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 93 毫秒
1.
The effect of 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE) in mixed membranes with 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) on interaction with a class A amphipathic peptide, Ac-DWLKAFYDKVAEKLKEAF-NH2 (Ac-18A-NH2), was investigated. The fluorescence lifetime of 2-(9-anthroyloxy)stearic acid and 2H NMR spectra were used to evaluate the penetration of water molecules into the membrane interface and the order of lipid acyl chains, respectively. The results demonstrated that DOPE in the mixed membranes decreased the fluorescence lifetime and increased the acyl-chain order, and that Ac-18A-NH2 affected them more for membranes with higher DOPE fractions. The partition coefficient (Kp) of the peptide to the mixed membranes was increased with the increase in the DOPE mole fractions. From the temperature dependence of the Kp values, the binding of Ac-18A-NH2 to POPC/DOPE mixed membranes was found to be entropy-driven. The formation of an α-helix at the membrane’s surface is supposed to induce positive curvature strain, which decreases the headgroup hydration and acyl-chain order of lipids. Thus, the binding of Ac-18A-NH2 to membranes is entropically more favorable at higher DOPE fractions since the peptide’s insertion into the membrane can decrease the order parameter and unfavorable headgroup hydration, which explains the enhanced peptide binding.  相似文献   

2.
An amphipathic class A peptide, Ac-18A-NH2, has been employed in modeling the -helical lipid-binding site of apolipoprotein A-I (apoA-I). To gain insight into the nature of protein–lipid interactions responsible for the ability of apoA-I to promote the efflux of intracellular cholesterol, the peptide disposition in model membranes composed of phosphatidylcholine (PC) and its mixture with cholesterol (Chol) has been characterized. By examining resonance energy transfer between the peptide Trp as a donor and anthrylvinyl-labeled PC as an acceptor it was found that Chol inclusion is conducive to shallower bilayer location of the Ac-18A-NH2 -helix. The limits for the Trp distance from the membrane center were estimated to be 1.5–1.7 nm (PC) and 1.9–2.1 nm (PC:Chol), indicating that in the PC bilayer the Trp resides at the level of the glycerol backbone and carbonyl groups while the region of the phosphocholine moieties is preferable for Trp location in the PC:Chol bilayer. These findings suggest that Chol can modulate the interactions between apoA-I and membrane lipids via reducing the depth of -helix bilayer penetration.Abbreviations apoA-I apolipoprotein A-I - AV-PC anthrylvinyl-labeled phosphatidylcholine - Chol cholesterol - HDL high-density lipoproteins - LUV large unilamellar vesicles - PC phosphatidylcholine - RET fluorescence resonance energy transfer  相似文献   

3.
Class A amphipathic helical peptides have been shown to mimic apolipoprotein A-I, the major protein component of high density lipoproteins and have been shown to inhibit atherosclerosis in several dyslipidemic mouse models. Previously we reported the NMR structure of Ac-18A-NH2, the base-line model class A amphipathic helical peptide in a 50% (v/v) trifluoroethanol-d3/water mixture, a membrane-mimic environment (Mishra, V. K., Palgunachari, M. N., Anantharamaiah, G. M., Jones, M. K., Segrest, J. P., and Krishna, N. R. (2001) Peptides 22, 567-573). The peptide Ac-18A-NH2 forms discoidal nascent high density lipoprotein-like particles with 1,2-dimyristoyl-sn-glycero-3-phosphocholine. Because subtle structural changes in the peptide.lipid complexes have been shown to be responsible for their antiatherogenic properties, we undertook high resolution NMR studies to deduce detailed structure of recombinant peptide.1,2-dimyristoyl-sn-glycero-3-phosphocholine complexes. The peptide adopts a well defined amphipathic alpha helical structure in association with the lipid at a 1:1 peptide:lipid weight ratio. Nuclear Overhauser effect spectroscopy revealed a number of intermolecular close contacts between the aromatic residues in the hydrophobic face of the helix and the lipid acyl chain protons. The pattern of observed peptide-lipid nuclear Overhauser effects is consistent with a parallel orientation of the amphipathic alpha helix, with respect to the plane of the lipid bilayer, on the edge of the disc (the belt model). Based on the results of chemical cross-linking and molecular modeling, we propose that peptide helices are arranged in a head to tail fashion to cover the edge of the disc. This arrangement of peptides is also consistent with the pKa values of the Lys residues determined previously. Taken together, these results provide for the first time a high resolution structural view of the peptide.lipid discoidal complexes formed by a class A amphipathic alpha helical peptide.  相似文献   

