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.
As models of ion channel proteins and naturally occurring pore-forming peptides, we designed a series of Aib rich peptides [Ac-(Aib-Xxx-Aib-Ala)(5)-NH(2) (Xxx = Lys, Glu, Ser, and Gly: BXBA-20)] to investigate the effects of the side chains of the amino acid residues Lys, Glu, Ser, and Gly on the conformation and electrophysiological properties of ion channels. The conformation of peptides and their affinity for phospholipid membranes were evaluated by CD spectroscopy. Patch-clamp experiments revealed that all BXBA-20 peptides form ion channels in DPhPC bilayers exhibiting clearly resolved transitions between the open and closed states. The channel forming frequency was in the order BKBA-20>BEBA-20>BSBA-20>BGBA-20. In the case of BKBA-20 and BEBA-20, the self-assembled conductive oligomers expressed homogeneous and voltage-independent single channel conductances. In contrast, heterogeneous conductance was observed in BSBA-20 and BGBA-20 ion channels under similar experimental conditions. From these results, we conclude that peptides with a high degree of helical conformation, high amphipathicity, high affinity for lipid membranes, and self-associating characters in vesicles are most suitable for inducing ion channels with a high frequency of occurrence. Moreover, BEBA-20, BSBA-20, and BGBA-20 channels were cation-selective, whereas the BKBA-20 channel was non-selective.  相似文献   

2.
We made use of a planar lipid bilayer system to examine the action of synthetic basic peptides which model the prepiece moiety of mitochondrial protein precursors and have antibacterial activity against Gram-positive bacteria. The sequences of the peptides used were as follows: Ac-(Ala-Arg-Leu)3-NHCH3 (3(3], Ac-(Leu-Ala-Arg-Leu)2-NHCH3 (4(2], Ac-(Leu-Ala-Arg-Leu)3-NHCH3 (4(3], Ac-(Leu-Leu-Ala-Arg-Leu)2-NHCH3 (5(2]. These peptides interacted differently with planar lipid bilayer membranes and membrane conductance increased by the formation of ion channels. The effects of the peptides on the macroscopic current-increase and on the probability of channel formation, at the single channel level were in the order of 4(3) greater than 4(2) approximately 5(2) much greater than 3(3), a finding which correlates with the antibacterial activity of these peptides. The micromolar (microM) order concentration at which the channel was formed resembles that causing antibacterial activity. Thus, the peptide antibacterial activity may occur through an increase in ion permeability of the bacterial membrane. The single-channel properties were investigated in detail using 4(3), the peptide with the highest ion channel-forming activity. Many types of channels were observed with respect to conductance (2-750 pS) and voltage dependency of gating. However, the channels were all cation-selective. These results suggest that the ion channels formed by peptide 4(3) may be able to take on a variety of conformations and/or assembly.  相似文献   

3.
Basic amphipathic alpha-helical peptides Ac-(Leu-Ala-Arg-Leu)3 or 4-NHCH3 (4(3) or 4(4)) and H-(Leu-Ala-Arg-Leu)3-(Leu-Arg-Ala-Leu)2 or 3-OH (4(5) or 4(6)) were synthesized and studied in terms of their interactions with phospholipid membranes, biological activity, and ion channel-forming ability. CD study of the peptides showed that they form alpha-helical structures in the presence of phospholipid liposomes and thus they have amphipathic distribution of the side chains along the axis of the helix. A leakage study of carboxyfluorescein encapsulated in phospholipid vesicles indicated that the peptides possess a highly potent ability to perturb the membrane structure. Membrane current measurements using the planar lipid bilayer technique revealed that the peptide 4(6), which was long enough to span the lipid bilayer in the alpha-helical structure, formed cation-selective ion channels at a concentration of 0.5 microM in a planar diphytanoylphosphatidylcholine bilayer. In contrast, other shorter peptides failed to form discrete and stable channels though they occasionally induced an increase in the membrane current with erratic conductance levels. The probability of detecting a conductance increase was in the order of 4(6) greater than 4(5) greater than 4(4) greater than 4(3), which corresponds to the order of the peptide chain lengths. Furthermore, 4(6) but not 4(5) showed an antimicrobial activity against both Gram-positive and -negative bacteria. The structure of ion channels formed by 4(6) and the relationship between the peptide chain length and biological activity of the synthetic peptides are discussed.  相似文献   

