biochimica et biophysica acta - connecting … this peptide, its release gives the bacterium a...

12
Antimicrobial HPA3NT3 peptide analogs: Placement of aromatic rings and positive charges are key determinants for cell selectivity and mechanism of action Jong-Kook Lee a, 1 , Seong-Cheol Park a, 1 , Kyung-Soo Hahm b , Yoonkyung Park a, c, a Research Center for Proteinaceous Materials (RCPM), Chosun University, Kwangju 501759, Republic of Korea b Bioleaders Corporation, 559 Yongsan-Dong, Yusong-Ku, Daejeon 305500, Republic of Korea c Department of Biotechnology, Chosun University, Kwangju 501759, Republic of Korea abstract article info Article history: Received 1 June 2012 Received in revised form 6 September 2012 Accepted 7 September 2012 Available online 13 September 2012 Keywords: Determinants of cell selectivity Antimicrobial action Resistance In an earlier study, we determined that HP(220) (residues 220 of parental HP derived from the N-terminus of the Helicobacter pylori ribosomal protein L1) and its analog, HPA3NT3, had potent antimicrobial effects. However, HPA3NT3 also showed undesirable cytotoxicity against HaCaT cells. In the present study, we designed peptide ana- logs including HPA3NT3-F1A (F1A), HPA3NT3-F8A (F8A), HPA3NT3-F1AF8A (F1AF8A), HPA3NT3-A1 (A1) and HPA3NT3-A2 (A2) in an effort to investigate the effects of amino acid substitutions in reducing their hydrophobicity or increasing their cationicity, and any resulting effects on their selectivity in their interactions with human cells and pathogens, as well as their mechanism of antimicrobial action. With the exception of HPA3NT3-A1, all of these pep- tides showed potent antimicrobial activity. Moreover, substitution of Ala for Phe at positions 1 and/or 8 of the HPA3NT3 peptides (F1A, -F8A and -F1AF8A) dramatically reduced their cytotoxicity. Thus the cytotoxicity of HPA3NT3 appears to be related to its Phe residues (positions 1 and 8), which strongly interact with sphingomyelin in the mammalian cell membrane. HPA3NT3 exerted its bactericidal effects through membrane permeabilization mediated by pore formation. In contrast, uorescent dye leakage and nucleic acid gel retardation assays showed that A2 acted by penetrating into the cytoplasm, where it bound to nucleic acids and inhibited protein synthesis. Notably, Staphylococcus aureus did not develop resistance to -A2 as it did with rifampin. These results suggest that the -A2 peptide could potentially serve as an effective antibiotic agent against multidrug-resistant bacterial strains. © 2012 Elsevier B.V. All rights reserved. 1. Introduction Antimicrobial peptides (AMPs) are small, cationic peptides that play an important role in the innate immune system [1]. In many species, they provide the rst line of immune defense against microbial infec- tion [2,3]. AMPs have advantages over small-molecule antibiotics in that they possess multiple modes of action, rapid kill kinetics, low levels of resistance-inducing behavior and little host toxicity. In addition, un- like conventional antibiotics, which have denite intracellular targets, AMPs generally do not have specic microbial cell targets; instead, they bind to bacterial cell membranes and perturb the membrane struc- ture. It is this mechanism of action that makes it unlikely that AMPs would induce bacterial resistance, since microbes would need to change their entire membrane lipid composition to evade them. Consequently, AMPs are an attractive therapeutic candidate for protection against drug-resistant organisms. Stomach mucosa infected with Helicobacter pylori, the bacterial pathogen associated with gastritis and peptic ulcers, typically shows massive inltration of inammatory cells and tissue destruction [4]. It has been suggested that the persistence of H. pylori within the mu- cosa is facilitated by cecropin-like peptides with antibacterial proper- ties, which are produced by the bacteria. Because H. pylori is resistant to this peptide, its release gives the bacterium a competitive advan- tage over other microorganisms [5]. The linear antimicrobial peptide HP(220) is a cationic α-helical peptide isolated from the N-terminal region of the H. pylori ribosomal protein L1 [6]. This peptide possesses several important functional char- acteristics: it is bactericidal, it is a neutrophil chemoattractant, and it ac- tivates phagocyte NADPH oxidase to produce reactive oxygen species [6]. In addition, we reported that HPA3NT3, a peptide derived from HP(220), also exerts potent antibacterial effects. The radius of the pore in the membrane of HP(220) was observed to be approximately 1.8 nm, whereas HPA3NT3, like melittin, has an apparent radius be- tween 3.3 and 4.8 nm [7]. Unfortunately, at higher concentrations HPA3NT3 is cytotoxic and is thus not suitable for in vivo use. Therefore, Biochimica et Biophysica Acta 1828 (2013) 443454 Abbreviations: Fmoc, 9-uorenylmethoxycarbonyl; DMF, dimethylformamide; DCM, dichloromethane; hRBCs, human red blood cells; HaCaT, human keratinocyte; CLSM, Confocal laser-scanning microscopy; ATCC, American Type Culture Collection; KCTC, Korean Collection for Type Cultures; CCARM, Collection of Antibiotics Resistance Microbes; CFU, colony forming unit; MIC, minimum inhibitory concentration; TAMRA, tetramethyl rhodamine; SUVs, small unilamellar vesicles; LUVs, large unilamellar ves- icles; TEM, transmission electron microscopy; UAC, uranyl acetate Corresponding author at: Department of Biotechnology, Chosun University, Kwangju 501759, Republic of Korea. Tel.: +82 62 2306854; fax: +82 62 2256758. E-mail address: [email protected] (Y. Park). 1 Share equal rst authorship. 0005-2736/$ see front matter © 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.bbamem.2012.09.005 Contents lists available at SciVerse ScienceDirect Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbamem

Upload: truongnhan

Post on 11-Jun-2019

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Biochimica et Biophysica Acta 1828 (2013) 443–454

Contents lists available at SciVerse ScienceDirect

Biochimica et Biophysica Acta

j ourna l homepage: www.e lsev ie r .com/ locate /bbamem

Antimicrobial HPA3NT3 peptide analogs: Placement of aromatic rings and positivecharges are key determinants for cell selectivity and mechanism of action

Jong-Kook Lee a,1, Seong-Cheol Park a,1, Kyung-Soo Hahm b, Yoonkyung Park a,c,⁎a Research Center for Proteinaceous Materials (RCPM), Chosun University, Kwangju 501‐759, Republic of Koreab Bioleaders Corporation, 559 Yongsan-Dong, Yusong-Ku, Daejeon 305‐500, Republic of Koreac Department of Biotechnology, Chosun University, Kwangju 501‐759, Republic of Korea

Abbreviations: Fmoc, 9-fluorenylmethoxycarbonyDCM, dichloromethane; hRBCs, human red blood cellsCLSM, Confocal laser-scanning microscopy; ATCC, AmeKCTC, Korean Collection for Type Cultures; CCARM, ColleMicrobes; CFU, colony forming unit; MIC, minimum inhtetramethyl rhodamine; SUVs, small unilamellar vesicleicles; TEM, transmission electron microscopy; UAC, ura⁎ Corresponding author at: Department of Biotechnol

Kwangju 501‐759, Republic of Korea. Tel.: +82 62 2306E-mail address: [email protected] (Y. Park).

