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.
Alamethicin is a hydrophobic antibiotic peptide 20 amino acids in length. It is predominantly helical and partitions into lipid bilayers mostly in transmembrane orientations. The rate of the peptide transverse diffusion (flip-flop) in palmitoyl-oleyl-phosphatidylcholine vesicles has been measured recently and the results suggest that it involves an energy barrier, presumably due to the free energy of transfer of the peptide termini across the bilayer. We used continuum-solvent model calculations, the known x-ray crystal structure of alamethicin and a simplified representation of the lipid bilayer as a slab of low dielectric constant to calculate the flip-flop rate. We assumed that the lipids adjust rapidly to each configuration of alamethicin in the bilayer because their motions are significantly faster than the average peptide flip-flop time. Thus, we considered the process as a sequence of discrete peptide-membrane configurations, representing critical steps in the diffusion, and estimated the transmembrane flip-flop rate from the calculated free energy of the system in each configuration. Our calculations indicate that the simplest possible pathway, i.e., the rotation of the helix around the bilayer midplane, involving the simultaneous burial of the two termini in the membrane, is energetically unfavorable. The most plausible alternative is a two-step process, comprised of a rotation of alamethicin around its C-terminus residue from the initial transmembrane orientation to a surface orientation, followed by a rotation around the N-terminus residue from the surface to the final reversed transmembrane orientation. This process involves the burial of one terminus at a time and is much more likely than the rotation of the helix around the bilayer midplane. Our calculations give flip-flop rates of approximately 10(-7)/s for this pathway, in accord with the measured value of 1.7 x 10(-6)/s.  相似文献   

2.
The topologies of zervamicin II and alamethicin, labeled with (15)N uniformly, selectively, or specifically, have been investigated by oriented proton-decoupled (15)N solid-state NMR spectroscopy. Whereas at lipid-to-peptide (L/P) ratios of 50 (wt/wt) zervamicin II exhibits transmembrane alignments in 1,2-dicapryl (di-C10:0-PC) and 1,2-dilauroyl (di-C12:0-PC) phosphatidylcholine bilayers, it adopts orientations predominantly parallel to the membrane surface when the lengths of the fatty acyl chains are extended. The orientational order of zervamicin II increases with higher phospholipid concentrations, and considerable line narrowing is obtained in di-C10:0-PC/zervamicin II membranes at L/P ratios of 100 (wt/wt). In contrast to zervamicin, alamethicin is transmembrane throughout most, if not all, of its length when reconstituted into 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine bilayers. The (31)P solid-state NMR spectra of all phospholipid/peptaibol samples investigated show a high degree of headgroup order, indicating that the peptides do not distort the bilayer structure. The observed differences in peptide orientation between zervamicin and alamethicin are discussed with reference to differences in their lengths, helical conformations, distribution of (hydroxy)proline residues, and hydrophobic moments. Possible implications for peptaibol voltage-gating are also described.  相似文献   

3.
Analogues of alamethicin, a 20-mer amphipathic helical peptide with ionophore activity, in the sequence of which all Aib residues were substituted by Ala (A1) or Leu (L1), were synthesized by the solid phase method, purified by high performance liquid chromatography and characterized by fast atomic bombardment mass spectrometry. Infrared and CD studies showed that A1 easily underwent a transconformation to beta-structure whereas L1 displayed a predominant alpha-helical character, thus being a potential ionophore model. Its voltage-dependent multistate activity in model membranes showed that Aib is not a requisite residue to observe an alamethicin-like behavior. However, as the lifetime of the single channels was much shorter than for alamethicin, the peptide chain was lengthened by a Leu (LL1) or a Ser (SL1) residue. The last peptide gave an increased channel lifetime, but the design of other non-Aib peptides, taking into account the hydroxyl C-terminus and side-chain interactions between helices in a barrel-stave bundle, is desirable to approach more closely the alamethicin activity.  相似文献   