4.
An ultimate goal of synthetic ion-channel peptide design is to construct stable and functional ion-conducting pores. It is expected that specific interhelical interactions would facilitate the association of helices in phospholipid membranes and the successive helix-bundle formation. In the present study, we rationally designed helix-bundle ion channels using the synthetic hybrid peptide K20E20, a disulfide dimer of cationic- and anionic-amphiphilic helices Ac-CGG-(BKBA) 5-NH 2 and Ac-CGG-(BEBA) 5-NH 2. Circular dichroism (CD) measurements in aqueous media implied helix stabilization in the peptide caused by the interhelical electrostatic interactions. In addition, CD spectra recorded in the presence of DPPC liposomes and dye-leakage measurements suggested a high degree of association of peptide monomers in phospholipid membranes as well as high affinities between peptide and lipid bilayers. These features allowed ion-channel formation at extremely low peptide concentrations (as low as 1 nM). According to electrophysiological analyses, stable helix bundles were constructed of six peptide helices by association of three K20E20 molecules. Helix-helix association in lipid membranes, peptide-membrane interactions, and ion-channel formation of K20E20 peptides were all facilitated by intramolecular electrostatic interactions between the helices of the hybrid peptide and were pH-dependent. Conductance through K20E20 ion channels decreased under acidic conditions because of the interruption of the salt bridges.  相似文献   

5.
Hydroxylation of proline residues in the Yaa position of the Gly-Xaa-Yaa repeated sequence to 4(R)-hydroxyproline is essential for the formation of the collagen triple helix. A small number of 3(S)-hydroxyproline residues are present in most collagens in the Xaa position. Neither the structural nor a biological role is known for 3(S)-hydroxyproline. To characterize the structural role of 3(S)-hydroxyproline, the peptide Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 was synthesized and analyzed by circular dichroism spectroscopy, analytical ultracentrifugation, and 1H nuclear magnetic resonance spectroscopy. At 4 degrees C in water the circular dichroism spectrum indicates that this peptide was in a polyproline-II-like secondary structure with a positive peak at 225 nm similar to Ac-(Gly-Pro-4(R)Hyp)10-NH2. The positive peak at 225 nm almost linearly decreases with increasing temperature to 95 degrees C without an obvious transition. Although the peptide Ac-(Gly-Pro-4(R)Hyp)10-NH2 forms a trimer at 10 degrees C, sedimentation equilibrium experiments indicate that Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 is a monomer in water at 7 degrees C. To study the role of 3(S)-hydroxyproline in the Yaa position, we synthesized Ac-(Gly-Pro-3(S)Hyp)10-NH2. This peptide also does not form a triple helix in water. 1H Nuclear magnetic resonance spectroscopy data (including line widths and nuclear Overhauser effects) are entirely consistent, with neither Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 nor Ac-(Gly-Pro-3(S)Hyp)10-NH2 forming a triple helix in water. Therefore 3(S)-hydroxyproline destabilizes the collagen triple helix in either position. In contrast, when 3(S)-hydroxyproline is inserted as a guest in the highly stable -Gly-Pro-4(R)Hyperepeated host sequence, Ac-(Gly-Pro-4(R)Hyp)3-Gly-3(S)Hyp-4(R)Hyp-(Gly-Pro-4(R)Hyp)4-Gly-Gly-NH2 forms as stable a trimer (Tm=49.6 degrees C) as Ac-(Gly-Pro-4(R)Hyp)8-Gly-Gly-NH2 (Tm=48.9 degrees C). Given that Ac-(Gly-Pro-4(R)Hyp)3-Gly-4(R)Hyp-Pro-(Gly-Pro-4(R)Hyp)4-Gly-Gly-NH2 forms a triple helix nearly as stable as the above two peptides (Tm=45.0 degrees C) and the knowledge that Ac-(Gly-4(R)Hyp-Pro)10-NH2 does not form a triple helix, we conclude that the host environment dominates the structure of host-guest peptides and that these peptides are not necessarily accurate predictors of triple helical stability.  相似文献   