1 Share equal first authorship.

0005-2736/$ – see front matter © 2012 Elsevier B.V. Allhttp://dx.doi.org/10.1016/j.bbamem.2012.09.005

a b s t r a c t

a r t i c l e i n f o

Article history:Received 1 June 2012Received in revised form 6 September 2012Accepted 7 September 2012Available online 13 September 2012

Keywords:Determinants of cell selectivityAntimicrobial actionResistance

In an earlier study, we determined that HP(2‐20) (residues 2‐20 of parental HP derived from the N-terminus of theHelicobacter pylori ribosomal protein L1) and its analog, HPA3NT3, had potent antimicrobial effects. However,HPA3NT3 also showed undesirable cytotoxicity against HaCaT cells. In the present study, we designed peptide ana-logs including HPA3NT3-F1A (‐F1A), HPA3NT3-F8A (‐F8A), HPA3NT3-F1AF8A (‐F1AF8A), HPA3NT3-A1 (‐A1) andHPA3NT3-A2 (‐A2) in an effort to investigate the effects of amino acid substitutions in reducing their hydrophobicityor increasing their cationicity, and any resulting effects on their selectivity in their interactionswith human cells andpathogens, as well as their mechanism of antimicrobial action.With the exception of HPA3NT3-A1, all of these pep-tides showed potent antimicrobial activity. Moreover, substitution of Ala for Phe at positions 1 and/or 8 of theHPA3NT3 peptides (‐F1A, -F8A and -F1AF8A) dramatically reduced their cytotoxicity. Thus the cytotoxicity ofHPA3NT3 appears to be related to its Phe residues (positions 1 and 8), which strongly interact with sphingomyelinin the mammalian cell membrane. HPA3NT3 exerted its bactericidal effects through membrane permeabilizationmediated by pore formation. In contrast, fluorescent dye leakage and nucleic acid gel retardation assays showedthat ‐A2 acted by penetrating into the cytoplasm, where it bound to nucleic acids and inhibited protein synthesis.Notably, Staphylococcus aureus did not develop resistance to -A2 as it did with rifampin. These results suggest thatthe -A2 peptide could potentially serve as an effective antibiotic agent against multidrug-resistant bacterial strains.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction

Antimicrobial peptides (AMPs) are small, cationic peptides that playan important role in the innate immune system [1]. In many species,they provide the first line of immune defense against microbial infec-tion [2,3]. AMPs have advantages over small-molecule antibiotics inthat they possessmultiplemodes of action, rapid kill kinetics, low levelsof resistance-inducing behavior and little host toxicity. In addition, un-like conventional antibiotics, which have definite intracellular targets,AMPs generally do not have specific microbial cell targets; instead,they bind to bacterial cellmembranes and perturb themembrane struc-ture. It is this mechanism of action that makes it unlikely that AMPs

l; DMF, dimethylformamide;; HaCaT, human keratinocyte;rican Type Culture Collection;ction of Antibiotics Resistanceibitory concentration; TAMRA,s; LUVs, large unilamellar ves-nyl acetateogy, Chosun University,854; fax: +82 62 2256758.

rights reserved.

would induce bacterial resistance, sincemicrobeswould need to changetheir entire membrane lipid composition to evade them. Consequently,AMPs are an attractive therapeutic candidate for protection againstdrug-resistant organisms.

Stomach mucosa infected with Helicobacter pylori, the bacterialpathogen associated with gastritis and peptic ulcers, typically showsmassive infiltration of inflammatory cells and tissue destruction [4].It has been suggested that the persistence of H. pylori within the mu-cosa is facilitated by cecropin-like peptides with antibacterial proper-ties, which are produced by the bacteria. Because H. pylori is resistantto this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5].

The linear antimicrobial peptide HP(2‐20) is a cationic α-helicalpeptide isolated from the N-terminal region of the H. pylori ribosomalprotein L1 [6]. This peptide possesses several important functional char-acteristics: it is bactericidal, it is a neutrophil chemoattractant, and it ac-tivates phagocyte NADPH oxidase to produce reactive oxygen species[6]. In addition, we reported that HPA3NT3, a peptide derived fromHP(2‐20), also exerts potent antibacterial effects. The radius of thepore in the membrane of HP(2‐20) was observed to be approximately1.8 nm, whereas HPA3NT3, like melittin, has an apparent radius be-tween 3.3 and 4.8 nm [7]. Unfortunately, at higher concentrationsHPA3NT3 is cytotoxic and is thus not suitable for in vivo use. Therefore,

Page 2: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

444 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

we sought a new peptide that exhibited no cytotoxicity at high concen-trations, yet retained its antibacterial activity. To that end, we designedand synthesized a set of HPA3NT3 analogs through amino acid substitu-tion. The antibiotic effects of the synthesized peptides were measuredagainst Gram-positive and Gram-negative bacteria and pathogenic fun-gal cells. The cytotoxicity of the peptides against human red blood cells(hRBCs), human keratinocyte cells, human cell membrane componentsand sphingomyelin (SM)was also examined. Finally, themode of actionof the peptides in amodel lipid bilayer and in bacterial and fungalmem-branes was investigated using various techniques.

2. Materials and methods

2.1. Materials

L-α-phosphatidylethanolamine (PE), sphingomyelin (SM), cho-lesterol (CH), L-α-phosphatidyl-DL-glycerol (PG) and egg yolk L-α-phosphatidylcholine (PC) were obtained from Avanti Polar Lipids(Alabaster, AL). 3,3′-diethylthio-dicarbocyanine iodide (DiSC3-5),carboxytetramethylrhodamine succinimidyl ester (TAMRA) andcalcein were from Molecular Probes (Eugene, OR). Plasmid DNA(pTYB2) was from New England Biolabs Ltd. Erythromycin, ampicil-lin, ciprofloxacin, rifampin and vancomycin were from Sigma-Aldrich Co. All other reagents used were of analytical grade.

2.2. Microbial strains

Escherichia coli (ATCC 25922), Staphylococcus aureus (ATCC 25923),S. epidermidis (ATCC 12228) and Pseudomonas aeruginosa (ATCC15692) were obtained from American Type Culture Collection.Bacillus subtilis (KCTC 1998), Listeria monocytogenes (KCTC 3710),Proteus vulgaris (KCTC 2433), Salmonella typhimurium (KCTC 1926),Aspergillus awamori (KCTC 6915), A. parasiticus (KCTC 6598), A. flavus(KCTC 6905), A. fumigates (KCTC 6145), Candida albicans (KCTC 7270)and Trichosporon beigellii (KCTC 7707) were from the Korean Collec-tion for Type Cultures. E. coli CCARM 1229, E. coli CCARM 1238, S. au-reus CCARM 3089, S. aureus CCARM 3114, Sal. typhimurium CCARM8009 and Sal. typhimurium CCARM 8013 were from the Culture Col-lection of Antibiotics Resistant Microbes at Seoul Women's Universi-ty, Republic of Korea. P. aeruginosa 4007 and 3547 were resistantstrains isolated from patients in a hospital using otitis media. Therifampin-resistant S. aureus PBEL 1 and vancomycin-resistant S. aureusPBEL 2 used in this study were induced from normal S. aureus (ATCC25923).

