Báo cáo khoa học: Interactions of the antimicrobial b-peptide b-17 with phospholipid vesicles differ from membrane interactions of magainins potx

9 266 0
Báo cáo khoa học: Interactions of the antimicrobial b-peptide b-17 with phospholipid vesicles differ from membrane interactions of magainins potx

Đang tải... (xem toàn văn)

Thông tin tài liệu

Interactions of the antimicrobial b-peptide b-17 with phospholipid vesicles differ from membrane interactions of magainins Raquel F. Epand 1 , Naoki Umezawa 2 , Emilie A. Porter 2 , Samuel H. Gellman 2 and Richard M. Epand 1 1 Department of Biochemistry, McMaster University Health Sciences Centre, Hamilton, Canada; 2 Department of Chemistry, University of Wisconsin, Wisconsin USA We have studied the interaction of b-17, a potent synthetic antimicrobial b-peptide, with phospholipids. We find that unlike other antimicrobial peptides such as magainin II, b-17 facilitates the formation of nonbilayer phases, indicating that the peptide promotes negative curvature. Studies of liposomal leakage also indicate a different mode of mem- brane interaction relative to magainin II, but both leakage and membrane binding show that b-17, like magainin II, has strong affinity for membranes containing anionic lipids. This is likely to be an important factor contributing to the anti- microbial specificity of the b-peptide. Keywords: b-peptide; antimicrobial; peptide–lipid inter- actions; membrane leakage; membrane intrinsic curvature. There is currently extensive interest in developing new antimicrobial agents to counter the growing clinical pro- blem of bacterial resistance to traditional antibiotics [1]. A large variety of natural peptides and synthetic derivatives display antimicrobial activity [2], and these peptides have been viewed as potential sources of new therapeutic agents. Nearly all of the synthetic derivatives have been constructed from a-amino acids, as are the natural host-defense peptides. Recently, antimicrobial activity has been reported for a number of b-amino acid oligomers (b-peptides) [3–8]. b-Peptides differ from conventional peptides (a-amino acid residues) in two important ways. First, the unnatural backbone of b-peptides confers resistance to degradation by proteolytic enzymes [8–10]. Second, b-peptides construc- ted from appropriately rigidified residues display higher conformational stability, on a per-residue basis, than do conventional peptides [6]. It is therefore of interest to compare the mechanism(s) of b-peptide antibacterial action with the antibacterial mechanisms of analogous conven- tional peptides. Here we characterize the interactions of one antimicrobial b-peptide, b-17 [7,8], with lipid vesicles, which are simplified models of cell membranes. b-17 was designed to mimic natural host-defense peptides like the magainins, which are thought to exert their antimicrobial effects by disrupting bacterial membranes. The structure of b-17 is shown in Fig. 1. Magainins are cationic peptides that contain 23 a-amino acid residues and adopt an amphiphilic a-helical conformation in the presence of membranes [11]. b-17 contains 17 b-amino acid residues and adopts a 12-helical conformation, which is defined by a network of 12-mem- bered ring C¼O(i) fi NH(i + 3) hydrogen bonds [12–15]. (The a-helix contains 13-membered ring C¼O(i) fi NH(i + 4) hydrogen bonds.) The b-peptide 12-helix has approximately 2.5 residues per turn and a rise of 5.5 A ˚ per turn; therefore, a 17-residue 12-helix should be similar in length to a 23-residue a-helix formed by a conventional peptide. b-17 contains only two types of b-amino acid residue, hydrophobic trans-2-aminocyclopentanecarboxylic acid (ACPC) and cationic trans-3-aminopyrrolidine-4- carboxylic acid (APC). The repeating APC-ACPC-APC- ACPC-ACPC pentad gives rise to a 12-helix with distinct cationic and hydrophobic sides, which mimics the amphi- philic a-helical conformation adopted by magainin peptides [11]. We have characterized interactions between b-17 and vesicles by differential scanning calorimetry, to gain insight on lipid phase transitions, and by circular dichroism (CD), to gain insight on b-peptide conformation in the presence and absence of membranes. We have also examined the ability of b-17 to lyse model membranes as a function of lipid composition, in an effort to explain the specificity of action and characterize the energetics of binding to the different liposomal systems used. Like magainins and other host-defense peptides, b-17 is more effective at killing bacterial cells than at inducing lysis of human red blood cells (which are taken to represent eukaryotic cells in general) [7,8]. The basis for this cell specificity among natural peptides remains a subject of debate [1,16–19], although electrostatic factors are thought to be important, as bacterial cell surfaces generally have a greater negative charge density than do eukaryotic cell surfaces [11]. Vesicle model studies have been conducted with many conventional antimicrobial peptides [18]; therefore, our results with b-17 have an ample Correspondence to R. M. Epand, Department of Biochemistry, 1200 Main Street West, McMaster University Health Sciences Centre, Hamilton, ON, L8N 3Z5 Canada. Fax: 905 521 1397, Tel.: 905 525 9140, E-mail: epand@mcmaster.ca Abbreviations: DPam 2 PtdEtn, dipalmitoleoyl-phosphatidylethanol- amine; Ole 2 PtdEtn, dioleoylphosphatidylethanolamine; Ole 2 PtdGro, dioleoylphosphatidylglycerol; Ole 2 PtdSer, dioleoylphosphatidyl- serine; eggPtdCho, phosphatidylcholine extracted from egg yolk; LUV, large unilamellar vesicle; PamOlePtdSer, 1-palmitoyl-2- oleoylphosphatidylserine; SUVs, small or sonicated unilamellar vesicles; T H , bilayer to hexagonal transition temperature. (Received 11 November 2002, revised 11 January 2003, accepted 27 January 2003) Eur. J. Biochem. 