6.
Bann JG  Bächinger HP  Peyton DH 《Biochemistry》2003,42(14):4042-4048
The glycopeptide Ac-(Gly-Pro-Thr(beta-Gal))(10)-NH(2) forms a collagen-like triple-helix. A (1)H NMR structural analysis is reported for the peptides Ac-(Gly-Pro-Thr)(n)-NH(2) and Ac-(Gly-Pro-Thr(beta-Gal))(n)-NH(2), where n = 1, 5, and 10. NMR assignments for the individual peptides are made using one- and two-dimensional TOCSY, ROESY, and NOESY experiments. The NMR and corroborating CD data show that Ac-(Gly-Pro-Thr)(n)-NH(2), n = 1, 5, or 10, as well as Ac-(Gly-Pro-Thr(beta-Gal))(n)-NH(2), n = 1 or 5 peptides are unable to form collagen-like triple-helical structures. Furthermore, the equilibrium ratio of cis to trans isomers of the Pro residues is unaffected by the presence of carbohydrate. For Ac-(Gly-Pro-Thr(beta-Gal))(10)-NH(2), the kinetics of amide (1)H exchange with solvent deuterium indicate a slow rate of exchange for both the Gly and the Thr amide. The data are thus consistent with a model in which the carbohydrate stabilizes the triple helix through an occlusion of water molecules and by hydrogen bonding but not through an influence on the cis to trans isomer ratio.  相似文献   

7.
The single-crystal structures of three collagen-like host-guest peptides, (Pro-Pro-Gly)(4) -Hyp-Yaa-Gly-(Pro-Pro-Gly)(4) [Yaa = Thr, Val, Ser; Hyp = (4R)-4-hydroxyproline] were analyzed at atomic resolution. These peptides adopted a 7/2-helical structure similar to that of the (Pro-Pro-Gly)(9) peptide. The stability of these triple helices showed a similar tendency to that observed in Ac-(Gly-Hyp-Yaa)(10) -NH(2) (Yaa = Thr, Val, Ser) peptides. On the basis of their detailed structures, the differences in the triple-helical stabilities of the peptides containing a Hyp-Thr-Gly, Hyp-Val-Gly, or Hyp-Ser-Gly sequence were explained in terms of van der Waals interactions and dipole-dipole interaction between the Hyp residue in the X position and the Yaa residue in the Y position involved in the Hyp(X):Yaa(Y) stacking pair. This idea also explains the inability of Ac-(Gly-Hyp-alloThr)(10) -NH(2) and Ac-(Gly-Hyp-Ala)(10) -NH(2) peptides to form triple helices. In the Hyp(X):Thr(Y), Hyp(X):Val(Y), and Hyp(X):Ser(Y) stacking pairs, the proline ring of the Hyp residues adopts an up-puckering conformation, in agreement with the residual preference of Hyp, but in disagreement with the positional preference of X in the Gly-Xaa-Yaa sequence.  相似文献   