4.
The aim of the present investigation is to determine the effect of α-helical propensity and sidechain hydrophobicity on the stability of amphipathic α-helices. Accordingly, a series of 18-residue amphipathic α-helical peptides has been synthesized as a model system where all 20 amino acid residues were substituted on the hydrophobic face of the amphipathic α-helix. In these experiments, all three parameters (sidechain hydrophobicity, α-helical propensity and helix stability) were measured on the same set of peptide analogues. For these peptide analogues that differ by only one amino acid residue, there was a 0.96 kcal/mole difference in α-helical propensity between the most (Ala) and the least (Gly) α-helical analogue, a 12.1-minute difference between the most (Phe) and the least (Asp) retentive analogue on the reversed-phase column, and a 32.3°C difference in melting temperatures between the most (Leu) and the least (Asp) stable analogue. The results show that the hydrophobicity and α-helical propensity of an amino acid sidechain are not correlated with each other, but each contributes to the stability of the amphipathic α-helix. More importantly, the combined effects of α-helical propensity and sidechain hydrophobicity at a ratio of about 2:1 had optimal correlation with α-helix stability. These results suggest that both α-helical propensity and sidechain hydrophobicity should be taken into consideration in the design of α-helical proteins with the desired stability.  相似文献   

5.
D J Tobias  C L Brooks 《Biochemistry》1991,30(24):6059-6070
We used molecular dynamics simulations to study the folding/unfolding of one of turn of an alpha helix in Ac-(Ala)3-NHMe and Ac-(Val)3-NHMe. Using specialized sampling techniques, we computed free energy surfaces as functions of a conformational coordinate that corresponds to alpha helices at small values and to extended conformations at large values. Analysis of the peptide conformations populated during the simulations showed that alpha helices, reverse turns, and extended conformations correspond to minima on the free energy surfaces of both peptides. The free energy difference between alpha helix and extended conformations, determined from the equilibrium constants for helix unfolding, is approximately -1 kcal/mol for Ac-(Ala)3-NHMe and -5 kcal/mol for Ac-(Val)3-NHMe. The mechanism observed in our simulations, which includes reverse turns as important intermediates along the helix folding/unfolding pathway, is consistent with a mechanism proposed previously. Our results predict that both peptides (but especially the Ala peptide) have a much larger equilibrium constant for helix initiation than is predicted by the helix-coil transition theory with the host-guest parameters. We also predict a much greater difference in the equilibrium constants than the theory predicts. Insofar as helix initiation is concerned, our results suggest that the large difference between the helical propensities of Ala and Val cannot be explained by simple concepts such as side-chain rotamer restriction or unfavorable steric interactions. Rather, the origin of the difference appears to be quite complicated because it involves subtle differences in the solvation of the two peptides. The two peptides have similar turn-extended equilibria but very different helix-turn equilibria, and the difference in helical propensities reflects the fact that the helix-turn equilibrium strongly favors the turns in Ac-(Val)3-NHMe, while it favors the helices in Ac-(Ala)3-NHMe. We also computed thermodynamic decompositions of the free energy surfaces, and these revealed that the helix-turn equilibria are vastly different primarily because the changes in peptide-water interactions that accompany helix-to-turn conformational changes are qualitatively different for the two peptides.  相似文献   