270, 1240–1248 (2003) Ó FEBS 2003 doi:10.1046/j.1432-1033.2003.03484.x basis for comparison. The lipids used in this study as components of lipid mixtures have the structures shown in Fig. 2. Materials and methods Materials Phospholipids were purchased from Avanti Polar Lipids (Alabaster, AL). The labeled lipids were made by chemical modification of phosphatidylethanolamine that was prepared from egg phosphatidylcholine by transphosphatidylation. [Ala8,13,18]-magainin II amide was purchased from the Sigma Chemical Co (St Louis, MO, USA). Synthesis of NBD-b-17 Fmoc-ACPC-OH and Fmoc-APC(Boc)-OH were synthe- sized as previously reported [20,21]. b-17 was prepared with a conventional automated solid-phase peptide synthesizer (Synergy 432 A; Applied Biosystems, Foster City, CA) using extended coupling and deprotection steps. After removal of the N-terminal Fmoc protecting group, the resin-bound peptide was treated with three equivalents of 4-fluoro-7-nitrobenzofurazan(4-fluoro-7-nitrobenz-2-oxa-1,3- diazole) (Molecular Probes, Eugene, OR) in dry dimethyl- formamide containing 5% v/v diisopropylethylamine. The reaction vessel was covered with foil and shaken at room temperature. After 48 h, the resin was washed with dimethylformamide and then CH 2 Cl 2 .NBD-labeledb-17 was cleaved from the resin and purified by RP-HPLC using a Vydac C4214TP510 column. Trifluoroacetic acid coun- terions were replaced with chloride ions through three dissolutions of the b-peptide in 50 m M HCl followed by lyophilization. The replacement of trifluoroacetic acid was confirmed by 19 F-NMR. A small amount of decomposition was seen during this step; the final purity of NBD-b-17 was  95%, and the purity of b-17 itself was >98%. Differential scanning calorimetry Lipid films were made from dipalmitoleoylphosphati- dylethanolamine (DPam 2 PtdEtn) dissolved in chloroform/ methanol (2 : 1, v/v). Increasing mole fractions of peptides dissolved in methanol were added and the solvent removed by evaporation with nitrogen. Final traces of organic solvent were removed in a vacuum chamber attached to a liquid nitrogen trap for 2–3 h. The lipid films were hydrated at room temperature by vortexing with 20 m M Pipes buffer containing 0.14 M NaCl, 1 m M EDTA and 0.002% sodium azide pH 7.4. The final lipid concentration was 5 mgÆmL )1 . Lipid suspensions were degassed under vacuum before being loaded into a NanoCal high sensitivity calorimeter (CSC, American Forks, UT). A heating scan rate of Fig. 1. The structure of b-17. Fig. 2. Structures of phospholipids used in these studies. (1) Dioleoylphosphatidyletha- nolamine (Ole 2 PtdEtn); (2) dielaidoylphos- phatidylethanolamine; (3) dipalmitoleoylphos- phatidylethanolamine (DPam 2 PtdEtn), and (4) dioleoylphosphatidylcholine (Ole 2 PtdCho) have zwitterionic headgroups. (5) Dioleoyl- phosphatidylglycerol (Ole 2 PtdGro); (6) diol- eoylphosphatidylserine (Ole 2 PtdSer) and (7) 1-palmitoyl-2-oleoylphosphatidylserine (PamOlePtdSer) have anionic headgroups. Lipids are sometimes referred to by their headgroups, e.g. phosphatidylethanolamine, phosphatidylcholine, phosphatidylglycerol, phosphatidylserine. Ó FEBS 2003 Antimicrobial b-peptide and phospholipid vesicles (Eur. J. Biochem. 270) 1241 0.75 KÆmin )1 was employed. The observed phase transition was fitted with parameters describing an equilibrium with a single van’t Hoff enthalpy and the transition temperature reported as that for the fitted curve. Data were analyzed with the program Origin 5.0. For some curves the endothermic peak was not symmetrical and it appeared that including more than one component would provide a better fit to the data. However, there was no progressive trend showing a systematic increase in one component relative to another as a function of b-peptide/lipid ratio. Fitting the transition to one component provides a measure of the average effect of the b-peptide on the curvature properties of the lipid. Circular dichroism (CD) The CD spectra were recorded using an AVIV model 61 DS CD instrument (AVIV Associates, Lakewood, NJ). The sample was contained in a 1-mm pathlength quartz cell that was maintained at 25 °C in a thermostated cell holder. The CD data are expressed as the mean residue ellipticity. All CD runs were made with b-peptide dissolved in 10 m M sodium phosphate buffer containing 0.14 M NaF and 1 m M EDTA at pH 7.4. For samples containing lipid, the lipid was first made into a dry film from a solution of chloroform/methanol, and then hydrated by vortexing with buffer. The lipid suspension was then sonicated to clarity to make small or sonicated unilamellar vesicles (SUVs). Large unilamellar vesicles Lipid films were made by dissolving appropriate amounts of lipid in a mixture of chloroform/methanol (2 : 1, v/v) and dried in a test tube under nitrogen. Final traces of solvent were removed in a vacuum chamber attached to a liquid nitrogen trap for 2–3 