2.3. Peptide synthesis, fluorescent labeling and purification

Peptides were synthesized using 9-fluorenylmethoxycarbonyl(Fmoc) solid-phasemethods on Rink amide 4-methyl benzhydrylamineresin (Novabiochem) (0.55 mmol/g) and a Liberty microwave peptidesynthesizer (CEM Co. Matthews, NC). To generate N-terminal fluores-cently labeled peptides, resin-bound peptides were treated with 20%(v/v) piperidine in dimethylformamide (DMF) to remove the Fmocprotecting group from theN-terminal amino acid. The resin-boundpep-tide was then reacted with rhodamine-SE in DMF (3–4 eq.) containing5% (v/v) diisopropylethylamine. After gently mixing for 24 h in thedark, the resins were washed thoroughly, first with DMF and thenwith dichloromethane (DCM). The peptides were cleaved from their re-spective resins, precipitated with ether, and extracted [8,9]. The crudepeptides were purified using reversed-phase preparative HPLC on aJupiter C18 column (250×21.2 mm, 15 μm, 300 Å) with an appropriate0–60% acetonitrile gradient in water containing 0.05% trifluoroaceticacid. The purity of the peptide was then determined by analyticalreversed-phase HPLC using a Jupiter Proteo C18 column (250×4.6 mm,90 Å, 4 μm). The molecular masses of the peptides were confirmed

using matrix-assisted laser desorption ionization mass spectrometry(MALDI II, Kratos Analytical Ins.) [7].

2.4. Antibacterial activity

Bacterial cells were cultured at 37 °C in appropriate culture media,and the antimicrobial activity of each peptide was determined inmicrodilution assays. Briefly, two-fold serial dilutions covering arange from 0.25 to 128 μM for each peptide were added to duplicatemedia containing cultures of test bacteria (5×105 CFUs/ml) at theirmid-logarithmic phase of growth. The samples were then incubatedfor 18–24 h at 37 °C. At the end of the incubation, minimum inhibito-ry concentrations (MICs) of the peptides were determined by evalu-ating growth based on the OD600 of the cultures. MICs were definedas the lowest peptide concentrations that completely inhibited bacte-rial growth [7].

2.5. Antifungal activity

In order to evaluate the activity of the peptides against various fun-gal pathogens, spores from 10-day-old cultures grown on PDA (potatodextrose agar) plates at 28 °C were collected using 0.08% TritonX-100. The spores were then suspended in a medium containing half10 mMMES buffer and half YPD (yeast peptone dextrose media: yeastextract 0.5%, peptone 1% and dextrose 2% at pH 5.0–5.5), and then incu-bated overnight. Eighty μL of the spore suspension (final concentrationof 104 spores/ml) in PD (potato dextrose)mediawas added to thewellsof a 96-well microtiter plate, afterwhich 20 μL of two-fold serially dilut-ed peptide solution in 10 mMMES buffer (pH 6.0) containing YPDor PDmedia was added to the wells. After 24 to 36 h of incubation at 28 °C,inhibition of fungal growth or germination was evaluated visuallyusing an inverted microscope, and by assessing the turbidity of eachwell by measuring the absorbance at 595 nm using a microtiter reader[10].

2.6. Hemolysis

Fresh hRBCs from healthy donors were centrifuged at 800×g andwashed with PBS until the supernatant was clear. Two-fold serial di-lutions of peptide in PBS were added to the wells of a 96-well plateplated, after which hRBCs were added to a final concentration of 8%(v/v). The samples were then incubated with mild agitation for 1 hat 37 °C. After then centrifuging the samples at 800×g for 10 min,the absorbance of the supernatant was measured at 414 nm. All mea-surements were made in triplicate [11,12], and the percent hemolysiswas calculated using Eq. (1):

% hemolysis ¼ Abs414 in the peptide solution−Abs414 inPBSð Þ=Abs414 in 0:1% Triton−X100−Abs414 in PBSð Þ� � 100

ð1Þ

where, 100% hemolysis was defined as the absorbance measuredfrom hRBCs exposed to 1% Triton X-100, and zero hemolysis wascharacterized with hRBCs alone in PBS.

2.7. Cell culture and cytotoxicity

To examine the cytotoxic effects of the peptides, HaCaT (culturedhuman keratinocyte) cells were cultured in Dulbecco's modified Eaglemedium (DMEM) supplemented with antibiotics (100 U/ml penicillin,100 μg/ml streptomycin) and 10% fetal calf serum at 37 °C in a humid-ified chamber under an atmosphere containing 5% CO2. Growth inhibi-tion was evaluated using MTT assays to assess cell viability. A total of4×103 cells/well was seeded onto a 96-well plate and incubated for24 h. Two-fold serial dilutions of peptides in DMEM were then addedto the plate and incubated for 24 h at 37 °C, after which 10 μL of

Page 3: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

445J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

5 mg/ml MTT was added to each well and incubated for an additional4 h. The supernatants were then aspirated, and 50 μl of DMSO wasadded to the wells to dissolve any remaining precipitate. Absorbanceat 570 nm was then measured using a microtiter reader [10,12].

2.8. Peptide-evoked membrane depolarization

E. coli were grown to mid-log phase at 37 °C with agitation, afterwhich they were washed once in buffer A (20 mM glucose, 5 mMHEPES, pH7.3) and resuspended to anOD600 of 0.05 in buffer A containing0.1 M KCl. The cells were then incubated with 1 μM DiSC3-5 until stablebaseline fluorescence was achieved. Peptides were then added to 480 μlof the bacterial suspension, afterwhich changes influorescencewere con-tinuously recorded (excitation wavelength, 622 nm; emission wave-length, 670 nm) for 30 min [7].

2.9. Time-kill kinetics of the peptides

The bactericidal kinetics of the peptides was assessed for compar-ison with the membrane depolarization activity. Suspensions of E. coli(OD600=0.05) were added to peptide solutions at the same concen-trations used in the membrane depolarization experiment. Bacteriawere then sampled after 0, 2, 4, 6, 8, 10, 15, 20, 30, 40 and 60 minof exposure to the peptides. The sampled bacteria were then diluted20-fold and plated on Luria broth agar. After incubation overnight,the colonies were counted [13].

2.10. SYTOX green uptake

E. coli cells were grown to mid-logarithmic phase at 37 °C, washed,and suspended (2×107 cells/ml) in 10 mM sodium phosphate buffer(pH 7.2). The cells were then incubated with 1 μM SYTOX green for15 min in the dark [14]. After the addition of peptides at the appropriateconcentrations, the time-dependent increases in fluorescence causedby the binding of the cationic dye to intracellular DNA were monitored(excitation wavelength, 485 nm; emission wavelength, 520 nm).

2.11. Confocal laser-scanning microscopy

To analyze the cellular distribution of the peptides, E. coli andC. albicans were incubated in the presence of tetramethyl rhodamine(TAMRA)-labeled peptide and observed on a confocal laser-scanningmicroscope. The cells were inoculated and adjusted following the pro-cedure used for the antimicrobial assay. TAMRA‐labeled peptides attheir respective MICs were added to 100 μl of the cell suspension.After incubation for 30 min, the cells were pelleted by centrifugationat 4000 rpm for 5 min andwashed three timeswith ice-cold PBS buffer.Localization of the TAMRA-labeled peptides was then examined usingan inverted LSM510 laser-scanning microscope (Carl Zeiss, Gőttingen,Germany) equipped with helium/neon laser (543 nm line). Theresulting images were recorded digitally in a 512×512 pixel format[15].

2.12. Preparation of liposomes and measurement of calcein leakage

Small and large unilamellar vesicles (SUVs and LUVs) were pre-pared using the sonication and freeze-thaw methods, respectively[16,17]. The desired mixtures of lipids were dissolved in chloroform,dried in a glass tube under nitrogen, and then lyophilized overnightto remove residual solvent. The dry lipid films were resuspended in1–2 ml of appropriate buffer in a water-bath for 30 min at 50 °C,and the resulting mixtures were vortexed. To prepare the SUVs, thelipid dispersions were sonicated in a bath-type sonicator for 30 minuntil the turbidity had cleared. LUVs were prepared using eightfreeze-thaw cycles under liquid nitrogen in a water bath at 50 °C.To ensure a homogeneous population of vesicles, after their

preparation, the suspensions were extruded 15 times through poly-carbonate membranes with 0.05 and 0.2 μm diameter pores forSUVs and LUVs, respectively. Their concentrations were determinedusing a standard phosphate assay [18]. The average diameters of thevesicles were confirmed using transmission electron microscopy(FEI Tecnai 12).