6.
Lysine-based amphipathic nonapeptides, including homochiral peptides [Ac-(l-Lys-l-Lys-Xaa)3-NH2 (Xaa = Gly, Ala, Aib, Ac5c, or Ac6c) and Ac-(d-Lys-d-Lys-Aib)3-NH2], a heterochiral peptide [Ac-(l-Lys-d-Lys-Aib)3-NH2], and a racemic mixture of diastereomeric peptides [Ac-(rac-Lys-rac-Lys-Aib)3-NH2] were designed and synthesized to investigate the relationship between their preferred secondary structures and their antimicrobial activity. Peptide 5, [Ac-(l-Lys-l-Lys-Ac6c)3-NH2] formed a stable α-helical structure and exhibited strong activity against Gram-negative bacteria (Escherichia coli and Pseudomonas aeruginosa).  相似文献   

7.
The conformational behaviour of delta Ala has been investigated by quantum mechanical method PCILO in the model dipeptide Ac-delta Ala-NHMe and in the model tripeptides Ac-X-delta Ala-NHMe with X = Gly, Ala, Val, Leu, Abu and Phe and is found to be quite different. The computational results suggest that in the model tripeptides the most stable conformation corresponds to phi 1 = -30 degrees, psi 1 = 120 degrees and phi 2 = psi 2 = 30 degrees in which the > C = 0 of the acetyl group is involved in hydrogen bond formation with N-H of the amide group. Similar results were obtained for the conformational behaviour of D-Ala in Ac-D-Ala-NHMe and Ac-Ala-D-Ala-NHMe. The conformational behaviour of the amino acids delta Ala, D-Ala, Val and Aib in model tripeptides have been utilized in the designing of left handed helical peptides. It is shown that the peptide HCO-(Ala-D-Ala)3-NHMe can adopt both left and right handed helix whereas in the peptide Ac-(Ala-delta Ala)3-NHMe the lowest energy conformer is beta-bend ribbon structure. Left handed helical structure with phi = 30 degrees, psi = 60 degrees for D-Ala residues and phi = psi = 30 degrees for delta Ala is found to be more stable by 4 kcal mole-1 than the corresponding right handed helical structure for the peptide Ac-(D-Ala-delta Ala)3-NHMe. In both the peptides Ac-(Val-delta Ala)3-NHMe and Ac-(D-Val-delta Ala)3-NHMe the most stable conformer is the left handed helix. Comparisons of results for Ac-(Ala-delta Ala)3-NHMe and Ac(Val-delta Ala)3-NHMe and Ac-(D-Ala-delta Ala)3-NHMe and Ac-(D-Val-delta Ala)3-NHMe also reveal that the Val residues facilitate the population of 3(10) left handed helix over the other conformers. It is also shown that the conformational behaviour of Aib residue depends on the chirality of neighbouring amino acids, i.e. Ac-(Aib-Ala)3-NHMe adopts right handed helical structure whereas Ac-(Aib-D-Ala)3-NHMe is found to be in left handed helical structure.  相似文献   

8.
We have shown that Ac-hE18A-NH2, a dual-domain cationic apolipoprotein-mimetic peptide, reduces plasma cholesterol levels in dyslipidemic mice. Two single-domain cationic peptides based on the lytic class L peptide 18L were developed to test the hypothesis that a single-domain cationic amphipathic peptide can reduce atherosclerosis in apolipoprotein (apo)E null mice when orally administered. To incorporate anti-inflammatory properties, aromatic residues were clustered in the nonpolar face similar to peptide 4F, resulting in modified 18L (m18L). To reduce lytic properties, the Lys residues of 18L were replaced with Arg with the resulting peptide called modified R18L (mR18L). Biophysical studies showed that mR18L had stronger interactions with lipids than did m18L. Peptide mR18L was also more effective than m18L in promoting LDL uptake by HepG2 cells. ApoE null mice received normal chow or chow containing m18L or mR18L for six weeks. A significant reduction in plasma cholesterol and aortic sinus lesion area was seen only in the mR18L group. Plasma from mice administered mR18L, unlike those from the control and m18L groups, did not enhance monocyte adhesion to endothelial cells. Thus oral administration of mR18L reduces plasma cholesterol and lesion formation and inhibits monocyte adhesion.  相似文献   