8.
We have studied the fusion of small unilamellar vesicles composed of egg PC and of a mixture of egg PC plus egg PA using various basic amphipathic peptides. Fusion was monitored by carboxyfluorescein leakage assay, light scattering, membrane intermixing assay, contents mixing assay and electron microscopy. Ac-(L-Leu-L-Ala-L-Arg-L-Leu)3-NHCH3 (peptide 4(3] and Ac-(L-Leu-L-Ala-L-Lys-L-Leu)3-NHCH3 (peptide 4'3), which have high hydrophobic moments, caused transformation of small unilamellar vesicles into larger and relatively homogeneous ones. Ac-(L-Leu-L-Leu-L-Ala-L-Arg-L-Leu)2-NHCH3 (5(2], which has medium hydrophobic moment, induced weak but appreciable fusion, while Ac-(L-Ala-L-Arg-L-Leu)3-NHCH3 (3(3] which has no helical structure did not show any fusion. However, peptides 4(3), 4'3 and 5(2) caused massive leakage of the contents from small unilamellar vesicles. These results indicated that interaction of the peptides with artificial membranes caused extensive perturbation of the lipid bilayer, followed by fusion. The fusogenic capacity of model basic peptides was correlated with the hydrophobic moment of each peptide when the peptides adopted an alpha-helical structure in the presence of acidic liposomes. Peptides 4(3) and 4'3 also showed weak fusogenic ability for neutral liposomes, while 5(2) and 3(3) showed no ability, suggesting that highly amphipathic peptides, such as 4(3), interact weakly but distinctly with neutral liposomes to fuse them.  相似文献   

9.
The collagen triple helix is one of the most abundant protein motifs in animals. The structural motif of collagen is the triple helix formed by the repeated sequence of -Gly-Xaa-Yaa-. Previous reports showed that H-(Pro-4(R)Hyp-Gly)(10)-OH (where '4(R)Hyp' is (2S,4R)-4-hydroxyproline) forms a trimeric structure, whereas H-(4(R)Hyp-Pro-Gly)(10)-OH does not form a triple helix. Compared with H-(Pro-Pro-Gly)(10)-OH, the melting temperature of H-(Pro-4(R)Hyp-Gly)(10)-OH is higher, suggesting that 4(R)Hyp in the Yaa position has a stabilizing effect. The inability of triple helix formation of H-(4(R)Hyp-Pro-Gly)(10)-OH has been explained by a stereoelectronic effect, but the details are unknown. In this study, we synthesized a peptide that contains 4(R)Hyp in both the Xaa and the Yaa positions, that is, Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2) and compared it to Ac-(Gly-Pro-4(R)Hyp)(10)-NH(2), and Ac-(Gly-4(R)Hyp-Pro)(10)-NH(2). Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2) showed a polyproline II-like circular dichroic spectrum in water. The thermal transition temperatures measured by circular dichroism and differential scanning calorimetry were slightly higher than the values measured for Ac-(Gly-Pro-4(R)Hyp)(10)-NH(2) under the same conditions. For Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2), the calorimetric and the van't Hoff transition enthalpy DeltaH were significantly smaller than that of Ac-(Gly-Pro-4(R)Hyp)(10)-NH(2). We postulate that the denatured states of the two peptides are significantly different, with Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2) forming a more polyproline II-like structure instead of a random coil. Two-dimensional nuclear Overhauser effect spectroscopy suggests that the triple helical structure of Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2) is more flexible than that of Ac-(Gly-Pro-4(R)Hyp)(10)-NH(2). This is confirmed by the kinetics of amide (1)H exchange with solvent deuterium of Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2), which is faster than that of Ac-(Gly-Pro-4(R)Hyp)(10)-NH(2). The higher transition temperature of Ac-(Gly-4(R)Hyp-4(R)Hyp)(10)-NH(2), can be explained by the higher trans/cis ratio of the Gly-4(R)Hyp peptide bonds than that of the Gly-Pro bonds, and this ratio compensates for the weaker interchain hydrogen bonds.  相似文献   

10.
All known naturally occurring linear cationic peptides adopt an amphipathic alpha-helical conformation upon binding to lipids as an initial step in the induction of cell leakage. We designed an 18-residue peptide, (KIGAKI)3-NH2, that has no amphipathic character as an alpha-helix but can form a highly amphipathic beta-sheet. When bound to lipids, (KIGAKI)3-NH2 did indeed form a beta-sheet structure as evidenced by Fourier transform infrared and circular dichroism spectroscopy. The antimicrobial activity of this peptide was compared with that of (KIAGKIA)3-NH2, and it was better than that of GMASKAGAIAGKIAKVALKAL-NH2 (PGLa) and (KLAGLAK)3-NH2, all of which form amphipathic alpha-helices when bound to membranes. (KIGAKI)3-NH2 was much less effective at inducing leakage in lipid vesicles composed of mixtures of the acidic lipid, phosphatidylglycerol, and the neutral lipid, phosphatidylcholine, as compared with the other peptides. However, when phosphatidylethanolamine replaced phosphatidylcholine, the lytic potency of PGLa and the alpha-helical model peptides was reduced, whereas that of (KIGAKI)3-NH2 was improved. Fluorescence experiments using analogs containing a single tryptophan residue showed significant differences between (KIGAKI)3-NH2 and the alpha-helical peptides in their interactions with lipid vesicles. Because the data suggest enhanced selectivity between bacterial and mammalian lipids, linear amphipathic beta-sheet peptides such as (KIGAKI)3-NH2 warrant further investigation as potential antimicrobial agents.  相似文献   