4.
Mechanism of alamethicin insertion into lipid bilayers.   总被引:8,自引:6,他引:2       下载免费PDF全文
K He  S J Ludtke  W T Heller    H W Huang 《Biophysical journal》1996,71(5):2669-2679
Alamethicin adsorbs on the membrane surface at low peptide concentrations. However, above a critical peptide-to-lipid ratio (P/L), a fraction of the peptide molecules insert in the membrane. This critical ratio is lipid dependent. For diphytanoyl phosphatidylcholine it is about 1/40. At even higher concentrations P/L > or = 1/15, all of the alamethicin inserts into the membrane and forms well-defined pores as detected by neutron in-plane scattering. A previous x-ray diffraction measurement showed that alamethicin adsorbed on the surface has the effect of thinning the bilayer in proportion to the peptide concentration. A theoretical study showed that the energy cost of membrane thinning can indeed lead to peptide insertion. This paper extends the previous studies to the high-concentration region P/L > 1/40. X-ray diffraction shows that the bilayer thickness increases with the peptide concentration for P/L > 1/23 as the insertion approaches 100%. The thickness change with the percentage of insertion is consistent with the assumption that the hydrocarbon region of the bilayer matches the hydrophobic region of the inserted peptide. The elastic energy of a lipid bilayer including both adsorption and insertion of peptide is discussed. The Gibbs free energy is calculated as a function of P/L and the percentage of insertion phi in a simplified one-dimensional model. The model exhibits an insertion phase transition in qualitative agreement with the data. We conclude that the membrane deformation energy is the major driving force for the alamethicin insertion transition.  相似文献   

5.
Lipid-alamethicin interactions influence alamethicin orientation   总被引:12,自引:9,他引:3       下载免费PDF全文
  相似文献   

6.
Alamethicin is a 20-amino-acid peptide that produces a voltage-dependent conductance in membranes. We investigated the state of aggregation of alamethicin in egg phosphatidylcholine and dioleoylphosphatidylcholine membranes by examining the EPR spectra obtained from an active analog of this peptide that is spin-labeled at its C-terminus. The dependence of both the linewidth and signal intensity as a function of peptide concentration exhibit exchange broadening as the peptide concentration is increased; however, the exchange rates are linear with peptide concentration as is expected for the simple diffusion of monomers. In addition, the spin-exchange rates obtained from the linebroadening are consistent with collisional rates that are predicted from free Brownian diffusion. The results provide strong evidence that in the absence of a membrane potential, alamethicin is largely monomeric in these membranes.  相似文献   

7.
A nitroxide spin label attached to the C-terminus of the channel forming peptide alamethicin produces an enhancement of the nuclear spin-lattice relaxation rates of peptide protons as a result of both intermolecular and intramolecular magnetic dipole-dipole interactions. The intermolecular contribution provides evidence that alamethicin monomers collide preferentially in a C-terminal-to-N-terminal configuration in methanol. From the intramolecular paramagnetic enhancement of nuclear spin-lattice relaxation times, effective distances between the unpaired electron on the nitroxide at the C-terminus of alamethicin and protons along the peptide backbone were calculated. These distances are much shorter than distances based on the reported crystal structure of alamethicin, and cannot be accounted for by motion in the bonds that attach the nitroxide to the peptide. In addition, the differences between distances deduced from the nuclear spin relaxation and the distances seen in the crystal structure increase toward the N-terminal end of the peptide. The simplest explanation for these data is that the alamethicin backbone suffers large structural fluctuations that yield shorter effective distances between the C-terminus and positions along the backbone. This finding can be interpreted in terms of a molecular mechanism for the voltage-gating of the alamethicin channel. When the distances between a paramagnetic center and a nucleus fluctuate, paramagnetic enhancements are expected to yield distances that are weighted by r-6, and distances calculated using the Solomon-Bloembergen equations may more nearly represent a distance of closest approach than a time average distance.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Alamethicin is a 19-residue hydrophobic peptide, which is extended by a C-terminal phenylalaninol but lacks residues that might anchor the ends of the peptide at the lipid-water interface. Voltage-dependent ion channels formed by alamethicin depend strongly in their characteristics on chain length of the host lipid membranes. EPR spectroscopy is used to investigate the dependence on lipid chain length of the incorporation of spin-labeled alamethicin in phosphatidylcholine bilayer membranes. The spin-label amino acid TOAC is substituted at residue positions n = 1, 8, or 16 in the sequence of alamethicin F50/5 [TOAC(n), Glu(OMe)(7,18,19)]. Polarity-dependent isotropic hyperfine couplings of the three TOAC derivatives indicate that alamethicin assumes approximately the same location, relative to the membrane midplane, in fluid diC(N)PtdCho bilayers with chain lengths ranging from N = 10-18. Residue TOAC(8) is situated closest to the bilayer midplane, whereas TOAC(16) is located farther from the midplane in the hydrophobic core of the opposing lipid leaflet, and TOAC(1) remains in the lipid polar headgroup region. Orientational order parameters indicate that the tilt of alamethicin relative to the membrane normal is relatively small, even at high temperatures in the fluid phase, and increases rather slowly with decreasing chain length (from 13 degrees to 23 degrees for N = 18 and 10, respectively, at 75 degrees C). This is insufficient for alamethicin to achieve hydrophobic matching. Alamethicin differs in its mode of incorporation from other helical peptides for which transmembrane orientation has been determined as a function of lipid chain length.  相似文献   