h. Dried films were kept under argon gas at )30 °C if not used immediately. Films were hydrated with buffer, vortexed extensively at room temperature and then subjected to five cycles of freezing and thawing. The homogeneous lipid suspensions were then further processed by 10 passes through two stacked 0.1 lm polycarbonate filters (Nucleopore Filtration Products, Pleasanton, CA) in a barrel extruder (Lipex Biomembranes, Vancouver, BC), at room temperature. Large unilamellar vesicles (LUVs) were kept on ice and used within a few hours of preparation. Lipid phosphorus was determined by the method of Ames [22]. Leakage studies Aqueous content leakage from liposomes was determined using the ANTS-DPX assay [23]. Lipid films were hydrated with 12.5 m M 8-aminonaphthalene-1,3,6-trisulfonic acid (ANTS), 45 m M p-xylene-bis-pyridinium bromide (DPX), 68 m M NaCl, 10 m M Hepes at pH 7.4. The osmolarity of this solution was adjusted to be equal to that of the buffer as measured with a cryoosmometer (Advanced Model 3MOplus Micro-Osmometer, Advanced Instruments Inc., Norwood, MA). LUVs of 0.1 lm diameter were prepared by extrusion as described above. After passage through a 2.5 · 20 cm column of Sephadex G-75, the void volume fractions were collected and the phospholipid concentration was determined by phosphate analysis. The fluorescence measurements were performed in 2 mL of buffer composed of 10 m M Hepes, 0.14 M NaCl, 1 m M EDTA, pH 7.4, in a quartz cuvette equilibrated at 37 °C with stirring. Aliquots of LUVs were added to the cuvette to a final lipid concentration of 25 l M . The fluorescence was recorded as a function of time using an excitation wavelength of 360 nm and an emission wavelength of 530 nm with 8 nm band- width slits. A 490-nm cutoff filter was placed in the emission path. Leakage was started by addition of the peptide in buffer solution. The value for 100% leakage was obtained adding 20 lL of a 10% Triton X-100 solution to the cuvette. Runs were performed in duplicate. Membrane binding NBD-labeled compounds have a low quantum yield in aqueous solution, but the quantum yield becomes higher when the probe enters a hydrophobic environment such as in a membrane [24]. To assess the degree of b-17 association with the membrane systems tested in leakage, a series of LUVs were made as previously described and aliquots of a suspension of LUVs were added sequentially to NBD- labeled a-peptide in the cuvette and the fluorescence intensity changes followed until a plateau was reached. Fluorescence intensity was measured with an Aminco- Bowman Series II spectrofluorimeter, with magnetic stir- ring, at 25 °C; excitation was set at 467 nm and emission at 539 nm, using 4 nm bandpass slits in excitation and emission and a 490-nm cut off filter in the emission path. Polarizers were set at 90° for the excitation and 0° for the emission to minimize scattering effects. Siliconized glass cuvettes were used, containing 2 mL of 10 m M Hepes buffer with 0.14 M NaCl and 1 m M EDTA at pH 7.4. b-Peptide solutions were made in the same buffer and added at a concentration of 0.4 l M to the cuvette, before addition of LUVs. After every addition of vesicles, the system was allowed to incubate for a few minutes. Controls with addition of LUVs to buffer only were also made and subtracted. The intensity changes were used to calculate binding affinity [25–27]. Isothermal titration calorimetry (ITC) The experiments to measure the heat of reaction at 30 °C were carried out on a VP-ITC isothermal titration calori- meter manufactured by MicroCal, Northampton, MA. The solutions were degassed under vacuum prior to being loaded in the calorimeter. The b-peptide solution (25–50 l M )was placed in the reaction cell and was titrated with 3–10 lLof 5–10 m M lipid mixtures in the form of LUVs delivered from a motor-driven syringe. The reference cell contained the same buffer as the solution placed in the cell. A buffer composed of 10 m M Hepes, 0.14 M NaCl, 1 m M EDTA at pH 7.4 was used in all titrations. The binding curves obtained were fitted with software provided by Microcal. Results Differential scanning calorimetry We measured how increasing mole fractions of b-17 changed the bilayer to hexagonal phase transition temperature (T H ) 1242 R. F. Epand et al.(Eur. J. Biochem. 270) Ó FEBS 2003 of a model unsaturated phospholipid, DPam 2 PtdEtn, to obtain an indication of the relationship between insertion of the peptide and the stability of the bilayer phase. DPam 2 PtdEtn is a zwitterionic lipid with a bilayer to hexagonal phase transition at 43 °C. Small mole fractions of additives that interact with this lipid have been shown to either lower this transition temperature, thereby destabi- lizing the lamellar phase and facilitating the conversion to a more highly curved nonlamellar phase, or vice versa.The change in transition temperature in the presence of small amounts of additives has been therefore used as a measure of propensity to induce curvature in the membrane [28]. Changes in membrane curvature are required for several of the proposed mechanisms by which antimicrobial peptides induce membrane leakage. Thus, the formation of a large pore or a carpet-like disruption on the surface of the bilayer requires increased positive membrane curvature. However, not all lytic peptides promote positive membrane curvature. For