The ability of the peptides to permeabilize liposomeswas assayed bymeasuring calcein leakage. Liposomeswith entrapped calceinwere pre-pared as described previously. In brief, the dried lipid was hydratedwith dye buffer solution (80 mM calcein, 10 mM HEPES, 30 mM NaCl,pH 7.4), vortexed for 1 min and incubated for 30 min at 50 °C. Thesuspension was then subjected to nine cycles of freeze-thaw in liquidnitrogen to create LUVs, and then extruded 30 times through polycar-bonate filters (two stacked 0.2-μm pore size filters) using an AvantiMini-Extruder (Avanti Polar Lipids Inc., Alabaster, AL). The calcein-containing LUVs were separated from the free calcein by gel filtrationchromatography on a Sephadex G-50 column, after which suspensionscontaining 2.5 μM lipid were incubated with various concentrations ofpeptide. The fluorescence of the released calcein was measured usinga spectrofluorometer (Perkin-Elmer LS55) at an excitation wavelengthof 480 nm and an emission wavelength of 520 nm. Complete (100%)release of calcein was achieved by adding Triton X-100 to a final con-centration of 0.03% (w/v). Spontaneous leakage was determined to benegligible in the time-frame of this experiment. The experimentswere conducted at 25 °C. The apparent percent calcein release was cal-culated using Eq. (2) [7,19]:

Release %ð Þ ¼ 100� F−F0ð Þ= Ft−F0ð Þ ð2Þ

2.13. Tryptophan fluorescence blue shift

The fluorescence emission spectrum of the Trp residues in thepeptides was monitored in aqueous buffer and in SUVs composed ofPE/PG (7:3, w/w), PC/SM (2:1, w/w) or PC/CH (2:1, w/w). Aliquotsof 2 μM peptide solution were added to 1 ml of 10 mM HEPES (pH7.2), or to vesicles at a total lipid concentration in the range from0 to 300 μM. Trp fluorescence was then excited at 280 nm, and theemission spectrumwas measured from 300 to 450 nm in 1-nm incre-ments with 1-s signal averaging. The wavelength with the maximumfluorescence emission was then plotted against the lipid concentra-tion. Trp fluorescence measurements were made using a Perkin-Elmer LS55 fluorometer.

2.14. DNA/RNA gel retardation assay

Plasmid pTYB 2 (NEB Ltd.) was purified using a plasmid extractionkit (Exprep™ Quick, GeneAll Biotechnology Co., Seoul, Republic ofKorea), while yeast RNAwas isolated using TRI reagent (MRC Inc. Cin-cinnati, OH). The resulting plasmid DNA (200 ng) or yeast RNA(10 μg) was mixed with increasing amounts of peptide in buffercontaining 10 mM Tris (pH 8.0), 1 mM EDTA, 5% glycerol, 20 mMKCl and 50 μg/ml BSA. The mixtures were incubated for 10 min at37 °C and then subjected to electrophoresis on a 0.5 or 1% agarosegel in TBE buffer, followed by ethidium bromide staining [20]. Gel re-tardation was visualized under UV illumination using a Bio-Rad GelDocumentation system (UK).

2.15. Transmission electron microscopy

The pTYB2 DNA was expressed in E. coli cells grown to mid-logphase at 37 °C. The cells were then suspended in 10 mM sodiumphosphate buffer to achieve an OD600 of 0.15. They were then sonicat-ed and centrifuged in order to collect the pTYB2 DNA. The pTYB2 DNAwas then negatively stained with 1% (w/v) uranyl acetate (UAC) inthe absence or presence of -A2. The DNA was visualized using a FEI

Page 4: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 1. Helical wheel diagram of HPA3NT3 and its analogs. The analogs were produced through amino acid substitution, as indicated.

446 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

Tecnai 12 transmission electron microscope (TEM) at nominal magni-fications of 67,000×(for OsmC), 52,000×(for Lon, CodWX and SOX)or 40,000×(for ClpB), and using an acceleration voltage of 120 kVand a magnification of 21 K and 52 K, as described [21].

2.16. Inhibition of expression of recombinant green fluorescent protein(GFP) in E. coli

E. coli BL21 (DE3) cells transfected with pET28a-GFP were grownto mid-log phase at 37 °C. After which aliquots were suspended in10 mM sodium phosphate buffer containing 10% LB media to anOD600 of 0.15. Peptides at 1/4-MIC, 1/2-MIC or MIC were added tothe mixtures, which were then incubated for 30 min. Thereafter,0.5 mM of IPTG was added to induce GFP expression, and the mixturewas incubated for 4 h at 37 °C. The cells were then centrifuged andwashed three times to remove unreacted peptide, lysed by sonicationand centrifuged again. The cytosolic proteins in the supernatant werethen visualized using fluorescence microscopy or 15% SDS-PAGE.

2.17. Resistance development

A single colony of S. aureus ATCC 25923 was cultured at 37 °C inappropriate culture media to determine the MICs of the two peptides(LL-37 and -A2), and rifampin and vancomycin, as described above.The final concentrations of the peptides ranged from 256 to0.125 μM, while those of rifampin and vancomycin ranged from 156

Table 1Amino acid sequences, retention times, and mean hydrophobicities (H) of HPA3NT3 and it

Name Sequence Remarks

HPA3NT3 FKRLKKLFKKIWNWK-NH2 Parent peptide-F1A AKRLKKLFKKIWNWK-NH2 HPA3NT3: F1 → A1

-F8A FKRLKKLAKKIWNWK-NH2 HPA3NT3: F8 → A8

-F1AF8A AKRLKKLAKKIWNWK-NH2 HPA3NT3: F1,F8 → A1,A8

-A1 AKRLKKLAKKIWKWL-NH2 -F1AF8A: N13,K15 → K13,L15

-A2 AKRLKKLAKKIWKWK-NH2 -F1AF8A: N13 → K13

a Mean retention times (min) analyzed by RP-HPLC.b Mean hydrophobicity (H) and mean relative hydrophobic moments (μH) were calculat

HydroMCalc.html).c Hemolysis andd cytotoxicity were assayed against hRBCs and HaCaT cells, respectively, at a peptide con

to 0.00119 μM and from 88.3 to 0.043 μM, respectively. Daily MICswere determined for 15 days for each agent using cells from thewell containing one-half the MIC (1/2 MIC well). In brief, cells fromthe 1/2 MIC well were suspended in appropriate culture media andincubated overnight at 37 °C, after which the suspensions were ad-justed to 5×105 CFU/ml in DMEM supplemented with 10% FBS andmixed with agents in the same concentration ranges. All MICs weredetermined in duplicate [22].

3. Results

3.1. Design and antimicrobial/cytotoxic activities of HPA3NT3 analogs

HPA3NT3 has an amphipathic helical structure, as indicated by thehelical wheel diagram in Fig. 1. The Phe residues at positions 1 and8 are situated at the hydrophobic face and are crowding the Trp res-idue at position 12. From this starting point, we designed analog pep-tides using three approaches: (1) reduce the hydrophobicity byreplacing Phe with Ala, (2) increase the cationicity by replacing Asnwith Lys, and (3) create completely amphipathic divisions in thehelical wheel diagram by replacing Asn with Lys in position 13 andLeu with Lys in position 15. The sequences of the peptides areshown in Table 1. We then investigated the antimicrobial activitiesof the peptides against eight pathogenic bacteria and six fungalstrains (Table 2). To evaluate cell selectivity, we determined the he-molytic and cytotoxic activities of the peptides against hRBCs and

s analogs.