9.
The amino acid Aib predisposes a peptide to be helical with context‐dependent preference for either 310‐ or α‐ or a mixed helical conformation. Short peptides also show an inherent tendency to be unfolded. To characterize helical and unfolded states adopted by water‐soluble Aib‐containing peptides, the conformational preference of Ac‐Ala‐Aib‐Ala‐Lys‐Ala‐Aib‐Lys‐Ala‐Lys‐Ala‐Aib‐Tyr‐NH2 was determined by CD, NMR and MD simulations as a function of temperature. Temperature‐dependent CD data indicated the contribution of two major components, each an admixture of helical and extended/polyproline II structures. Both right‐ and left‐handed helical conformations were detected from deconvolution of CD data and 13C NMR experiments. The presence of a helical backbone, more pronounced at the N‐terminal, and a temperature‐induced shift in α‐helix/310‐helix equilibrium, more pronounced at the C‐terminal, emerged from NMR data. Starting from polyproline II, the N‐terminal of the peptide folded into a helical backbone in MD simulations within 5 ns at 60°C. Longer simulations showed a mixed‐helical backbone to be stable over the entire peptide at 5°C while at 60°C the mixed‐helix was either stable at the N‐terminus or occurred in short stretches through out the peptide, along with a significant population of polyproline II. Our results point towards conformational heterogeneity of water‐soluble Aib‐based peptide helices and the associated subtleties. The problem of analyzing CD and NMR data of both left‐ and right‐handed helices are discussed, especially the validity of the ellipticity ratio [θ]222/[θ]207, as a reporter of α‐/310‐ population ratio, in right‐ and left‐handed helical mixtures. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

10.
Many of the bilayer spanning segments of membrane transport proteins contain proline residues, and most of them are believed to occur in α-helical form. A proline residue in the middle of an α-helix is known to produce a bend in the helix, and recent studies have focused on characterizing such a bend at atomic level. In the present case, molecular dynamics (MD) studies are carried out on helix F model of bacteriorhodopsin (BR) Ace-(Ala)7-Trp-(Ala)2-Tyr-Pro-(Ala)2-Trp-(Ala)8-NHMe and compared with Ace-(Ala)7-Trp-(Ala)2-Tyr-(Ala)3-Trp-(Ala)8-NHMe in which the proline is replaced by alanine. The bend in the helix is characterized by structural parameters such as kink angle (α), wobble angle (θ), virtual torsion angle (ρ), and the hydrogen bond distance d (Op?3 … Np+1). The average values and the flexibility involved in these parameters are evaluated. The correlation among the bend related parameters are estimated. The equilibrium side chain orientations of tryptophan and tyrosine residues are discussed and compared with those found in the recently proposed model of bacteriorhodopsin. Finally, a detailed characterization of the bend in terms of secondary structures such as αI, αII and goniometric helices are discussed, which can be useful in the interpretation of the experimental results on the secondary structures of membrane proteins involving the proline residue. © 1993 Wiley-Liss, Inc.  相似文献   

11.
To better understand the structural determinants of the physical-chemical and the biological properties of Ac-18A-NH(2) (acetyl-AspTrpLeuLysAlaPheTyrAspLysValAlaGluLysLeuLysGluAlaPhe-amide), we have determined its structure in 50% (v/v) trifluroethanol (TFE-d(3))/water mixture (5 mM potassium phosphate, pH 5.5, 310K) using two-dimensional proton NMR spectroscopy. Stereospecific assignments have been made for C(beta)H protons (all the residues except Ala and Val) and gammaCH(3) (Val) groups. Nuclear Overhauser effects are observed between the nonpolar side chains spaced at (i) and (i + 4) position in the primary sequence, e.g., Trp2 and Phe6, and Phe6 and Val10. This suggests that in addition to N-terminal acetyl and C-terminal amide groups, the amphipathic alpha helical structure of Ac-18A-NH(2) is further stabilized by interactions between the hydrophobic residues on the nonpolar face of the helix.  相似文献   