example, the peptide mastoparan, isolated from wasp venom, has the opposite effect on membrane curvature. It promotes negative curvature. How this effect of promoting negative curvature is related to the lytic action of masto- paran is not well understood, except for the fact that it would tend to destabilize the bilayer phase toward the formation of inverted phase morphology. This is not meant to imply that a peptide that promotes negative curvature will cause the formation of nonlamellar structures, but rather that it will facilitate the formation of transient structures with increased negative curvature. b-17 also promotes negative curvature and lowers the transition temperature of DPam 2 PtdEtn by )708 ± 104 °CÆmol )1 fraction of peptide added. This is opposite to the direction of the shift of T H of +1800 ± 300 °CÆmol )1 found with the addition of magainin II [29] and does not correspond to the type of curvature required for a large toroidal pore. We have also compared our previous differential scanning calori- metry results using magainin II with those of the more potent magainin analog, [Ala8,13,18]-magainin II amide [30]. We find that the peptide-induced changes in the transition of DPam 2 PtdEtn are similar for both magainin II and for [Ala8,13,18]-magainin II amide (data not shown). In a toroidal pore there are two types of curvatures to be considered, i.e. one in the plane of the membrane and the other along the bilayer normal; for large pores, the positive curvature required along the bilayer normal predominates. If b-17 promoted the formation of a similar pore arrange- ment, but the diameter of the pore were smaller than that of magainin, then the negative curvature in the dimension around the rim of the pore would predominate. However, the pore formed by b-17 does not appear to be very small as it allows the release of b-galactosidase from bacterial cells at a similar rate to magainin [8]. It is thus possible that b-17 lyses membranes by a pore-like mechanism, but the properties of the pore would have to be different from that formed with magainin in order to be consistent with the different curvature tendencies of these peptides. CD Solutions of b-17 in water were diluted into phosphate buffer, and the far UV CD spectra were measured in the presence and absence of SUVs (Fig. 3). When visually transparent SUVs of lipid mixtures are added to a solution of the b-peptide, there is a red shift in the CD spectra with an increase in signal intensity. Similar changes in 12-helical b-peptides are observed in going from water to methanol or trifluoroethanol [31,32]. Trifluoroethanol and methanol are known to stabilize helical conformations of both b-peptides [33] and b-peptides [3,32]. Thus the CD spectra in Fig. 3 suggest that the 12-helical conformation of b-17 is modestly stabilized in the presence of vesicles relative to aqueous solution. DeGrado et al. [3] have observed strong helix stabilization by vesicles for a different class of b-peptides that are more conformationally flexible than is b-17. Leakage Leakage experiments were designed to obtain information regarding the role of anionic headgroup, fatty acyl chain unsaturation and vesicle curvature in the action of b-17 on model membranes. For this purpose, different anionic lipids dioleoylphosphatidylglycerol (Ole 2 PtdGro) Fig. 3. Circular dichroism spectra of the b-17 peptide at a concentration of 100 l M . b-17 alone in buffer (solid line); b-17 with SUVs of Ole 2 PtdEtn : Ole 2 PtdSer (1 : 1) at a lipid to peptide molar ratio (L : P) ¼ 10 (dotted line); b-17 with SUVs of Ole 2 PtdEtn : Ole 2 PtdGro (1:1)ataL:P¼ 10 (dashed line). Ó FEBS 2003 Antimicrobial b-peptide and phospholipid vesicles (Eur. J. Biochem. 270) 1243 or dioleoylphosphatidylserine (Ole 2 PtdSer) were mixed with dioleoylphosphatidylethanolamine (Ole 2 PtdEtn), dielaidoyl- phosphatidylethanolamine or phosphatidylcholine extrac- ted from egg yolk (eggPtdCho), as the zwitterionic lipid. Phosphatidylethanolamine has a smaller headgroup and therefore a larger tendency to promote negative curvature compared to phosphatidylcholine. We included Ole 2 PtdEtn and dielaidoylphosphatidylethanolamine, two species of phosphatidylethanolamine, as components of these mix- tures, because the trans double bond at position 9–10 in dielaidoylphosphatidylethanolamine results in less deviation of the acyl chain from the bilayer normal than does the cis double bond of Ole 2 PtdEtn. Thus dielaidoylphosphatidyl- ethanolamine has less intrinsic negative curvature than Ole 2 PtdEtn, which is illustrated by the fact that pure dielaidoylphosphatidylethanolamine has a T H of 65 °C compared with 10 °C for pure Ole 2 PtdEtn, despite their similarity in chemical structure. The presence of the anionic lipids in the membrane allows for the formation of stable liposomes at room temperature in the presence of Ole 2 PtdEtn; pure PtdEtn does not easily form liposomes, presumably due to limited hydration of the headgroup and the higher energy cost of bending the membrane. In addition, the presence of the anionic lipid PtdGro in these mixtures is important as this is a major component of prokaryotic membranes, while the use of phosphatidylserine mimics the anionic lipid that becomes exposed to the extracellular side of the bilayer in cancer cells or in apoptotic mammalian cells. The zwitterionic lipid was mixed with the negatively charged lipid Ole 2 PtdSer or Ole 2 PtdGro at a 1 : 2 molar ratio. The rate and final extents of leakage is shown in Fig. 4A,D. The leakage of aqueous contents from lipo- somes of different lipid compositions is summarized by showing the dependence of the percentage leakage at 300 s on the b-peptide to lipid ratio (Fig. 5). In Fig. 6, leakage Fig. 4. Leakage from 25 l M LUVs of Ole 2 PtdEtn/Ole 2 PtdGro (1 : 2) (A); dielaidoylphosphatidylethanolamine/Ole 2 PtdGro (1 : 2) (B); eggPtdCho/ Ole 2 PtdGro (1 : 2) (C) or eggPtdCho (D) induced by b-17. Numbers identifying each curve correspond to the micromolar concentration of b-17. 