closely examine its role in virion maturation, we performed immunoelectron microscopy of MHV-68-infected cells and found that ORF52 localized to both mature, extracellular virions and immature viral particles in the cytoplasm. ORF52 consists of three α-helices followed by one β-strand. To understand the structural requirements for ORF52 function, we constructed mutants of ORF52 and examined their ability to complement an ORF52-null MHV-68 virus. Mutations in conserved residues in the N-terminal α1-helix and C terminus, or deletion of the α2-helix, resulted in a loss-of-function phenotype. Furthermore, the α1-helix was crucial for the predominantly punctate cytoplasmic localization of ORF52, while the α2-helix was a key domain for ORF52 dimerization. Immunoprecipitation experiments demonstrated that ORF52 interacts with another MHV-68 tegument protein, ORF42; however, a single point mutation in R95 in the C terminus of ORF52 led to the loss of this interaction. Moreover, the homologues of MHV-68 ORF52 in Kaposi's sarcoma-associated herpesvirus and Epstein-Barr virus complement the defect in ORF52-null MHV-68 and interact with MHV-68 ORF52. Taken together, these data uncover the relationship between the α-helical structure and the molecular basis for ORF52 function. This is the first structure-based functional domain mapping study for an essential gammaherpesvirus tegument protein.  相似文献   

14.
Conformational changes of bovine α-lactalbumin in sodium dodecyl sulfate (SDS) solution were studied with the circular dichroism (CD) method using a dilute phosphate buffer ofpH 7.0 and ionic strength 0.014. The proportions of α-helix and β-structure in α-lactalbumin were 34% and 12%, respectively, in the absence of SDS. In the SDS solution, the helicity increased to 44%, while the β-structure disappeared. In order to verify the structural change from β-structure to α-helix, the moiety, assuming the β-structure in the α-lactalbumin, was isolated by a chymotryptic digestion. The structure of this α-lactalbumin fragment, Phe31-Ile59, was almost disordered. However, the fragment adopted a considerable amount of α-helical structure in the SDS solution. On the other hand, the tertiary structure of α-lactalbumin, detected by changes of CD in the near-ultraviolet region, began to be disrupted before the secondary structural change in the surfactant solution. Dodecyl sulfate ions of 80 mol were cooperatively bound to α-lactalbumin. Although the removal of the bound dodecyl sulfate ions was tried by the dialysis against the phosphate buffer for 5 days, 4 mol dodecyl sulfates remained per mole of the protein. The remaining amount agreed with the number of stoichiometric binding site, determined by the Scatchard plot, indicating that the stoichiometric binding was so tight.  相似文献   

15.
The hexapeptide Ser-Gly-Ala-Gly-Lys-Thr has been synthesized and characterized. It was designed as a minimal soluble peptide that would be likely to have the phosphate-binding properties observed in the P-loops of proteins that bind the β-phosphate of GTP or ATP. The β-phosphate in such proteins is bound by a combination of the side chain ε-amino group of the lysine residue plus the concavity formed by successive main chain peptide NH groups called a nest, which is favored by the glycines. The hexapeptide is shown to bind HPO(4) (2-) strongly at neutral pH. The affinities of the various ionized species of phosphate and hexapeptide are analyzed, showing that they increase with pH. It is likely the main chain NH groups of the hexapeptide bind phosphate in much the same way as the corresponding P-loop atoms bind the phosphate ligand in proteins. Most proteinaceous P-loops are situated at the N-termini of α-helices, and this observation has frequently been considered a key aspect of these binding sites. Such a hexapeptide in isolation seems unlikely to form an α-helix, an expectation in accord with the CD spectra examined; this suggests that being at the N-terminus of an α-helix is not essential for phosphate binding. An unexpected finding about the hexapeptide-HPO(4) (2-) complex is that the side chain ε-amino group of the lysine occurs in its deprotonated form, which appears to bind HPO(4) (2-) via an N···H-O-P hydrogen bond.  相似文献   

16.
The cytoplasmic N-terminal domain of the human ether-a-go-go related gene (hERG) K+ channel is critical for the slow deactivation kinetics of the channel. However, the mechanism(s) by which the N-terminal domain regulates deactivation remains to be determined. Here we show that the solution NMR structure of the N-terminal 135 residues of hERG contains a previously described Per-Arnt-Sim (PAS) domain (residues 26-135) as well as an amphipathic α-helix (residues 13-23) and an initial unstructured segment (residues 2-9). Deletion of residues 2-25, only the unstructured segment (residues 2-9) or replacement of the α-helix with a flexible linker all result in enhanced rates of deactivation. Thus, both the initial flexible segment and the α-helix are required but neither is sufficient to confer slow deactivation kinetics. Alanine scanning mutagenesis identified R5 and G6 in the initial flexible segment as critical for slow deactivation. Alanine mutants in the helical region had less dramatic phenotypes. We propose that the PAS domain is bound close to the central core of the channel and that the N-terminal α-helix ensures that the flexible tail is correctly orientated for interaction with the activation gating machinery to stabilize the open state of the channel.  相似文献   