Retentiontimea

Hb Hemolysis(%)c

Cytotoxicity(%)d

20.94 −0.60 69.2 85.119.45 −1.34 7.1 68.317.93 −1.34 4.6 60.116.25 −2.08 0 7.819.87 −0.96 20.9 78.715.20 −2.26 0 11.7

ed to CSS scale using HydroMCalc (http://www.bbcm.univ.trieste.it/~tossi/HydroCalc/

centration of 250 μM.

Page 5: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Table 2Antimicrobial activity of HPA3NT3 and its analogs against various pathogens.

Microorganism MICs (μM)

1⁎ 2⁎ 3⁎ 4⁎ 5⁎ 6⁎

Gram (−) bacteriaE. coli 4 4 4 32 32 4P. aeruginosa 2 1 2 2 32 2P. vulgaris 8 4 4 4 8–16 2–4S. typhimurium 2 1 2 2 8 1–2

Gram (+) bacteriaS. aureus 2 2 4 16 16 4L. monocytogenes 2 2 2 2 16 2S. epidermidis 2–4 2–4 2–4 16 8 2–4B. subtilis 2 2 4 4 8 2

YeastC. albicans 8 8 8 16 16 8T. beigelii 16 16 16 32 >128 16

FungiA. awamori 16 16 16 16 64 8A. flavus 16 16 16 32 64 8A. fumigatus 4 8 8 4 16 4A. parasiticus 8 8 8 8 16 8

⁎ 1, 2, 3, 4, 5 and 6 indicate HPA3NT3, -F1A, -F8A, -F1AF8A, -A1 and -A2, respectively.All assayswere performed in a 10 mM sodiumphosphate (pH 7.2) containing 10% culturemedia.

447J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

HaCaT cells, respectively (Table 1). The results showed that the anti-microbial activities of the single substitutes (‐F1A and -F8A) wereclose to or slightly greater than that of HPA3NT3, and their cytotoxicactivities were somewhat lower. The -F1AF8A double substitute wasconsiderably less hemolytic than the single substitutes, and its MICagainst the various microbes was similar to or less than that ofHPA3NT3 (Table 2). In contrast, -A2 showed greater antimicrobial ac-tivity than -F1AF8A or HPA3NT3 against some of the microbes tested

Fig. 2.Membrane binding and dye release with artificial zwitterionic vesicles. Emission maxiC) and peptide induced calcein leakage from LUVs (B and D). Data in A and B were obtaine

(Table 2). Furthermore, -A2 exhibited no hemolytic activity, even at250 μM (Table 1), nor did it show significant cytotoxicity againstHaCaT cells (Table 1). Of the several analogs synthesized, only themechanism of action of -A2 was studied in detail, since it had thehighest antimicrobial activity with the least cytotoxicity.

3.2. Membrane binding and dye release with artificialzwitterionic vesicles

To determine how the hemolytic and cytotoxic activities of HPA3NT3were ameliorated in ‐F1AF8A, binding and permeabilization activity wasassessed based on Trp fluorescence and calcein leakage using artificial,PC/CH (2:1, w/w) and PC/SM (2:1, w/w) LUVS. PC, CH and SM arecommon lipids found in the mammalian erythrocyte membrane.HPA3NT3, ‐F1A, -F1AF8A and -A1 did not bind to PC/CH vesicles, andthere was no significant calcein leakage from those vesicles (Fig. 2Aand B). In contrast, HPA3NT3 strongly bound to PC/SM vesicles, resultingin a shift in the Trp fluorescence maximum from 351.5 to 339 nm in thepresence of 300 μM vesicles (peptide/lipid molar ratio, 1:150) (Fig. 2C).Moreover, HPA3NT3 induced a corresponding loss of calcein from thePC/SM vesicles (Fig. 2D) -F1A and -F1AF8A did not bind to the PC/SMLUVs and thus did not induce calcein leakage. However, Fig. 2C and Dshows that treatment with -A1 shifted the Trp fluorescence maximumand caused a 40% loss of calcein from PC/SM vesicles, a result reflectingits greater hydrophobicity. In 10 mM HEPES buffer (pH 7.2), HPA3NT3,-F1A and -F1AF8A exhibited fluorescence emission maxima at 351 nm,which corroborates the earlier finding that Trp residues in these peptidesare located in a hydrophilic environment. However, when lipid vesicleswere present, as the Trp residues in ‐A1 moved to a less polar environ-ment the emission maximum shifted to 349 nm. This observationcould explain the apparent cytotoxicity of -A1, but we do not have the

ma from Trp in peptides in the presence of the indicated concentrations or LUVs (A andd with PC/CH (2:1, w/w) vesicles, those in C and D with PC/SM (2:1, w/w) vesicles.

Page 6: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 3. Blue shift in Trp fluorescence and calcein release from PE:PG liposomes. Emission maxima from Trp in peptides in the presence of PE/PG LUVs (A) and calcein leakage fromthe vesicles (B).

448 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

necessary orientation or structure details available for the present studyto provide a conclusive determination.

3.3. Blue shift in Trp fluorescence and calcein release fromPE:PG liposomes

The actions of the peptides towards PE:PG (7:3, w/w) LUVs, whichmimic a bacterial membrane, were determined using Trp blue shiftassays and calcein leakage assays. The degree to which the peptidesinteracted with the membrane (1:25 peptide-to-lipid ratio) was inthe order, HPA3NT3>-A2>-F1AF8A>-A1 (Fig. 3A). The largest blueshift was seen with HPA3NT3, which suggests the Trp side chainpartitioned preferentially into a more rigid, hydrophobic environ-ment in the PE:PG LUVs [23]. Treating PE:PG LUVs with HPA3NT3 ata 1:10 P/L molar ratio resulted in a maximum leakage of 100%,

Fig. 4. Confocal laser scanning micrograph images of E. coli cells treated with TAMRA-TAMRA-labeled peptide for 10 min at 37 °C in PBS (pH 7.2). From left to right: TAMRA, dif

whereas -A2 released a maximum of approximately 32% of theentrapped calcein (Fig. 3B).

3.4. Intermembrane and intracellular localization of the peptides

The morphology of E. coli cells incubated in the presence of the pep-tides was examined using light microscopy during the antimicrobialassay. Whereas untreated cells tended to gather together (data notshown), cells treated with HPA3NT3 were scattered (data not shown).HPA3NT3 reduced the bacterial population at the 1/2-MIC, yet thecells retained a shape similar to normal E. coli. Interestingly, in the pres-ence of -A2 the cells appeared to be elongated in shape (data notshown), a result suggesting that cell divisionwas inhibited. The locationof fluorescently labeled peptide in both of the treated E. coli cultureswas monitored using confocal laser scanning microscopy. TAMRA-HPA3NT3 (at itsMIC) accumulated to a greater degree at the cell surface

labeled- TAMRA-HPA3NT3 (a) or TAMRA-A2 (b). The cells were treated with MICferential interference contrast (DIC), merged images.