12.
The effects on protein stability of negatively charged Glu side chains at different positions along the length of the α-helix were investigated in the two-stranded α-helical coiled-coil. A native coiled-coil has been designed which consists of two identical 35 residue polypeptide chains with a heptad repeat QgVaGbAcLdQeKf and a Cys residue at position 2 to allow the formation of an interchain 2-2′ disulphide bridge. This coiled-coil contains no intra- or interchain electrostatic interactions and served as a control for peptides in which Glu was substituted for Gln in the e or g heptad positions. The effect of the substitutions on stability was determined by urea denaturation at 20°C with the degree of unfolding monitored by circular dichroism spectroscopy. A Glu substituted for Gln near the N-terminus in each chain of the coiled-coil stabilizes the coiled-coil at pH 7, consistent with the charge–helix dipole interaction model. This stability increase is modulated by pH change and the addition of salt (KCl or guanidine hydrochloride), confirming the electrostatic nature of the effect. In contrast, Glu substitution in the middle of the helix destabilizes the coiled-coil because of the lower helical propensity and hydrophobicity of Glu compared with Gln at pH 7. Taking the intrinsic differences into account, the apparent charge–helix dipole interaction at the N-terminus is approximately 0.35 kcal/mol per Glu substitution. A Glu substitution at the C-terminus destabilizes the coiled-coil more than in the middle owing to the combined effects of intrinsic destabilization and unfavourable charge–helix dipole interaction with the negative pole of the helix dipole. The estimated destabilizing charge–helix dipole interaction of 0.08 kcal/mol is smaller than the stabilizing interaction at the N-terminus. The presence of a 2-2′disulphide bridge appears to have little influence on the magnitude of the charge–helix dipole interactions at either end of the coiled-coil. © 1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Model class A amphipathic helical peptides mimic several properties of apolipoprotein A-I (apoA-I), the major protein component of high density lipoproteins. Previously, we reported the NMR structures of Ac-18A-NH(2) (renamed as 2F because of two phenylalanines), the base-line model class A amphipathic helical peptide in the presence of lipid ( Mishra, V. K., Anantharamaiah, G. M., Segrest, J. P., Palgunachari, M. N., Chaddha, M., Simon Sham, S. W., and Krishna, N. R. (2006) J Biol. Chem. 281, 6511-6519 ). Substitution of two Leu residues on the nonpolar face (Leu(3) and Leu(14)) with Phe residues produced the peptide 4F (so named because of four phenylalanines), which has been extensively studied for its anti-inflammatory and antiatherogenic properties. Like 2F, 4F also forms discoidal nascent high density lipoprotein-like particles with 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC). Since subtle structural changes in the peptide-lipid complexes have been shown to be responsible for their antiatherogenic properties, we undertook high resolution NMR studies to deduce detailed structure of 4F in 4F.DMPC discs. Like 2F, 4F adopts a well defined amphipathic alpha-helical structure in association with the lipid at a 1:1 peptide/lipid weight ratio. Nuclear Overhauser effect (NOE) spectroscopy revealed a number of intermolecular close contacts between the aromatic residues in the hydrophobic face of the helix and the lipid acyl chain protons. Similar to 2F, the pattern of observed peptide-lipid NOEs is consistent with a parallel orientation of the amphipathic alpha helix, with respect to the plane of the lipid bilayer, on the edge of the disc (the belt model). However, in contrast to 2F in 2F.DMPC, 4F in the 4F.DMPC complex is located closer to the lipid headgroup as evidenced by a number of NOEs between 4F and DMPC headgroup protons. These NOEs are absent in the 2F.DMPC complex. In addition, the conformation of the DMPC sn-3 chain in 4F.DMPC complex is different than in the 2F.DMPC complex as evidenced by the NOE between lipid 2.CH and betaCH(2) protons in 4F.DMPC, but not in 2F.DMPC, complex. Based on the results of this study, we infer that the antiatherogenic properties of 4F may result from its preferential interaction with lipid headgroups.  相似文献   