9.
One area of the search for hormonal signaling systems regulators is development of peptides that correspond to the cytoplasmic regions of G protein-coupled receptors (GPCR). Modification of such peptides with hydrophobic radicals increases their efficiency and selectivity. However, at present it has not been studied how the activity of the peptide depends on the localization of hydrophobic radicals, their number, and chemical nature. The aim of this work consisted in synthesis of peptide 562–572 derivatives modified by fatty-acid radicals and corresponding to the C-terminal region of the luteinizing hormone receptor (LHR) and in the study of regulatory effects of the acylated LHR peptides on the basal and hormone-stimulated activity of adenylyl cyclase (AC) in rat tissues. To elucidate the effects of localization of hydrophobic radicals and of their number, modifications of peptide 562–572 were carried out only at the N-or at the C-terminus or at both ends. To study the effect of hydrophobicity, residues of palmitic (Pal) and decanoic (Dec) acids were chosen. Using a solid-phase strategy synthesis was performed of the unmodified peptide NKDTKIAKK-Nle-A562-572-KA (1) and five of its acylated analogues, N[K(Dec)]DTKIAKK-Nle-A562-572-KA (2), NKDTKIAKK-Nle-A562-572-[K(Dec)]A (3), N[K(Dec)]DTKIAKK-Nle-A562-572-[K(Dec)]A (4), N[K(Pal)]DTKIAKK-Nle-A562-572-KA (5), and NKDTKIAKK-Nle-A562-572-[K(Pal)]A (6). Peptide 6 modified with palmitate at the C-terminus to a large extent increased the basal AC activity and reduced the AC stimulating effect of human chorionic gonadotropin (hCG) in testes of rats; peptides 3 and 4 modified with decanoate at the C-terminus were less effective, but exceeded in activity the unmodified peptide 1; and peptides 2 and 5 acylated at the N-terminus were little active. The action of peptides was characterized by tissue and the receptor specificity. Thus, modification of the LHR peptide 562–572 with fatty-acid radicals at the C-terminus enhances its regulatory effect on the functional activity of the adenylyl cyclase system in rat testes, which indicates a promising modification of GPCR peptides with hydrophobic radicals. These data confirm the hypothesis that the hydrophobic radical is to be localized in the locus of GPCR peptide, where a transmembrane domain is located in the receptor.  相似文献   