1244 R. F. Epand et al.(Eur. J. Biochem. 270) Ó FEBS 2003 from Ole 2 PtdEtn LUVs containing the anionic lipids Ole 2 PtdGro, Ole 2 PtdSer, 1-palmitoyl-2-oleoylphosphatidyl- serine (PamOlePtdSer) or phosphatidylserine extracted from bovine brain is shown. Substituting either bovine brain phosphatidylserine; dielaidoylphosphatidylethanol- amine, or PamOlePtdSer for Ole 2 PtdSer had little effect, highlighting again the predominance of headgroup charge over chain length or degree of unsaturation of the fatty acid in controlling a-peptide induced leakage. EggPtdCho alone had a lower rate of leakage than did the lipid mixtures, implying the requirement for an anionic lipid. This result is expected for a peptide with antimicrobial activity, as microbial membranes have a net negative surface charge. Membrane binding All vesicle systems tested that contained Ole 2 PtdGro were found to bind strongly to NBD-b-17 peptide, as would be expected of a cationic molecule (Fig. 7). Binding to LUVs of Ole 2 PtdEtn/Ole 2 PtdGro and dielaidoylphosphatidyletha- nolamine/Ole 2 PtdGro was virtually indistinguishable, while the lowest binding affinity was found for Ole 2 PtdGro/ Ole 2 PtdSer. Isothermal titration calorimetry (ITC) Isothermal titration calorimetry was performed with b-17 in the cell and incremental addition of vesicles. The results of titration of LUVs of dielaidoylphosphatidylethanolamine/ Ole 2 PtdGro (1 : 2) into a solution of b-17 is given in Fig. 8A. Similar titrations were performed for mixtures of Ole 2 PtdEtn/Ole 2 PtdGro(2:1)aswellasOle 2 PtdEtn/ Ole 2 PtdSer (2 : 1) (not shown). The titration for pure eggPtdCho LUVs did not evolve significant heat with either b-17 or NBD-b-17 (Fig. 8B). However, addition of Ole 2 Ptd- Gro to eggPtdCho produced a titration curve similar to that of the other LUV containing anionic lipids. The two modes of binding observed could be explained by aggregation at the high concentrations of peptide required for the ITC experiment and in the presence of high local concentrations of LUVs after each addition. Fig. 5. Extent of leakage from 25 l M LUVs induced by b-17 after 300 s as a function of P/L. The LUVs used were Ole 2 PtdEtn/Ole 2 PtdGro (1 : 2) (j), dielaidoylphosphatidylethanolamine/Ole 2 PtdGro (1 : 2) (m), eggPtdCho/Ole 2 PtdGro (1 : 2) (h), eggPtdCho (e). Fig. 6. Extent of leakage from 25 l M LUVs induced by b-17 after 300 s as a function of P/L. TheLUVsusedwereOle 2 PtdEtn/ Ole 2 PtdGro (1 : 2) (j), Ole 2 PtdEtn/bovine brain phosphatidyl- serine (1 : 2) (h), Ole 2 PtdEtn/Ole 2 PtdSer (1 : 2) (,), Ole 2 PtdEtn/ PamOlePtdSer (1 : 2) (.). Fig. 7. Binding isotherms for NBD-b-17 with LUVs of Ole 2 PtdEtn/Ole 2 PtdGro (1 : 2) (m), dielaidoylphosphatidylethanolamine/ Ole 2 PtdGro (1 : 2) (j), Ole 2 PtdEtn/Ole 2 Ptd- Ser (1 : 2) (d), eggPtdCho/Ole 2 PtdGro (1 : 2) (h). Ó FEBS 2003 Antimicrobial b-peptide and phospholipid vesicles (Eur. J. Biochem. 270) 1245 For this reason, thermodynamic data was not extracted to evaluate binding. The titrations however, allowed discrimination between the behaviour of the peptide in lipids containing Ole 2 PtdGro and with eggPtdCho alone. Discussion In this study we have characterized the interactions of b-17 with model membranes, in order to rationalize the obser- vations that the antimicrobial activities and hemolytic activities of b-17 and those of [Ala8,13,18]-magainin II amide [30], are similar [7,8]. This finding indicates that these peptides must both cross the bacterial cell wall at similar rates. It has been observed that although a cyclic derivative of magainin has similar activity against liposomes when bound to the membrane, its activity against biological membranes can be altered by its ability to access the membrane surface [34]. Both b-17 and [Ala8,13,18]-maga- inin II amide have in common their affinity for anionic lipids, but we found significant differences between b-17 and magainin II or [Ala8,13,18]-magainin II amide in the way these antimicrobial peptides interact with vesicle systems. While both magainin II [29] and [Ala8,13,18]-magainin II amide raise the T H of DPam 2 PtdEtn, b-17 lowers it. The fact that magainin raises the T H is in accord with the suggested pore mechanism [35] in which the peptide has to promote a large positive curvature of the lipid monolayer lining the pore. It is less clear how the promotion of positive curvature relates to the action of the b-17 peptide. The fact that homologous antimicrobial peptides can have different effects on liposomal membranes is not unique to the present paper. It has recently been shown that a group of model Leu-containing diastereomeric peptides micellized vesicles while the Ile-containing diastereomeric