11.
Bax and Bid are proapoptotic proteins of the Bcl-2 family that regulate the release of apoptogenic factors from mitochondria. Although they localize constitutively in the cytoplasm, their apoptotic function is exerted at the mitochondrial outer membrane, and is related to their ability to form transbilayer pores. Here we report the poration activity of fragments from these two proteins, containing the first alpha-helix of a colicinlike hydrophobic hairpin (alpha-helix 5 of Bax and alpha-helix 6 of Bid). Both peptides readily bind to synthetic lipid vesicles, where they adopt predominantly alpha-helical structures and induce the release of entrapped calcein. In planar lipid membranes they form ion conducting channels, which in the case of the Bax-derived peptide are characterized by a two-stage pattern, a large conductivity and lipid-charge-dependent ionic selectivity. These features, together with the influence of intrinsic lipid curvature on the poration activity and the existence of two helical stretches of different orientations for the membrane-bound peptide, suggest that it forms mixed lipidic/peptidic pores of toroidal structure. In contrast, the assayed Bid fragment shows a markedly different behavior, characterized by the formation of discrete, steplike channels in planar lipid bilayers, as expected for a peptidic pore lined by a bundle of helices.  相似文献   

12.
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   

13.
Conformational properties and interactions with lipid membranes were studied for the chemically synthesized peptides PC(1-37) and PC(1-43), corresponding to the N-terminal 37 and 43 residues, respectively, of the transit peptide of the precursor to plastocyanin of Silene pratensis. PC(1-43) covers the entire chloroplast targeting domain of the transit peptide. CD spectra of PC(1-37) and PC(1-43) showed that both peptides have little ordered structure in aqueous solutions but form partially helical conformations in the presence of detergent micelles or in methanol. Vesicle disruption and direct-binding experiments revealed, however, that neither PC(1-37) nor PC(1-43) had a high affinity for lipid membranes. Since in the intact plastocyanin transit peptide the chloroplast-targeting domain is followed by a hydrophobic thylakoid-transfer domain, the plastocyanin precursor may well be transported to the chloroplast surface first with the aid of the thylakoid-transfer domain. The chloroplast-targeting domain may then form a helical structure in the lipid environments, and a chloroplast-specific motif displayed on the helical structure may be recognized by a receptor protein located at the chloroplast envelope membranes.  相似文献   