Page 7: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 5. Kill kinetics and membrane permeation. Time-kill kinetic (A), time-dependent depolarization (B) and influx of SYTOX green (C and D) in E. coli. A, E. coli cells were exposedto peptides at 1x or 2x MIC, and incubated for 0–100 min. Data were recorded at 5-min intervals. B, E. coli cells in mid-logarithmic phase were adjusted to OD600=0.05and pre-equilibrated for 60 min with diS-C35. HPA3NT3 (solid line) or -A2 (dotted line) was then added to 4 μM, and changes in fluorescence were monitored continuously(Ex. 622 nm and Em. 670 nm) and plotted as arbitrary units. The second arrow indicates the addition of Triton X-100 (0.1%) to confirm the maximum depolarization. C and D,E. coli (2×107 cells/ml) were incubated with 1 μM SYTOX green for 15 min. HPA3NT3 (C) or -A2 (D) were added when the basal fluorescence reached a steady value, and the in-crease of fluorescence was recorded at the indicated times (Ex. 485 nm and Em. 520 nm).

449J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

than in the cytoplasm, while the reverse was true for TAMRA-A2(Fig. 4). This suggests that -A2 was translocated to the cytoplasmwhere it potentially interacted with DNA, RNA and related proteins.

3.5. Kill kinetics and membrane permeation

Fig. 5A shows time-kill curves for HPA3NT3 and -A2 against E. coli at1x and 2x MIC. The kill kinetics for HPA3NT3 were faster than for -A2;moreover, HPA3NT3 killed 100% of the bacterial cells while -A2 achievedonly partial elimination of the bacteria. To evaluate peptide-inducedmembrane permeabilization, we assessed the ability of the peptides toinduce membrane depolarization in E. coli. HPA3NT3 maximallydepolarized E. coli cells at 4 μM but -A2 elicited only half the maximumat the same concentration, even though both peptides had the sameMIC (Fig. 5B). This result suggests that these peptides interact with theE. coli membrane differently, which in turn suggests the mechanismsthat their antimicrobial actions also likely differ. We monitored the pep-tides' permeabilization of E. colimembranes by measuring the uptake ofthe normally impermeant nucleic acid binding molecule, SYTOX green(Fig. 5C and D). The results indicated that HPA3NT3 induces greater up-take of the dye than does -A2. The rate of permeabilization was partiallyconcentration-dependent and increased with increasing concentrationsof HPA3NT3 up to 8 μM. At concentrations above 8 μM there was nochange in the kinetics of SYTOX green uptake, a result suggesting thatpermeabilization by HPA3NT3 is saturable. The maximum uptake of

-A2 occurred at 1/2-MIC (2 μM) and there was no change in the perme-abilization at higher concentrations (4, 8 and 16 μM).

3.6. Peptide binding to DNA and RNA

We next examined peptide binding to plasmid DNA and yeastRNA. The peptides were mixed with a fixed amount of plasmid(pTYB2) DNA (pDNA) (Fig. 6A) or yeast RNA (Fig. 6B), after whichthe samples were electrophoresed on agarose gels. The -A2 complete-ly retarded pDNA migration at a peptide/pDNA weight ratio of 0.5(Fig. 6A, bottom of the right panel). When treated with indolicidin,which is known to bind to pDNA at a ratio of 1 (Fig. 6A, top of theright panel) [24], complete retardation of the DNA migration was ob-served. In contrast, magainin-II did not affect the pDNA migration,even at a peptide/pDNA weight ratio of 4, indicating that magainin-IIhad no interaction with the pDNA (Fig. 6A, top of the left panel). In ad-dition, Fig. 6B shows that the binding of -A2 to yeast RNA was greaterthan that observed with buforin 2. RNA binding was observed, even atlow peptide concentrations, as indicated by the slight downward shiftin the RNA band. In order to confirm the -A2 binding to pTYB2, the pep-tide and the plasmid were mixed together and then examined usingtransmission electron microscopy. Fig. 6C shows that the pTYB2 DNAdid not aggregate in the absence of peptide (Fig. 6C, left panel). Onthe other hand, the pTYB2 DNA clumped together in the presence of

Page 8: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 6. Peptide binding to DNA and RNA. Gel retardation analysis of peptides binding to DNA (A) and RNA (B) and negative staining of the DNA-peptide complex (C). A and B, Bind-ing was assayed based on the inhibitory effect of the peptides on the migration of DNA or RNA. The indicated amounts of peptide were mixed with 200 ng of plasmid DNA (pTYB 2,A) or 10 μg of yeast RNA (B) and incubated for 10 min at 37 °C. The reaction mixtures were then applied to a 1% (A) or 0.5% (B) agarose electrophoresis gel. C, pTYB2 DNA in theabsence or presence of -A2 was negatively stained with 1% UAC and visualized using transmission electron microscopy. Values at the top are peptide to DNA or RNA ratios.

450 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

the two concentrations of ‐A2, confirming the interaction between ‐A2and pTYB2 DNA.

3.7. Inhibition of recombinant GFP expression in E. coli

To investigate the effect of intracellular -A2 on GFP expression inE. coli cells, expressed GFP was visualized using either fluorescent mi-croscopy or 15% SDS-PAGE. Fig. 7A shows that untreated cells pro-duced strong fluorescence, but that the fluorescence declined whenthe cells were exposed to increasing concentrations of -A2. In addi-tion, when we used SDS‐PAGE to assess GFP expression in E. coli inthe presence and absence of the peptide, we found that GFP expres-sion gradually declined with increasing concentrations of -A2 until,at 1x MIC, the GFP band was completely absent (Fig. 7B). This resultsuggested that -A2 enters E. coli cells, binds to nucleic acids, and in-hibits protein synthesis.

3.8. -A2 does not induce resistance, in vitro

The development of resistance to -A2, and to rifampin and vancomy-cin (positive controls), was evaluated using S. aureus (ATCC 25923,antibiotic-susceptible strain) after its growth in the presence of the re-spective antibiotics. TheMIC for rifampin, which is known to quickly in-duce resistance through chromosomal pointmutation [25], significantly

increased 8192-fold over the course of 12 passages. In contrast, theMICfor vancomycin increased only 4-fold, and there was no change in theMIC for -A2. Thus, unlike rifampin, exposure to -A2 did not lead to thedevelopment of resistance in S. aureus (Fig. 8). We also assessed thebactericidal activity of -A2 against various other antibiotic- andmultidrug‐resistant strains. The results are summarized in Table 3.

4. Discussion

Because AMPs mainly interact with lipid membranes through elec-trostatic and hydrophobic interactions, the structure and function ofthe transmembrane domain of membrane proteins have been used tocharacterize and evaluate the mechanisms of action of antimicrobialpeptides atmembrane-water interfaces and in lipid bilayers. The explo-ration of these mechanisms has led to the development of models in-cluding the barrel-stave model [26,27], toroidal pore formation [28],detergent-like model [29] and the translocation model [30]. Studies oftransmembrane proteins/AMPs have suggested that the indole sidechain of Trp is ideally suited for interacting with the polar–nonpolar in-terface, with the highly hydrophobic aromatic ring preferentially buriedin the hydrophobic region of the lipid bilayer [31]. Hydrophobic Phe res-idues also contain an aromatic ring, which mainly situates in the trans-membrane portion of membrane proteins or in the hydrophobic regionof the lipid bilayer [32]. This clearly suggests that Phe and Trp have

Page 9: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 7. Inhibition of recombinant GFP expression in E. coli. Inhibition of GFP expression by -A2. A, After incubation for 30 min in the presence of the indicated concentration of -A2,E. coli BL21(DE3) cells transfected with pET28(a)-GFP were exposed to 0.5 mM IPTG for 2 h to induce GFP expression: a, phase-contrast images of the cells; b, GFP fluorescence fromthe same cells. B, SDS-PAGE analysis of total soluble protein collected from GFP-expressing cells incubated in the absence or presence of peptide.