14.
Cholesterol (Chol) in phosphatidylcholine large unilamellar vesicles (PC LUV) modulated interaction of the bilayers with a class A amphipathic peptide, Ac-18A-NH2: Chol increased the peptide binding capacity and reduced the affinity together with the peptide-induced leakage of calcein from LUV. Similar effects of Chol have been observed on the interaction of LUV with apoA-I [Saito, H., Miyako, Y., Handa, T., and Miyajima, K. (1997) J. Lipid Res. 38, 287-294]. Circular dichroism (CD) spectra of the peptide indicated a similar helical structure formation in LUV with and without Chol. The fluorescence spectral shift, quantum yield, anisotropy, and acrylamide-quenching of the peptide Trp indicated that in PC:Chol (3:2) LUV, Ac-18A-NH2 was located in a more polar membrane environment with increased motional freedom and greater accessibility to the aqueous medium. Fluorescence energy transfer from the Trp indole ring to acceptors situated at different depths in the bilayers revealed that the amphipathic peptide penetrated the hydrophobic interior of PC bilayers, while the peptide was located at the polar zwitterionic surface in PC:Chol LUV. The inclusion of Chol causes the headgroup separation of PC at the surface of LUV and increases the binding maximum of the wedge-shaped amphipathic peptide without disrupting the membrane structure. In addition, the rigidifying effect of Chol on PC acyl chains prevents the penetration of the peptide into the bilayer interior. These findings imply that Chol in membranes affects the binding and motional freedom of exchangeable plasma apolipoproteins containing class A amphipathic sequences, e.g., apoA-I and apoCs.  相似文献   

15.
A series of 14 residue amphipathic α-helical peptides, in which the sidechains of glutamic acid and lysine have been covalently joined, was synthesized in order to determine the effect of spacing, position and orientation of these lactam bridges. It was found that although an (i, i+3) spacing would position the lactam bridge on the same face of the helix, these lactams with 18-member rings were actually helix-destabilizing regardless of position or location. On the other hand, (i, i+4) lactams with 21-member rings were helix-stabilizing but this was dependent on orientation. Glutamic acid-lysine lactams increased the helical content of the peptide when compared with their linear homologue in benign conditions (50 mM KH2PO4, 100 mM KCl, pH 7). Two Glu-Lys (i, i+4) lactams located at the N- and C-termini gave rise to a peptide with greater than 99% helical content in benign conditions. Peptides with Lys-Glu oriented lactams were random structures in benign conditions but in the presence of 50% TFE could be induced into a helical conformation. The stability of the single-stranded α-helices, as measured by thermal denaturations in 25% TFE indicated that Glu-Lys oriented lactam bridges stabilized the helical conformation relative to the linear unbridged peptide. One Glu-Lys lactam in the middle of the peptide was more effective at stabilizing helical structure than two Glu-Lys lactams positioned one at each end of the molecule. The lactams with the Lys-Glu orientation were destabilizing relative to the unbridged peptide. This study demonstrates that correct orientation and position of a lactam bridge is critical in order to design peptides with high helical content in aqueous media.  相似文献   