10.
Alamethicin is a 19-amino-acid residue hydrophobic peptide of the peptaibol family that has been the object of numerous studies for its ability to produce voltage-dependent ion channels in membranes. In this work, for the first time electron paramagnetic resonance spectroscopy was applied to study the interaction of alamethicin with oriented bicelles. We highlighted the effects of increasing peptide concentrations on both the peptide and the membrane in identical conditions, by adopting a twofold spin labeling approach, placing a nitroxide moiety either on the peptide or on the phospholipids. The employment of bicelles affords additional spectral resolution, thanks to the formation of a macroscopically oriented phase that allows to gain information on alamethicin orientation and dynamics. Moreover, the high viscosity of the bicellar solution permits the investigation of the peptide aggregation properties at physiological temperature. We observed that, at 35 °C, alamethicin adopts a transmembrane orientation with the peptide axis forming an average angle of 25° with respect to the bilayer normal. Moreover, alamethicin maintains its dynamics and helical tilt constant at all concentrations studied. On the other hand, by increasing the peptide concentration, the bilayer experiences an exponential decrease of the order parameter, but does not undergo micellization, even at the highest peptide to lipid ratio studied (1:20). Finally, the aggregation of the peptide at physiological temperature shows that the peptide is monomeric at peptide to lipid ratios lower than 1:50, then it aggregates with a rather broad distribution in the number of peptides (from 6 to 8) per oligomer.  相似文献   

11.
Infection by enveloped viruses initially involves membrane fusion between viral and host cell membranes. The fusion peptide plays a crucial role in triggering this reaction. To clarify how the fusion peptide exerts this specific function, we carried out biophysical studies of three fusion peptide analogs of influenza virus hemagglutinin HA2, namely E5, G13L, and L17A. E5 exhibits an activity similar to the native fusion peptide, whereas G13L and L17A, which are two point mutants of the E5 analog, possess much less fusion activity. Our CD data showed that the conformations of these three analogs in SDS micelles are pH-dependent, with higher alpha-helical contents at acidic pH. Tryptophan fluorescence emission experiments indicated that these three analogs insert deeper into lipid bilayers at acidic pH. The three-dimensional structure of the E5 analog in SDS micelles at pH 4.0 revealed that two segments, Leu(2)-Glu(11) and Trp(14)-Ile(18), form amphipathic helical conformations, with Gly(12)-Gly(13) forming a hinge. The hydrophobic residues in the N- and C-terminal helices form a hydrophobic cluster. At neutral pH, however, the C-terminal helix of Trp(14)-Ile(18) reduces dramatically, and the hydrophobic core observed at acidic pH is severely disrupted. We suggest that the disruption of the C-terminal helix renders the E5 analog fusion-inactive at neutral pH. Furthermore, the decrease of the hinge and the reduction of fusion activity in G13L reveal the importance of the hinge in fusion activity. Also, the decrease in the C-terminal helix and the reduction of fusion activity in L17A demonstrates the importance of the C-terminal helix in fusion activity. Based on these biophysical studies, we propose a model that illustrates the structural change of the HA2 fusion peptide analog and explains how the analog interacts with the lipid bilayer at different pH values.  相似文献   

12.
The minimal structure in the heat-stable inhibitor protein of cAMP-dependent protein kinase required for a low nanomolar potency of inhibition is the peptide Thr6-Tyr-Ala-Asp-Phe-Ile-Ala-Ser-Gly-Arg-Thr-Gly-Arg-Arg-Asn-Ala-+ ++Ile22-NH2 (PKI-(6-22)-amide). While primary structural determinants for interaction with the protein kinase are distributed throughout the 17 residues of this peptide, we have previously shown that phenylalanine 10 in the NH2-terminal portion is a particularly important determinant for high affinity binding (Glass, D. B., Cheng, H.-C., Mende-Mueller, L., Reed, J., and Walsh, D. A. (1989) J. Biol. Chem. 264, 8802-8810). To investigate this requirement further, peptide analogs of PKI-(6-22)-amide in which various natural and nonstandard amino acids are substituted for phenylalanine 10 have been synthesized and tested for inhibitory potency against the catalytic subunit of the protein kinase. Consistent with the importance of the hydrophobicity of phenylalanine, an alanine 10 substitution analog exhibited a 270-fold decrease in inhibitory potency, whereas the leucine 10 analog lost only 33-fold in activity as compared to the parent peptide PKI-(6-22)-amide. Peptides containing the spatial conformation analogs D-phenylalanine, homophenylalanine, or phenylglycine were 60-120-fold less potent than the parent peptide. Peptides containing various para-substituted phenylalanines at position 10 were only 5-11-fold less potent. One exception to this was (4'-azidophenylalanine 10)PKI-(6-22)-amide, which was nearly equipotent with the parent inhibitor. The most potent analogs were those peptides containing highly aromatic residues at position 10. The 2'-thienylalanine 10, tryptophan (formyl) 10, tryptophan 10, and the 1'-naphthylalanine 10 analogs were 3-fold less potent, equipotent, slightly more potent, and 4-fold more potent than the parent peptide inhibitor, respectively. We conclude that phenylalanine 10 in PKI-(6-22)-amide, and presumably in the native protein inhibitor, interacts through specific hydrophobic and/or aromatic binding to a hydrophobic pocket or cleft near the active site of the protein kinase.  相似文献   