451J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

strongly hydrophobic interactions with zwitterionic liposomes that re-semblemammalianmembranes. In addition, positively charged Lys andArg residues have relatively long aliphatic side chains that prefer thepolar region of the negatively charged phosphate groups in microbialmembranes [31,33,34]. Thus, amino acid side chains play a key role inthe antimicrobial and cytotoxic activities of AMPs, as well as determin-ing their selectivity formicrobial andmammalianmembranes [35]. Thisfinding suggests that synthetic AMPs could be specifically designed inorder to interact with specific targets of the microbial cell wall (sur-rounding the cytoplasmic membrane), based on their amino acid com-bination, charge and hydrophobicity/hydrophilicity [36,37].

With this in mind, we designed several analogs from the parentpeptide, HPA3NT3, which is a truncated analog of HP(2‐20) derivedfrom the N-terminus of the Helicobactor pylori ribosomal protein L1[7]. Through a pore-forming mechanism, HPA3NT3 exerts potent an-timicrobial effects against an even broader spectrum of microbes thandoes HP(2‐20) [11]. Unfortunately, HPA3NT3 cannot be utilized invivo due to its cytolytic activity at efficacious dose concentrations.We suggest that the binding of HPA3NT3 to the plasma membraneis mediated by the hydrophilic face, and that the inserts into the hy-drophobic face of the bilayer due to the presence of two Trp residuesat that end. If this is so, how does HPA3NT3 recognize the mammalian

plasma membrane? We propose that recognition is via the aromaticrings of Phe and Trp, which are packed on the hydrophobic face ofthe amphipathic helical structure (Fig. 1) and are known to interactwith CH and SM in order to exert its hemolytic and cytolytic effects[7]. In addition, the strong hydrophobicity of HPA3NT3 may enableits insertion into the lipid interior of the mammalian membranewhich is mainly composed of zwitterionic lipids.

In our analysis, Trp was never substituted because it is conservedin all natural AMPs and is essential for their antibacterial activity; itsindole side chain strongly interacts with bacterial phospholipid mem-branes, thereby promoting toxicity [38], so that replacing Trp withanother amino acid diminishes an AMP's antimicrobial activity [39].Instead, HPA3NT3 analogs were produced by substituting one orboth of the Phe residues (position 1 and/or 8) with an Ala residueto reduce the peptide's hydrophobicity. As shown in Table 1, the re-tention times and hydrophobicity of -F1A, -F8A and -F1AF8A werelower than those of HPA3NT3. The hydrophobicity (H), net chargeand hydrophobicmoment are three important physicochemical param-eters useful when evaluating membrane-interfacial and pore-formingAMPs andmembrane proteins [11,39]. The bacterial plasmamembraneconsists mainly of zwitterionic PE combined with a large amount of an-ionic lipids, such as PG and cardiolipin, whereas the membrane of

Page 10: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 8. -A2 does not induce resistance in vitro. Induction of resistance to antimicrobialpeptides and conventional antibiotics. The Y-axis indicates the relative change in theMIC from the first passage.

Table 3Antibacterial activity of HPA3NT3 and -A2 against drug-resistant strains.

Strains MIC (μM)

HPA3NT3 -A2 Erya Ampa Cipa

E. coli CCARM 1229§ 2 1 – >512 512E. coli CCARM 1238§ 2 2 – >512 512P. aeruginosa 3547¶ 4 2 – – 256P. aeruginosa 4007¶ 1 1 – – 256S. aureus CCARM 3089§ 2 2 >512 – –

S. aureus CCARM 3114§ 4 4 >512 – –

S. aureus PBEL 1⁎ 1 1 – – –

S. aureus PBEL 2⁎ 1 1 – – –

S. typhimurium CCARM 8009§ 4 1 – >512 –

S. typhimurium CCARM 8013§ 1 1 – >512 –

§ These strains (with CCARM numbers) were obtained from the Culture Collection ofAntimicrobial Resistant Microbes, in Korea.

¶ These resistant strains were isolated from patients in a hospital using otitis media.⁎ S. aureus PBEL 1 and 2 were, respectively, rifampin-resistant and

vancomycin-resistant strains induced in this study. All peptides were assayed in DMEMsupplemented with 10% FBS. Ery, Amp and Cip are erythromycin, ampicillin and cipro-floxacin, respectively.

452 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

healthy human cells appears to be uncharged in the outer leaflet whichconsists almost exclusively of the zwitterionic phospholipids PC andsphingomyelin, as well as cholesterol. Substituting Ala for Phe in -F1A,-F8A and -F1AF8A clearly diminished their cytotoxicity in associationwith reduced hydrophobicity and altered hydrophobic moment. In par-ticular, Phe residues have a strong affinity for SM, and the analog pep-tides showed less binding to and permeabilization of PC/SM vesicles.This probably reflects the absence of the Phe aromatic rings, which areclosely packed at the same face of HPA3NT3 (Fig. 1), and are requiredfor insertion of the peptide into SM-containing membranes. On theother hand, none of the tested peptides effectively bound to PC/CH ves-icles.Moreover, our results indicate that the hemolytic and cytotoxic ac-tivities of these peptides were directly related to the SM content in theouter leaflets of hRBCs and HaCaT cells.

The amino acids were substituted while maintaining the amphi-pathic organization of the HPA3NT3 peptide (Fig. 1). Augmentationof the hydrophobicity of the peptides through substitution of Lysand Leu for Asn and Lys at positions 13 and 15 of -F1AF8A (‐A1) dra-matically increased hemolysis and cytotoxicity. This suggests that thehydrophobicity is a key determinant of the hemolytic and cytotoxicproperties of an AMP [7] (Table 1).

Another peptide analog was designed to increase the net positivecharge by substituting Lys for Leu at position 15 without changing theLys residue at position 13 (‐A2). The Lys residue at position 13 of -A1was left unchanged because it was previously shown that positivecharge substitution did not affect the α‐helical structure or cytotoxicactivity [40], but the hydrophobic Leu residue at position 15 of -A1had critical effects on its cytotoxicity [7] (Table 1, Fig. 2) andantibacterial activity [41] (Tables 2 and 3, Fig. 3).

AMPs generally kill bacteria through permeabilization of the cyto-plasmic membrane or by binding to intracellular targets, such asnucleic acids [1]. It is reasonable to assume that AMPs containingHPA3NT3-like motifs interact with and damage the cytoplasmicmembrane, and do not bind to nucleic acids [11]. However a substitu-tion at position 13 appears to change the mechanism of action(Fig. 3). The bactericidal activity of -A2 was confirmed using

E. coli-based assays designed to evaluate its killing kinetics, whichshowed that cell viability reduced significantly within 40 min oftreatment. In addition, HPA3NT3 and -A2 had the same MIC asE. coli, but the killing kinetics of -A2 was slower than those forHPA3NT3, a finding which suggested that a different mechanism con-tributes to -A2 killing action (Fig. 5A). The antimicrobial action ofsome AMPs involved their insertion into the cytoplasmic membrane,pore formation, and resultant dissipation of the membrane potentialand disruption of lipid asymmetry [26–30]. We examined the interac-tion of -A2 with the membrane using bacterial cells. We found that-A2 completely dissipated the membrane potential of E. coli at itsMIC, whereas HPA3NT3 induced minimal depolarization (Fig. 5B).The -A2 also induced increasing efflux of SYTOX green (Fig. 5C andD), concentration-dependently, indicating that it effected both theouter and inner membrane. This notion was supported by the findingthat TAMRA-A2 localized mainly in the membrane or cytoplasm ofE. coli, depending upon its concentration. This result suggested that-A2's bactericidal activity reflects both an action at the membraneand an intracellular action in the cytoplasm (Fig. 4).