homologs fused model membranes [36]. The lipid dependence of leakage is also very different for magainin II and for b-17. In part, this reflects their different curvature tendencies. For example, replacing Ole 2 PtdEtn with dielaidoylphosphatidylethanolamine (Fig. 4A,B), which has less negative intrinsic curvature, reduces the lysis caused by b-17 as can be seen by the lower rate of contents release caused by 5 l M b-17 with dielaidoylphosphatidyl- ethanolamine:Ole 2 PtdGro (1 : 2) compared to 5 l M b-17 with Ole 2 PtdEtn:Ole 2 PtdGro (1 : 2). Similarly 10 l M b-17 is required for complete leakage to be achieved with dielaidoylphosphatidylethanolamine/Ole 2 PtdGro (1 : 2) as opposed to 6.5 l M b-17 with Ole 2 PtdEtn/Ole 2 PtdGro (1 : 2). In contrast, magainin II promotes greater lysis of vesicles when Ole 2 PtdEtn is substituted with dielaidoyl- phosphatidylethanolamine [29]. The effects of changes in lipid composition (to modulate membrane curvature) on liposomal leakage rates have not been studied for [Ala8,13,18]-magainin II amide. Magainin II also exhibits much more lysis with liposomes containing PtdGro than with liposomes containing PtdSer [29]. We found that b-17 (Fig. 6) did not discriminate, in terms of inducing leakage, between vesicles containing Ole 2 PtdGro or Ole 2 PtdSer, even though the binding affinity of b-17 to PtdSer-containing vesicles is lower than for PtdGro-containing vesicles. Magainin II has been shown in several studies to have antitumor activity [37–45]. With the greater interaction of b-17 with PtdSer-containing liposomes, as compared with magainin, one would anticipate that b-17 would be a more potent anticancer agent. Both magainin II and b-17 lyse anionic liposomes with greater potency than they lyse zwitterionic lipsosomes (Fig. 4C,D), and they both bind more strongly to anionic membranes than to the zwitterionic membrane, as seen by ITC. Therefore, their microbial specificity is likely a consequence of improved electrostatic interactions with microbial vs. eukaryotic membranes. For several anti- microbial peptides it is believed that the basis of the selective toxicity towards bacteria is that the peptide is cationic and can therefore partition to a greater extent into microbial membranes, which have exposed anionic lipids, than into eukaryotic cell membranes. Alternatively, the antimicrobial activity could result from interaction of the peptide with a specific component of microbial membranes. Recently it has been shown that the antimicrobial peptide nicin has greater specificity for microbial membranes because of a specific interaction with lipid II, a component of these membranes [46–49]. Like other cationic antimicrobial peptides, b-17 displays specificity for anionic lipids. However, b-17 differs from Fig. 8. ITC titrations at 30 °C. (A) Successive injections of 10 lLof 3.7 m M LUVs of dielaidoylphosphatidylethanolamine/Ole 2 PtdGro (1 : 2) into 12 l M b-17 placed in the calorimeter cell. Similar titrations were obtained with other lipid mixtures containing Ole 2 PtdGro. (B) successive injections of 10 lLof5m M eggPtdCho LUVs into 25 l M of b-17placedinthecell. 1246 R. F. Epand et al.(Eur. J. Biochem. 270) Ó FEBS 2003 several other such antimicrobial peptides, such as magainin II and its analog [Ala8,13,18]-magainin II amide, in that b-17 promotes negative rather than positive membrane curvature. Therefore it is likely that there are differences in the manner in which b-17 affects membrane properties, compared with magainin. One manifestation of these differences is the lack of specificity of interaction of the peptide with PtdSer vs. PtdGro. Despite these mechanistic differences, however, b-17 behaves as a potent antimicrobial peptide having a minimal inhibitory concentration against several bacteria at least as good as or better than that of [Ala8,13,18]-magainin II amide [8]. [Ala8,13,18]-magainin II amide is an analog of the natural peptide magainin II that has a greater antimicrobial activity [30]. This analog is more hydrophobic than magainin II, having three Ala substitu- ting for a Ser and two Gly. The antimicrobial activity of this analog for several bacteria is of a similar potency to its hemolytic activity [8] in contrast to magainin II where there is approximately 100-fold greater antimicrobial activity compared with hemolytic activity [50]. Being more hydro- phobic, [Ala8,13,18]-magainin II amide will likely penetrate more deeply into the membrane than the parent peptide, magainin II. Deeper membrane penetration also facilitates the promotion of negative curvature, as the regions of the bilayer below the pivotal plane are expanded more than regions close to the interface. Thus, although [Ala8,13,18]- magainin II amide still induces positive curvature effects on membranes, as measured by shifts in T H , its effects on membranes would tend to be somewhat different from that of magainin II. When the [Ala8,13,18]-magainin II amide analog is compared with b-17 by HPLC analysis, both peptides have similar overall hydrophobicities, with similar retention times. However, the effects of these peptides on the T H of DPam 2 PtdEtn suggest that the b-17hasagreater depth of penetration. Nevertheless the two peptides exhibit similar microbial specificity. Acknowledgements This work was supported by the Canadian Institutes of Health Research, Grant 7654 and by the US National Institutes of Health (GM56414). N.U. was supported in part by a Research Fellowship for Young Scientists from the Japan Society for the Promotion of Science, and E.A.P. was supported in part by a Biotechnology Training Grant from NIGMS. References 1. Lohner, K. (2001) Development of Novel Antimicrobial Agents: Emerging Strategies. Horizon Scientific Press, Norfolk, England. 