13.
The effects of the hydrophobicity and the distribution of hydrophobic residues on the surfaces of some designed alpha-helical transmembrane peptides (acetyl-K2-L(m)-A(n)-K2-amide, where m + n = 24) on their solution behavior and interactions with phospholipids were examined. We find that although these peptides exhibit strong alpha-helix forming propensities in water, membrane-mimetic media, and lipid model membranes, the stability of the helices decreases as the Leu content decreases. Also, their binding to reversed phase high-performance liquid chromatography columns is largely determined by their hydrophobicity and generally decreases with decreases in the Leu/Ala ratio. However, the retention of these peptides by such columns is also affected by the distribution of hydrophobic residues on their helical surfaces, being further enhanced when peptide helical hydrophobic moments are increased by clustering hydrophobic residues on one side of the helix. This clustering of hydrophobic residues also increases peptide propensity for self-aggregation in aqueous media and enhances partitioning of the peptide into lipid bilayer membranes. We also find that the peptides LA3LA2 [acetyl-K2-(LAAALAA)3LAA-K2-amide] and particularly LA6 [acetyl-K2-(LAAAAAA)3LAA-K2-amide] associate less strongly with and perturb the thermotropic phase behavior of phosphatidylcholine bilayers much less than peptides with higher L/A ratios. These results are consistent with free energies calculated for the partitioning of these peptides between water and phospholipid bilayers, which suggest that LA3LA2 has an equal tendency to partition into water and into the hydrophobic core of phospholipid model membranes, whereas LA6 should strongly prefer the aqueous phase. We conclude that for alpha-helical peptides of this type, Leu/Ala ratios of greater than 7/17 are required for stable transmembrane associations with phospholipid bilayers.  相似文献   

14.
Ampullosporin A and alamethicin are two members of the peptaibol family of antimicrobial peptides. These compounds are produced by fungi and are characterized by a high content of hydrophobic amino acids, and in particular the α-tetrasubstituted amino acid residue α-aminoisobutyric acid. Here ampullosporin A and alamethicin were uniformly labeled with 15N, purified and reconstituted into oriented phophatidylcholine lipid bilayers and investigated by proton-decoupled 15N and 31P solid-state NMR spectroscopy. Whereas alamethicin (20 amino acid residues) adopts transmembrane alignments in 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) or 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) membranes the much shorter ampullosporin A (15 residues) exhibits comparable configurations only in thin membranes. In contrast the latter compound is oriented parallel to the membrane surface in 1,2-dimyristoleoyl-sn-glycero-3-phosphocholine and POPC bilayers indicating that hydrophobic mismatch has a decisive effect on the membrane topology of these peptides. Two-dimensional 15N chemical shift - 1H-15N dipolar coupling solid-state NMR correlation spectroscopy suggests that in their transmembrane configuration both peptides adopt mixed α-/310-helical structures which can be explained by the restraints imposed by the membranes and the bulky α-aminoisobutyric acid residues. The 15N solid-state NMR spectra also provide detailed information on the helical tilt angles. The results are discussed with regard to the antimicrobial activities of the peptides.  相似文献   