14.
We compared the properties of two peptides of identical size and amino acid composition, Ac-(LKKL)(5)-NHEt and Ac-(KL)(10)-NHEt. Both are amphipathic, but only Ac-(LKKL)(5)-NHEt is a potent promoter of negative curvature. CD studies performed in the presence of lipids confirmed that under these conditions Ac-(LKKL)(5)-NHEt forms an alpha-helix, and Ac-(KL)(10)-NHEt adopts a beta structure. We studied their binding affinity by centrifugation and isothermal titration calorimetry techniques. The Ac-(LKKL)(5)-NHEt bound to zwitterionic and anionic liposomes, while Ac-(KL)(10)-NHEt interacted mainly with anionic liposomes. Ac-(LKKL)(5)-NHEt was more lytic than Ac-(KL)(10)-NHEt for zwitterionic palmitoyloleoylphosphatidylcholine (POPC) liposomes, and for liposomes composed of lipids extracted from either sheep or human erythrocytes (RBC). Both peptides had similar lytic and lipid mixing activities for liposomes containing anionic lipids. Both peptides were highly hemolytic, with Ac-(LKKL)(5)-NHEt active against sheep RBC and Ac-(KL)(10)-NHEt more active against human RBC. From their respective minimal effective concentrations (MECs) as antimicrobial agents, we judged Ac-(KL)(10)-NHEt to be 2 to 5-fold more potent than Ac-(LKKL)(5)-NHEt in media that contained physiological concentrations of NaCl. Notwithstanding, both peptides had MECs <1 microg/mL for Escherichia coli and Pseudomonas aeruginosa and <4 microg/mL for Staphylococcus aureus and methicillin-resistant Staphylococcus aureus. Although selectivity of antimicrobial peptides for bacterial membranes may result, in part, from the preferential display of anionic residues in these membranes, inability to interact with or bind to zwitterionic phospholipids offers no guarantee that the peptide will lack appreciable cytotoxicity for host cells.  相似文献   

15.
M E Hiltz  A Catania  J M Lipton 《Peptides》1991,12(4):767-771
D-Amino acid substitutions in the anti-inflammatory/antipyretic Ac-alpha-MSH(11-13)-NH2 tripeptide of Ac-alpha-MSH(1-13)-NH2 were made and the altered peptides were injected in mice treated with picryl chloride. Ear swelling, measured 3 and 6 h after application of the irritant, was reduced by IP injections of Ac-alpha-MSH(11-13)-NH2, in confirmation of previous observations. Ac-[D-Lys11]alpha-MSH(11-13)-NH2 effected similar anti-inflammatory activity but Ac-[D-Pro12]alpha-MSH(11-13)-NH2 was inactive. Ac-[D-Val13]alpha-MSH(11-13)-NH2 and Ac-[D-Lys11,D-Val13]alpha-MSH(11-13)-NH2 generally had greater anti-inflammatory activity than the parent tripeptide molecule; the dose-response relations exhibited the bell-shaped characteristics seen previously with MSH peptides. The results indicate that the L-Pro12 is essential for the anti-inflammatory activity of Ac-alpha-MSH(11-13)-NH2 whereas the L-Lys11 is not. D-Val13 substitution increased anti-inflammatory activity approximately four-fold over Ac-alpha-MSH(11-13)-NH2. These results provide new structure-activity relationships of the anti-inflammatory Ac-alpha-MSH(11-13)-NH2 molecule. The data support the developing idea that alpha-MSH and its COOH-terminal fragments modulate host responses, perhaps by antagonizing the actions of cytokines.  相似文献   

16.
The conformational behaviour of deltaZPhe has been investigated in the model dipeptide Ac-deltaZPhe-NHMe and in the model tripeptides Ac-X-deltaZPhe-NHMe with X=Gly,Ala,Val,Leu,Abu,Aib and Phe and is found to be quite different. In the model tripeptides with X=Ala,Val,Leu,Abu,Phe the most stable structure corresponds to phi1=-30 degrees, psi1=120 degrees and phi2=psi2=30 degrees. This structure is stabilized by the hydrogen bond formation between C=O of acetyl group and the NH of the amide group, resulting in the formation of a 10-membered ring but not a 3(10) helical structure. In the peptides Ac-Aib-deltaZPhe-NHMe and Ac-(Aib-deltaZPhe)3-NHMe, the helical conformers with phi = +/-30 degrees, psi = +/-60 degrees for Aib residue and phi=psi= +/-30 degrees for deltaZPhe are predicted to be most stable. The computational studies for the positional preferences of deltaZPhe residue in the peptide containing one deltaZPhe and nine Ala residues reveal the formation of a 3(10) helical structure in all the cases with terminal preferences for deltaZPhe. The conformational behaviour of Ac-(deltaZPhe)n-NHMe with n< or =4 is predicted to be very labile. With n > 4, degenerate conformational states with phi,psi values of 0 degrees +/- 90 degrees adopt helical structures which are stabilized by carbonyl-carbonyl interactions and the N-H-pi interactions between the amino group of every deltaZPhe residue with one C-C edge of its own phenyl ring. The results are in agreement with the experimental finding that screw sense of helix for peptides containing deltaZPhe residues is ambiguous in solution. The helical structures stabilized by hydrogen bond formation are found to be at least 3kCalmol(-1) less stable. Conformational studies have also been carried out for the peptide Ac-(deltaEPhe)6-NHMe and the peptide Ac-deltaAla-(deltaZPhe)6-NHMe containing deltaAla residue at the N-terminal. The N-H-pi interactions are absent in peptide Ac-(deltaEPhe)6-NHMe.  相似文献   