2. Epand, R.M. & Vogel, H.J. (1999) Diversity of antimicrobial peptides and their mechanisms of action. Biochim. Biophys. Acta 1462, 11–28. 3. Hamuro, Y., Schneider, J.P. & DeGrado, W.F. (1999) De novo design of antibacterial beta-peptides. J. Am. Chem. Soc. 121, 12200–12201. 4. Arvidsson, P.I., Frackenpohl, J., Ryder, N.S., Liechty, B., Peter- sen, F., Zimmermann, H., Camenisch, G.P., Woessner, R. & Seebach, D. (2001) On the antimicrobial and hemolytic activities of amphiphilic beta-peptides. Chembiochem. 2, 771–773. 5. Liu, D. & DeGrado, W.F. (2001) De novo design, synthesis and characterization of antimicrobial-b-peptides. J. Am. Chem. Soc. 123, 7553–7559. 6. Cheng, R.P., Gellman, S.H. & DeGrado, W.F. (2001) Beta-peptides: from structure to function. Chem. Rev. 101, 3219– 3232. 7. Porter, E.A., Wang, X., Lee, H.S., Weisblum, B. & Gellman, S.H. (2000) Non-haemolytic beta-amino-acid oligomers. Nature 404, 565. 8. Porter,E.A.,Weisblum,B.&Gellman,S.H.(2002)Mimicryof host-defense peptides by unnatural oligomers: Antimicrobial beta- peptides. J. Am. Chem. Soc. 124, 7324–7330. 9. Hintermann, T. & Seebach, D. (1997) The biological stability of beta–peptides: No interactions between alpha- and beta-peptidic structures. Chimia 51, 244–247. 10. Seebach, D., Abele, S., Schreiber, J.V., Martinoni, B., Nussbaum, A.K.,Schild,H.,Schulz,H.,Hennecke,H.,Woessner,R.& Bitsch, F. (1998) Biological and pharmacokinetic studies with beta-peptides. Chimia 52, 734–739. 11. Matsuzaki, K., Nakamura, A., Murase, O., Sugishita, K., Fujii, N. & Miyajima, K. (1997) Modulation of magainin 2–lipid bilayer interactions by peptide charge. Biochemistry 36, 2104–2111. 12. Appella, D.H., Christianson, L.A., Klein, D.A., Powell, D.R., Huang, X., Barchi, J.J. Jr & Gellman, S.H. (1997) Residue-based control of helix shape in beta-peptide oligomers. Nature 387, 381–384. 13. Appella, D.H., Christianson, L.A., Klein, D.A., Richards, M.R., Powell, D.R. & Gellman, S.H. (1999) Synthesis and structural characterization of helix-forming beta-peptides: trans-2-amino- cyclopentanecarboxylic acid oligomers. J. Am. Chem. Soc. 121, 7574–7581. 14. Barchi, J.J., Huang, X.L., Appella, D.H., Christianson, L.A., Durell, S.R. & Gellman, S.H. (2000) Solution conformations of helix-forming beta-amino acid homooligomers. J. Am. Chem. Soc. 122, 2711–2718. 15. Wang, X.F., Espinosa, J.F. & Gellman, S.H. (2000) 12-helix for- mation in aqueous solution with short beta-peptides containing pyrrolidine-based residues. J. Am. Chem. Soc. 122, 4821–4822. 16. Kobayashi, S., Takeshima, K., Park, C.B., Kim, S.C. & Matsu- zaki, K. (2000) Interactions of the novel antimicrobial peptide buforin 2 with lipid bilayers: proline as a translocation promoting factor. Biochemistry 39, 8648–8654. 17. Matsuzaki, K. (1998) Magainins as paradigm for the mode of action of pore forming polypeptides. Biochim. Biophys. Acta 1376, 391–400. 18. Matsuzaki, K. (1999) Why and how are peptide–lipid interactions utilized for self-defense? Magainins and tachyplesins as arche- types. Biochim. Biophys. Acta 1462, 1–10. 19. Shai, Y. & Oren, Z. (2001) From ÔcarpetÕ mechanism to de-novo designed diastereomeric cell-selective antimicrobial peptides. Peptides 22, 1629–1641. 20. Lee, H.S., LePlae, P.R., Porter, E.A. & Gellman, S.H. (2001) An efficient route to either enantiomer of orthogonally protected trans-3-aminopyrrolidine-4-carboxylic acid. J. Org. Chem. 66, 3597–3599. 21. LePlae, P.R., Umezawa, N., Lee, H.S. & Gellman, S.H. (2001) An efficient route to either enantiomer of trans-2-aminocyclopen- tanecarboxylic acid. J. Org. Chem. 66, 5629–5632. 22. Ames, B.N. (1966) Assay of inorganic phosphate, total phosphate and phosphatases. Methods Enzymol. 8, 115–118. 23. Ellens, H., Bentz, J. & Szoka, F.C. (1985) H + -andCa 2+ -induced fusion and destabilization of liposomes. Biochemistry 24, 3099– 3106. 24. Rajarathnam, K., Hochman, J., Schindler, M. & Ferguson-Miller, S. (1989) Synthesis, location, and lateral mobility of fluorescently labeled ubiquinone 10 in mitochondrial and artificial membranes. Biochemistry 28, 3168–3176. 25. Kim, J., Blackshear, P.J., Johnson, J.D. & McLaughlin, S. (1994) Phosphorylation reverses the membrane association of peptides Ó FEBS 2003 Antimicrobial b-peptide and phospholipid vesicles (Eur. J. Biochem. 270) 1247 that correspond to the basic domains of MARCKS and neuro- modulin. Biophys. J. 67, 227–237. 26. Peitzsch, R.M. & McLaughlin, S. (1993) Binding of acylated peptides and fatty acids to phospholipid vesicles: pertinence to myristoylated proteins. Biochemistry 32, 10436–10443. 27. Ladokhin, A.S., Selsted, M.E. & White, S.H. (1997) Bilayer interactions of indolicidin, a small antimicrobial peptide rich in tryptophan, proline, and basic amino acids. Biophys. J. 72, 794– 805. 28. Epand, R.M. (1985) Diacylglycerols, lysolecithin, or hydrocar- bons markedly alter the bilayer to hexagonal phase transition temperature of phosphatidylethanolamines. Biochemistry 24, 7092–7095. 