17.
Leader DP  Milner-White EJ 《Proteins》2011,79(3):1010-1019
We prepared a set of about 2000 α-helices from a relational database of high-resolution three-dimensional structures of globular proteins, and identified additional main chain i ← i+3 hydrogen bonds at the ends of the helices (i.e., where the hydrogen bonding potential is not fulfilled by canonical i ← i+4 hydrogen bonds). About one-third of α-helices have such additional hydrogen bonds at the N-terminus, and more than half do so at the C-terminus. Although many of these additional hydrogen bonds at the C-terminus are associated with Schellman loops, the majority are not. We compared the dihedral angles at the termini of α-helices having or lacking the additional hydrogen bonds. Significant differences were found, especially at the C-terminus, where the dihedral angles at positions C2 and C1 in the absence of additional hydrogen bonds deviate substantially from those occurring within the α-helix. Using a novel approach we show how the structure of the C-terminus of the α-helix can emerge from that of constituent overlapping α-turns and β-turns, which individually show a variation in dihedral angles at different positions. We have also considered the direction of propagation of the α-helix using this approach. If one assumes that helices start as a single α-turn and grow by successive addition of further α-turns, the paths for growth in the N → C and C → N directions differ in a way that suggests that extension in the C → N direction is favored.  相似文献   

18.
《Biophysical chemistry》1986,25(2):201-213
Interactions between B-DNA and homopolymeric α-helices of glycine, alanine, serine, asparagine and aspartic acid have been studied theoretically. The complexation energy has been minimised taking into account the interactions between DNA and the polypeptides as well as the internal energy of the α-helix and the interaction energy of counterions with the complex. The results obtained indicate the important role of strong hydrogen bonds between the peptide side chains and nucleic acid phosphate groups, these bonds being much stronger than specific interactions with the base-pairs. The formation of these structural bonds depends on the size of the α-helix, which in turn determines whether bridging across the major groove is possible. The steric role of the methyl group of thymine in orienting the peptide helix and the role of DNA screening cations in complex stabilization are also significant.  相似文献   

19.
Poly-L -lysine exists as an α-helix at high pH and a random coil at neutral pH. When the α-helix is heated above 27°C, the macromolecule undergoes a conformational transition to a β-sheet. In this study, the stability of the secondary structure of poly-L -lysine in solutions subjected to shear flow, at temperatures below the α-helix to β-sheet transition temperature, were examined using Raman spectroscopy and CD. Solutions initially in the α-helical state showed time-dependent increases in viscosity with shearing, rising as much as an order of magnitude. Visual observation and turbidity measurements showed the formation of a gel-like phase under flow. Laser Raman measurements demonstrated the presence of small amounts of β-sheet structure evidenced by the amide I band at 1666 cm−1. CD measurements indicated that solutions of predominantly α-helical conformation at 20°C transformed into 85% α-helix and 15% β-sheet after being sheared for 20 min. However, on continued shearing the content of β-sheet conformation decreased. The observed phenomena were explained in terms of a “zipping-up” molecular model based on flow enhanced hydrophobic interactions similar to that observed in gel-forming flexible polymers. © 1998 John Wiley & Sons, Inc. Biopoly 45: 239–246, 1998  相似文献   

20.
Cellobiose phosphorylase (CBP) catalyzes the reversible phosphorolysis of cellobiose to produce α-D-glucopyranosyl phosphate (Glc1P) and D-glucose. It is an essential enzyme for the metabolism of cello-oligosaccharides in a ruminal bacterium, Ruminococcus albus. In this study, recombinant R. albus CBP (RaCBP) produced in Escherichia coli was characterized. It showed highest activity at pH 6.2 at 50 °C, and was stable in a pH range of 5.5-8.8 and at below 40 °C. It phosphorolyzed only cellobiose efficiently, and the reaction proceeded through a random-ordered bi bi mechanism, by which inorganic phosphate and cellobiose bind in random order and D-glucose is released before Glc1P. In the synthetic reaction, RaCBP showed highest activity to D-glucose, followed by 6-deoxy-D-glucose. D-Mannose, 2-deoxy-D-glucose, D-glucosamine, D-xylose, 1,5-anhydro-D-glucitol, and gentiobiose also served as acceptors, although the activities for them were much lower than for D-glucose. D-Glucose acted as a competitive-uncompetitive inhibitor of the reverse synthetic reaction, which bound not only the Glc1P site (competitive) but also the ternary enzyme-Glc1P-D-glucose complex (uncompetitive).  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