17.
Using molecular dynamics simulations, we studied the structure, interhelix interactions, and dynamics of transmembrane proteins. Specifically, we investigated homooligomeric helical bundle systems consisting of synthetic α-helices with either the sequence Ac-(LSLLLSL)3-NH2 (LS2) or Ac-(LSSLLSL)3-NH2 (LS3). The LS2 and LS3 helical peptides are designed to have amphipathic characteristics that form ion channels in membrane. We simulated bundles containing one to six peptides that were embedded in palmitoyl-oleoyl-phosphatidylcholine (POPC) lipid bilayer and placed between two lamellae of water. We aim to provide a fundamental understanding of how amphipathic helical peptides interact with each other and their dynamical behaviors in different homooligomeric states. To understand structural properties, we examined the helix lengths, tilt angles of individual helices and the entire bundle, interhelix distances, interhelix cross-angles, helix hydrophobic-to-hydrophilic vector projections, and the average number of interhelix hydrophilic (serine–serine) contacts lining the pore of the transmembrane channel. To analyze dynamical properties, we calculated the rotational autocorrelation function of each helix and the cross-correlation of the rotational velocity between adjacent helices. The observed structural and dynamical characteristics show that higher order bundles containing four to six peptides are composed of multiple lower order bundles of one to three peptides. For example, the LS2 channel was found to be stable in a tetrameric bundle composed of a “dimer of dimers.” In addition, we observed that there is a minimum of two strong hydrophilic contacts between a pair of adjacent helices in the dimer to tetramer systems and only one strong hydrophilic interhelix contact in helix pairs of the pentamer and hexamer systems. We believe these results are general and can be applied to more complex ion channels, providing insight into ion channel stability and assembly.  相似文献   

18.
Using molecular dynamics simulations, we studied the structure, interhelix interactions, and dynamics of transmembrane proteins. Specifically, we investigated homooligomeric helical bundle systems consisting of synthetic α-helices with either the sequence Ac-(LSLLLSL)3-NH2 (LS2) or Ac-(LSSLLSL)3-NH2 (LS3). The LS2 and LS3 helical peptides are designed to have amphipathic characteristics that form ion channels in membrane. We simulated bundles containing one to six peptides that were embedded in palmitoyl-oleoyl-phosphatidylcholine (POPC) lipid bilayer and placed between two lamellae of water. We aim to provide a fundamental understanding of how amphipathic helical peptides interact with each other and their dynamical behaviors in different homooligomeric states. To understand structural properties, we examined the helix lengths, tilt angles of individual helices and the entire bundle, interhelix distances, interhelix cross-angles, helix hydrophobic-to-hydrophilic vector projections, and the average number of interhelix hydrophilic (serine–serine) contacts lining the pore of the transmembrane channel. To analyze dynamical properties, we calculated the rotational autocorrelation function of each helix and the cross-correlation of the rotational velocity between adjacent helices. The observed structural and dynamical characteristics show that higher order bundles containing four to six peptides are composed of multiple lower order bundles of one to three peptides. For example, the LS2 channel was found to be stable in a tetrameric bundle composed of a “dimer of dimers.” In addition, we observed that there is a minimum of two strong hydrophilic contacts between a pair of adjacent helices in the dimer to tetramer systems and only one strong hydrophilic interhelix contact in helix pairs of the pentamer and hexamer systems. We believe these results are general and can be applied to more complex ion channels, providing insight into ion channel stability and assembly.  相似文献   