16.
The loop following helix α2 in glutathione transferase P1-1 has two conserved residues, Cys48 and Tyr50, important for glutathione (GSH) binding and catalytic activity. Chemical modification of Cys48 thwarts the catalytic activity of the enzyme, and mutation of Tyr50 generally decreases the kcat value and the affinity for GSH in a differential manner. Cys48 and Tyr50 were targeted by site-specific mutations and chemical modifications in order to investigate how the α2 loop modulates GSH binding and catalysis. Mutation of Cys48 into Ala increased KMGSH 24-fold and decreased the binding energy of GSH by 1.5 kcal/mol. Furthermore, the protein stability against thermal inactivation and chemical denaturation decreased. The crystal structure of the Cys-free variant was determined, and its similarity to the wild-type structure suggests that the mutation of Cys48 increases the flexibility of the α2 loop rather than dislocating the GSH-interacting residues. On the other hand, replacement of Tyr50 with Cys, producing mutant Y50C, increased the Gibbs free energy of the catalyzed reaction by 4.8 kcal/mol, lowered the affinity for S-hexyl glutathione by 2.2 kcal/mol, and decreased the thermal stability. The targeted alkylation of Cys50 in Y50C increased the affinity for GSH and protein stability. Characterization of the most active alkylated variants, S-n-butyl-, S-n-pentyl-, and S-cyclobutylmethyl-Y50C, indicated that the affinity for GSH is restored by stabilizing the α2 loop through positioning of the key residue into the lock structure of the neighboring subunit. In addition, kcat can be further modulated by varying the structure of the key residue side chain, which impinges on the rate-limiting step of catalysis.  相似文献   

17.
Alzheimer’s disease (AD) is a progressive neurodegenerative disorder that is characterized by loss of intellectual functioning of brain and memory loss. According to amyloid cascade hypothesis, aggregation of amyloid-β42 (Aβ42) peptide can generate toxic oligomers and their accumulation in the brain is responsible for the onset of AD. In spite of carrying out a large number of experimental studies on inhibition of Aβ42 aggregation by small molecules, the detailed inhibitory mechanism remains elusive. In the present study, comparable molecular dynamics (MD) simulations were performed to elucidate the inhibitory mechanism of a sulfonamide inhibitor C1 (2,5-dichloro-N-(4-piperidinophenyl)-3-thiophenesulfonamide), reported for its in vitro and in vivo anti-aggregation activity against Aβ42. MD simulations reveal that C1 stabilizes native α-helix conformation of Aβ42 by interacting with key residues in the central helix region (13–26) with hydrogen bonds and ππ interactions. C1 lowers the solvent-accessible surface area of the central hydrophobic core (CHC), KLVFF (16–20), that confirms burial of hydrophobic residues leading to the dominance of helical conformation in the CHC region. The binding free energy analysis with MM–PBSA demonstrates that Ala2, Phe4, Tyr10, Gln15, Lys16, Leu17, Val18, Phe19, Phe20, Glu22, and Met35 contribute maximum to binding free energy (?43.1 kcal/mol) between C1 and Aβ42 monomer. Overall, MD simulations reveal that C1 inhibits Aβ42 aggregation by stabilizing native helical conformation and inhibiting the formation of aggregation-prone β-sheet conformation. The present results will shed light on the underlying inhibitory mechanism of small molecules that show potential in vitro anti-aggregation activity against Aβ42.  相似文献   