29. Matsuzaki, K., Sugishita, K., Ishibe, N., Ueha, M., Nakata, S., Miyajima, K. & Epand, R.M. (1998) Relationship of membrane curvature to the formation of pores by magainin 2. Biochemistry 37, 11856–11863. 30. Chen, H.C., Brown, J.H., Morell, J.L. & Huang, C.M. (1988) Synthetic magainin analogues with improved antimicrobial activity. FEBS Lett. 236, 462–466. 31. Applequist, J., Bode, K.A., Appella, D.H., Christianson, L.A. & Gellman, S.H. (1998) Theoretical and experimental circular dichroic spectra of the novel helical foldamer poly[(1R,2R)-trans- 2-aminocyclopentanecarboxylic acid]. J. Am. Chem. Soc. 120, 4891–4892. 32. Lee, H.S., Syud, F.A., Wang, X.F. & Gellman, S.H. (2001) Diversity in short beta-peptide 12-helices: High-resolution struc- tural analysis in aqueous solution of a hexamer containing sulfo- nylated pyrrolidine residues. J. Am. Chem. Soc. 123, 7721–7722. 33. Epand, R.F. & Scheraga, H.A. (1968) The helix-coil transition of poly- L -lysine in methanol-water solvent mixtures. Biopolymers 6, 1383–1386. 34. Unger, T., Oren, Z. & Shai, Y. (2001) The effect of cyclization of magainin 2 and melittin analogues on structure, function, and model membrane interactions: implication to their mode of action. Biochemistry 40, 6388–6397. 35. Matsuzaki, K., Murase, O. & Miyajima, K. (1995) Kinetics of pore formation by an antimicrobial peptide, magainin 2, in phospholipid bilayers. Biochemistry 34, 12553–12559. 36. Avrahami, D., Oren, Z. & Shai, Y. (2001) Effect of multiple ali- phatic amino acids substitutions on the structure, function, and mode of action of diastereomeric membrane active peptides. Biochemistry 40, 12591–12603. 37. Baker, M.A., Maloy, W.L., Zasloff, M. & Jacob, L.S. (1993) Anticancer efficacy of magainin 2 and analogue peptides. Cancer Res. 53, 3052–3057. 38. Cruciani, R.A., Barker, J.L., Zasloff, M., Chen, H.C. & Cola- monici, O. (1991) Antibiotic magainins exert cytolytic activity against transformed cell lines through channel formation. Proc. NatlAcad.Sci.USA88, 3792–3796. 39. Haimovich, B. & Tanaka, J.C. (1995) Magainin-induced cyto- toxicity in eukaryotic cells: kinetics, dose–response and channel characteristics. Biochim. Biophys. Acta 1240, 149–158. 40. Jacob, L. & Zasloff, M. (1994) Potential therapeutic applications of magainins and other antimicrobial agents of animal origin. Ciba Found. Symp. 186, 197–216. 41. Ohsaki, Y., Gazdar, A.F., Chen, H.C. & Johnson, B.E. (1992) Antitumor activity of magainin analogues against human lung cancer cell lines. Cancer Res. 52, 3534–3538. 42. Shin, S.Y., Lee, M.K., Kim, K.L. & Hahm, K.S. (1997) Structure- antitumor and hemolytic activity relationships of synthetic pep- tides derived from cecropin A-magainin 2 and cecropin A-melittin hybrid peptides. J. Pept. Res. 50, 279–285. 43. Shin, S.Y., Kang, J.H., Jang, S.Y., Kim, Y., Kim, K.L. & Hahm, K.S. (2000) Effects of the hinge region of cecropin A(1–8)-magainin 2 (1–12), a synthetic antimicrobial peptide, on liposomes, bacterial and tumor cells. Biochim. Biophys. Acta 1463, 209–218. 44. Soballe, P.W., Maloy, W.L., Myrga, M.L., Jacob, L.S. & Herlyn, M. (1995) Experimental local therapy of human melanoma with lytic magainin peptides. Int. J. Cancer 60, 280–284. 45. Westerhoff, H.V., Zasloff, M., Rosner, J.L., Hendler, R.W., De Waal, A., Vaz, G.A., Jongsma, P.M., Riethorst, A. & Juretic, D. (1995) Functional synergism of the magainins PGLa and magai- nin 2 in Escherichia coli, tumor cells and liposomes. Eur.J.Bio- chem. 228, 257–264. 46. Breukink, E., Bonev, B.B., Wiedemann, I., Sahl, H.G., Watts, A. & de Kruijff, B. (2001) Specific interaction of the lantibiotic nisin with lipid II leads to highly efficient pore formation. Biophys. J. 80, 2A. 47. Breukink, E., Ganz, P., de Kruijff, B. & Seelig, J. (2000) Binding of nisin Z to bilayer vesicles as determined with isothermal titration calorimetry. Biochemistry 39, 10247–10254. 48. Breukink, E., Wiedemann, I., van Kraaij, C., Kuipers, O.P., Sahl, H.G. & de Kruijff, B. (1999) Use of the cell wall precursor lipid II by a pore-forming peptide antibiotic. Science 286, 2361– 2364. 49. Wiedemann, I., Breukink, E., van Kraaij, C., Kuipers, O.P., Bierbaum, G., de Kruijff, B. & Sahl, H.A. (2001) Specific binding of nisin to the peptidoglycan precursor lipid II combines pore formation and inhibition of cell wall biosynthesis for potent antibiotic activity. J. Biol. Chem. 276, 1772–1779. 50. Kobayashi, S., Hirakura, Y. & Matsuzaki, K. (2001) Bacteria- selective synergism between the antimicrobial peptides alpha- helical magainin 2 and cyclic beta-sheet tachyplesin I: toward cocktail therapy. Biochemistry 40, 14330–14335. 1248 R. F. Epand et al.(Eur. J. Biochem. 270) Ó FEBS 2003 . Interactions of the antimicrobial b-peptide b-17 with phospholipid vesicles differ from membrane interactions of magainins Raquel F. Epand 1 , Naoki Umezawa 2 ,. that there are differences in the manner in which b-17 affects membrane properties, compared with magainin. One manifestation of these differences is the lack of specificity of interaction of the peptide. predominates. If b-17 promoted the formation of a similar pore arrange- ment, but the diameter of the pore were smaller than that of magainin, then the negative curvature in the dimension around the rim of the

Ngày đăng: 31/03/2014, 07:20

Từ khóa liên quan

Tài liệu cùng người dùng

Tài liệu liên quan