15.
Two spin-labeled derivatives of the ion conductive peptide alamethicin were synthesized and used to examine its binding and state of aggregation. One derivative was spin labeled at the C-terminus and the other, a leucine analogue, was spin labeled at the N-terminus. In methanol, both the C and N terminal labeled peptides were monomeric. In aqueous solution, the C-terminal derivative was monomeric at low concentrations, but aggregated at higher concentrations with a critical concentration of 23 microM. In the membrane, the C-terminal label was localized to the membrane-aqueous interface using 13C-NMR, and could assume more than one orientation. The membrane binding of the C-terminal derivative was examined using EPR, and it exhibited a cooperativity seen previously for native alamethicin. However, this cooperativity was not the result of an aggregation of the peptide in the membrane. When the spectra of either the C or N-terminal labeled peptide were examined over a wide range of membrane lipid to peptide ratios, no evidence for aggregation could be found and the peptides remained monomeric under all conditions examined. Because electrical measurements on this peptide provide strong evidence for an ion-conductive aggregate, the ion-conductive form of alamethicin likely represents a minor fraction of the total membrane bound peptide.  相似文献   

16.
Alamethicin is a 20-residue, hydrophobic, helical peptide, which forms voltage-sensitive ion channels in lipid membranes. The helicogenic, nitroxyl amino acid TOAC was substituted isosterically for Aib at residue positions 1, 8, or 16 in a F50/5 alamethicin analog to enable EPR studies. Electron spin-echo envelope modulation (ESEEM) spectroscopy was used to investigate the water exposure of TOAC-alamethicin introduced into membranes of saturated or unsaturated diacyl phosphatidylcholines that were dispersed in D2O. Echo-detected EPR spectra were used to assess the degree of assembly of the peptide in the membrane, via the instantaneous diffusion from intermolecular spin-spin interactions. The profile of residue exposure to water differs between membranes of saturated and unsaturated lipids. In monounsaturated dioleoyl phosphatidylcholine, D2O-ESEEM intensities decrease from TOAC1 to TOAC8 and TOAC16 but not uniformly. This is consistent with a transmembrane orientation for the protoassembled state, in which TOAC16 is located in the bilayer leaflet opposite to that of TOAC1 and TOAC8. Relative to the monomer in fluid bilayers, assembled alamethicin is disposed asymmetrically about the bilayer midplane. In saturated dimyristoyl phosphatidylcholine, the D2O-ESEEM intensity is greatest for TOAC8, indicating a more superficial location for alamethicin, which correlates with the difference in orientation between gel- and fluid-phase membranes found by conventional EPR of TOAC-alamethicin in aligned phosphatidylcholine bilayers. Increasing alamethicin/lipid ratio in saturated phosphatidylcholine shifts the profile of water exposure toward that with unsaturated lipid, consistent with proposals of a critical concentration for switching between the two different membrane-associated states.  相似文献   

17.
Both structural and functional studies suggest that pH gating of the inward rectifier potassium (K) channel, Kir1.1 (ROMK), is mediated by the convergence of 4 hydrophobic leucines (one from each subunit) near the cytoplasmic bundle-crossing of the inner transmembrane helices. We tested this hypothesis by moving the putative leucine gate from the L160-Kir1.1b to other positions along the inner transmembrane helix, and measuring inward current and conductance as functions of internal pH, using the Xenopus oocyte heterologous expression system. Results of these studies indicated that it was possible to replace the putative inward rectifier pH gate at L160-Kir1.1b by either a leucine or methionine at 157-Kir1.1b (G157L-L160G or G157M-L160G). Although both leucine and methionine gated the channel at 157-Kir1.1b, residues of similar hydrophobicity (tyrosine and valine) did not. Hence, hydrophobicity was a necessary but not a sufficient condition for steric gating at 157. This was in contrast to the 160-Kir1.1b locus, where side-chain hydrophobicity was both a necessary and sufficient property for steric gating. Homology models were constructed for all mutants that expressed significant whole-cell currents, using the closed-state coordinates of the prokaryotic inward rectifier, KirBac1.1. Models of mutants that retained pH gating were too narrow at the bundle crossing to permit hydrated K ion permeation in the closed-state. On the other hand, mutants that lost pH gating had ample space at the bundle crossing for hydrated K permeation in the closed-state. These results support our hypothesis that hydrophobic leucines at the cytoplasmic end of the inner transmembrane helices comprise the principal pH gate of Kir1.1, a gate that can be relocated from 160-Kir1.1b to 157-Kir1.1b.  相似文献   