18.
Four peptides capable of forming an amphipathic alpha-helix have been synthesized and their conformational and lipid-binding properties studied. These peptides have been designed to vary the alpha-helix-forming potential as well as the charge distribution of the model peptide. The resulting peptide analogs and their complexes with dimyristoyl phosphatidylcholine were studied by using right angle light scattering, negative stain electron microscopy, nondenaturing gradient gel electrophoresis, circular dichroism, intrinsic tryptophan fluorescence, and differential scanning calorimetry techniques. The four analogs, [Glu4,9, Leu11,17] (reverse-18A, [Glu4,9, Leu5,11,17] reverse-18A, [Glu1,8, Leu11,17] 18A, and [Glu1,8, Leu5,11,17] 18A were derived from a model amphipathic peptide Asp-Trp-Leu-Lys-Ala-Phe-Tyr-Asp-Lys-Val-Ala-Glu-Lys-Leu-Lys-Glu-Ala-Phe (18A) whose lipid-associating properties strongly mimic apolipoprotein A-I or derived from Lys-Trp-Leu-Asp-Ala-Phe-Tyr-Lys-Asp-Val-Ala-Lys-Glu-Leu-Glu-Lys-Ala-Phe (reverse-18A), a peptide with little affinity for lipid and having a reversed charge distribution compared to the 18A peptide. We have shown that by substituting glutamic acid and leucine for aspartic acid and alanine, respectively, in a weak lipid-associating amphipathic helix peptide, the lipid-associating ability can be increased. Thus, peptides with both kinds of charge distribution can associate with the lipid. The ability of the peptide to disrupt phospholipid bilayers, however, is higher for 18A analogs compared to the reverse-18A analogs even after increasing the helix-forming potential and hydrophobicity. In addition to forming smaller lipoprotein particles, the modified 18A analogs were much superior to the modified reverse-18A analogs in their ability to activate the enzyme lecithin:cholesterol acyltransferase. This demonstrates that the positions of charged residues in the amphipathic helix play an important role in lecithin:cholesterol acyltransferase activation.  相似文献   

19.
The amphipathic helix hypothesis for the lipid-associating domains of exchangeable plasma apolipoproteins has been further studied by analysis of the structure of the complexes formed between four synthetic peptide analogs of the amphipathic helix and dimyristoyl phosphatidylcholine (DMPC). Density gradient ultracentrifugation, negative stain electron microscopy, nondenaturing gradient gel electrophoresis, 1H NMR, high sensitivity differential scanning calorimetry, and circular dichroism were the techniques used in these studies. The two analogs Asp-Trp-Leu-Lys-Ala-Phe-Tyr-Asp-Lys-Val-Ala-Glu-Lys-Leu-Lys-Glu-Ala-Phe (18A) and 18A-Pro-18A whose sequences most strongly mimic native amphipathic sequences were found also most strongly to mimic apolipoprotein A-I in DMPC complex structure. The covalently linked dimer of the prototype amphipathic analog 18A, 18A-Pro-18A, appears to have greater lipid affinity than 18A. This presumably is the result of the cooperativity provided by two covalently linked lipid-associating domains in 18A-Pro-18A. The studies further suggest that the charge-reversed analog of the prototype 18A, reverse-18A, has the lowest lipid affinity of the four analogs studied and forms only marginally stable discoidal DMPC complexes. We postulate that this low lipid affinity is due predominantly, but not necessarily exclusively, to the lack of a hydrophobic contribution of lysine residues at the polar-nonpolar interface of reverse-18A versus 18A. The intermediate lipid affinity of des-Val10-18A, the fourth analog peptide, to produce a rank order of 18A-Pro-18A greater than 18A greater than des-Val10-18A greater than reverse-18A, supports this interpretation. Des-Val10-18A which has Val deleted from 18A has an amphipathic helical structure partially disrupted by the shift of 2 lysine residues away from the polar-nonpolar interface.  相似文献   

20.
The interaction of a model Lys flanked α-helical peptides K2-X24-K2, (X = A,I,L,L+A,V) with lipid bilayers composed of dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylcholine (DPPC) both, in a gel and in a liquid-crystalline state, has been studied by molecular dynamics simulations. It has been shown that these peptides cause disordering of the lipid bilayer in the gel state but only small changes have been monitored in a liquid-crystalline state. The peptides affect ordering of the surrounding lipids depending on the helix stability which is determined by amino acid side chains – their volume, shape, etc. We have shown that the helix does not keep the linear shape in all simulations but often bends or breaks. During some simulations with a very small difference between hydrophobic length of peptide and membrane thickness the peptide exhibits negligible tilt. At the same time changes in peptide conformations during simulations resulted in appearance of superhelix.  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