19.
Lipid-peptide interaction has been investigated using cationic amphiphilic alpha-helical peptides and systematically varying their hydrophobic-hydrophilic balance (HHB). The influence of the peptides on neutral and acidic liposomes was examined by 1) Trp fluorescence quenched by brominated phospholipid, 2) membrane-clearing ability, 3) size determination of liposomes by dynamic light scattering, 4) morphological observation by electron microscopy, and 5) ability to form planar lipid bilayers from channels. The peptides examined consist of hydrophobic Leu and hydrophilic Lys residues with ratios 13:5, 11:7, 9:9, 7:11, and 5:13 (abbreviated as Hels 13-5, 11-7, 9-9, 7-11, and 5-13, respectively; Kiyota, T., S. Lee, and G. Sugihara. 1996. Biochemistry. 35:13196-13204). The most hydrophobic peptide (Hel 13-5) induced a twisted ribbon-like fibril structure for egg PC liposomes. In a 3/1 (egg PC/egg PG) lipid mixture, Hel 13-5 addition caused fusion of the liposomes. Hel 13-5 formed ion channels in neutral lipid bilayer (egg PE/egg PC = 7/3) at low peptide concentrations, but not in an acidic bilayer (egg PE/brain PS = 7/3). The peptides with hydrophobicity less than Hel 13-5 (Hels 11-7 and Hel 9-9) were able to partially immerse their hydrophobic part of the amphiphilic helix in lipid bilayers and fragment liposome to small bicelles or micelles, and then the bicelles aggregated to form a larger assembly. Peptides Hel 11-7 and Hel 9-9 each formed strong ion channels. Peptides (Hel 7-11 and Hel 5-13) with a more hydrophilic HHB interacted with an acidic lipid bilayer by charge interaction, in which the former immerses the hydrophobic part in lipid bilayer, and the latter did not immerse, and formed large assemblies by aggregation of original liposomes. The present study clearly showed that hydrophobic-hydrophilic balance of a peptide is a crucial factor in understanding lipid-peptide interactions.  相似文献   

20.
We have prepared several alpha-melanotropin (alpha-MSH) analogues with tyrosine substituted for methionine at the 4-position and determined their melanotropic activities on the frog (Rana pipiens), lizard (Anolis carolinensis) and S-91 (Cloudman) mouse melanoma adenylate cyclase bioassays. The potencies of Ac-[Tyr4]-alpha-MSH4-10-NH2 and Ac-[Tyr4]-alpha-MSH4-11-NH2 were compared with alpha-MSH and with their corresponding methionine and norleucine substituted analogues. The Tyr-4 analogues were found to be less active than the Nle-4 analogues on both the frog and lizard assays. Ac-[Tyr4]-alpha-MSH4-10-NH2 was found to be less active than Ac-[Tyr4]-alpha-MSH4-11-NH2 on the lizard bioassay, but more active than the longer fragment on the frog skin assay. Ac-[Tyr4]-alpha-MSH4-10-NH2 exhibited extremely prolonged biological activity on frog skin, but not on lizard skin, while the melanotropic activity of Ac-[Tyr4]-alpha-MSH4-11-NH2 was rapidly reversed on both assay systems. The increased potency of Ac-[Tyr4]-alpha-MSH4-10-NH2 over Ac-[Tyr4]-alpha-MSH4-11-NH2 on frog melanocytes may be related to the fact that the shorter 4-10 analogue exhibits prolonged biological activity. Interestingly, it was found that both Tyr-4 analogues were partial agonists on the mouse melanoma adenylate cyclase bioassay, and stimulated the enzyme to only about 50% of the maximal activity of alpha-MSH. We reported previously that replacement of L-Phe-7 by its D-enantiomer in [Nle4]-alpha-MSH and its Nle-4 containing analogues resulted in peptides with increased potency and in some instances prolonged activity.(ABSTRACT TRUNCATED AT 250 WORDS)  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