18.
Truncations at the carboxyl termini of G protein-coupled receptors result in defective receptor biogenesis and comprise a number of inherited disorders. In order to evaluate the structural role of the C-terminus in G protein-coupled receptor biogenesis, we generated a series of deletion and substitution mutations in the dopamine D1 receptor and visualized receptor subcellular localization by fusion to a green fluorescent protein. Alanine substitutions of several hydrophobic residues within the proximal C-terminus resulted in receptor transport arrest in the ER. Agonist binding and coupling to adenylyl cyclase was also abolished. In contrast, substitutions conserving C-terminal hydrophobicity produced normal cell surface receptor expression, binding, and stimulatory function. A mechanism for the role of the C-terminus in D1 receptor transport was investigated by searching for candidate protein interactions. The D1 receptor was found to co-precipitate and associate in vitro directly with the gamma-subunit of the COPI coatomer complex. In vitro pull-down assays confirmed that only the D1 C-terminus is required for COPI association, and that identical mutations causing disruption of receptor transport to the cell surface also disrupted binding to COPI. Furthermore, conservative mutations in the D1 C-terminus restored COPI association just as they restored cell surface transport. These results suggest that association between the coatomer complex and hydrophobic residues within the proximal C-terminus of the D1 receptor may serve an important role in receptor transport.  相似文献   

19.
Pleurocidin (Ple) is a peptide derived from the winter flounder. In our previous study, we reported the antifungal effect of Ple and its mode of action. To develop novel antifungal peptides useful as therapeutic agents, two analogs, with amino acid substitutions, were designed to decrease the net hydrophobicity by Arg (R) or Ser (S)-substitution at the hydrophobic face of Ple without changing the amphipathic structure. By substituting Ser, the hydrophobicity of the peptide (anal-S) was decreased, and by substituting Arg, though the hydrophobicity of the peptide (anal-R) was decreased, the cationicity of this peptide was increased. CD measurements showed the substitution of Arg or Ser decrease the α-helical conformation of analog peptides. Studies with analog peptides have shown decreases in hydrophobicity and α-helicity do not affect antifungal activity but decrease hemolytic activity. These results suggest that highly hydrophobic and α-helical natures are not desirable in the design of antimicrobial peptides.  相似文献   

20.
In order to test the influence of chemical modifications designed to allow covalent coupling of channel-forming peptide motifs into variable sized oligomers, a series of alamethicin derivatives was prepared. The building block encompassing the N-terminal 1-17 residues of alamethicin behaved normally in the conductance assay on planar lipid bilayers, albeit at higher concentration and with a slightly reduced voltage-dependence. A linker Ac-K-OCH(2)C(6)H(4)CH(3)p attached via the epsilon amino group of lysine to the C-terminus of alamethicin(1-17) increased membrane affinity. The latter was further enhanced in a dimer and a tetramer in which alamethicin(1-17) chains were tethered to di- or tetra-lysine linkers, respectively, but macroscopic current-voltage curves displayed much reduced voltage-dependencies and reversed hysteresis. An usual behaviour with high voltage-dependence was restored with the modified dimer of alamethicin(1-17) in which alanine separated the two consecutive lysine residues in the linker. Of special interest was the development of a 'negative resistance' branch in macroscopic current-voltage curves for low concentrations of this dimer with the more flexible linker. Single channel events displayed only one single open state with fast kinetics and whose conductance matches that of the alamethicin heptamer or octamer.  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