Interestingly, in this study we found that -A2 peptide entering thecytoplasm through self-formed pores specifically bound to DNA(Fig. 6A and C) or RNA (Fig. 6B), a finding which is expected to inhibitprotein synthesis (Fig. 7B). This hypothesis of its mechanism of actionexplains its favorable killing kinetics and the increased efflux ofSYTOX green over the MIC value.

It is noteworthy that the mechanism by which -A2 penetratedthe bacterial cell membrane relied on the substitution of the aminoacids at positions 1, 8 and 13 of HPA3NT3. We suggest that themechanism by which -A2 binds to nucleic acids involves its stronginteraction with phosphodiester bonds via its positively chargedLys, with the Trp residue stacking between the nucleotide base or ri-bose in each strand of the nucleic acid duplex. Several AMPs areknown to exhibit nucleic acid binding properties [40,42]; these in-clude cationic and hydrophobic-rich antimicrobial peptides, such asindolicidin and LfcinB, as well as the scramble peptide PAF26[24,43]. It can be reasonably expected that -A2 kills bacterial cellsby targeting DNA, RNA or proteins in order to inhibit synthesisof nucleic acids and proteins. Notably, in contrast to some conven-tional antibiotics, S. aureus remained susceptible to -A2, even afterprolonged exposure (Fig. 8).

The modes of action of HPA3NT3 and -A2 are summarized in Fig. 9.HPA3NT3 exerts its bactericidal effects by intercalating into the mem-brane, most likely through interaction between its Phe residues andSM, and by inducing pore formation. In contrast, -A2 acts by penetratinginto the cytoplasm, binding to nucleic acids, and thus inhibiting protein

Page 11: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

Fig. 9. The modes of action of HPA3NT3 and -A2. Schematic representation of the proposed mechanisms of action of HPA3NT3 and -A2 in bacterial cells.

453J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

synthesis. Notably, S. aureus did not develop resistance to -A2 as it doeswith rifampin. Collectively, these results suggest that -A2 could serve asan effective antibiotic agent against multidrug-resistant bacterialstrains.

Acknowledgments

This work was supported by a National Research Foundationof Korea (NRF) grant funded by the Korea government (MEST)(No. 2011‐0017532).

References

[1] K.A. Brogden, Antimicrobial peptides: pore formers or metabolic inhibitors inbacteria? Nature 3 (2005) 238–250.

[2] M. Zasloff, Antimicrobial peptides of multicellular organisms, Nature 415 (2002)389–395.

[3] R.E.W. Hancock, R. Lehrer, Cationic peptides: a new source of antibiotics, TrendsBiotechnol. 16 (1998) 82–88.

[4] L.A. Allen, Modulating phagocyte activation: the pros and cons of Helicobacterpylori virulence factors, J. Exp. Med. 191 (2000) 1451–1454.

[5] K. Pütsep, C.I. Bränden, H.G. Boman, S. Normark, Antibacterial peptide fromH. pylori, Nature 398 (1999) 671–672.

[6] J. Bylund, T. Christophe, F. Boulay, T. Nyström, A. Karlsson, C. Dahlgren,Proinflammatory activity of a cecropin-like antibacterial peptide from Helicobacterpylori, Antimicrob. Agents Chemother. 45 (2001) 1700–1704.

[7] S.C. Park, M.H. Kim, M.A. Hossain, S.Y. Shin, Y. Kim, L. Stella, J.D. Wade, Y. Park, K.S.Hahm, Amphipathic alpha-helical peptide, HP (2‐20), and its analogues derivedfrom Helicobacter pylori: pore formation mechanism in various lipid composi-tions, Biochim. Biophys. Acta Biomembr. 1778 (2008) 229–241.

[8] J.K. Ghosh, D. Shaool, P. Guillaud, L. Cicéron, D.Mazier, I. Kustanovich, Y. Shai, A.Mor,Selective cytotoxicity of dermaseptin S3 toward intraerythrocytic Plasmodiumfalciparum and the underlying molecular basis, J. Biol. Chem. 272 (1997)31609–31616.

[9] D. Rapaport, Y. Shai, Aggregation and organization of pardaxin in phospholipidmembranes, J. Biol. Chem. 267 (1992) 6502–6509.

[10] S.C. Park, J.R. Lee, S.O. Shin, Y. Park, S.Y. Lee, K.S. Hahm, Characterization of aheat-stable protein with antimicrobial activity from Arabidopsis thaliana,Biochem. Biophys. Res. Commun. 362 (2007) 562–567.

[11] Y. Park, S.C. Park, H.K. Park, S.Y. Shin, Y. Kim, K.S. Hahm, Structure-activity rela-tionship of HP (2‐20) analog peptide: enhanced antimicrobial activity byN-terminal random coil region deletion, Biopolymers 88 (2007) 199–207.

[12] Y.H. Chen, J.T. Yang, K.H. Chau, Determination of the helix and beta form of proteins inaqueous solution by circular dichroism, Biochemistry 13 (1974) 3350–3359.

[13] K. Marynka, S. Rotem, I. Portnaya, U. Cogan, A. Mor, In vitro discriminativeantipseudomonal properties resulting from acyl substitution of N-terminal se-quence of dermaseptin S4 derivatives, Chem. Biol. 14 (2007) 75–85.

[14] M.L. Mangoni, N. Papo, D. Barra, M. Simmaco, A. Bozzi, A. Di Giulio, A.C.Rinaldi, Effects of the antimicrobial peptide temporin L on cell morphology,membrane permeability and viability of Escherichia coli, Biochem. J. 380(2004) 859–865.

[15] S.C. Park, J.R. Lee, S.O. Shin, J.H. Jung, Y.M. Lee, H. Son, Y. Park, S.Y. Lee, K.S. Hahm,Purification and characterization of an antifungal protein, C-FKBP, from Chinesecabbage, J. Agric. Food Chem. 55 (2007) 5277–5281.

[16] D.G. Lee, H.N. Kim, Y. Park, H.K. Kim, B.H. Choi, C.H. Choi, K.S. Hahm, Design ofnovel analogue peptides with potent antibiotic activity based on the antimicrobi-al peptide, HP (2‐20), derived from N-terminus of Helicobacter pylori ribosomalprotein L1, Biochim. Biophys. Acta 1598 (2002) 185–194.

Page 12: Biochimica et Biophysica Acta - COnnecting … this peptide, its release gives the bacterium a competitive advan-tage over other microorganisms [5]. The linear antimicrobial peptide

454 J.-K. Lee et al. / Biochimica et Biophysica Acta 1828 (2013) 443–454

[17] L.D. Mayer, M.J. Hope, P.R. Cullis, Vesicles of variable sizes produced by a rapid ex-trusion procedure, Biochim. Biophys. Acta 858 (1986) 161–168.

[18] J.C.M. Stewart, Colorimetric determination of phospholipids with ammoniumferrothiocyanate, Anal. Biochem. 104 (1980) 10–14.

[19] J.K. Ghosh, S.G. Peisajovich, M. Ovadia, Y. Shai, Serum-induced leakage of lipo-some contents, J. Biol. Chem. 273 (1998) 27182–27190.

[20] C.H. Hsu, C. Chen, M.L. Jou, A.Y. Lee, Y.C. Lin, Y.P. Yu, W.T. Huang, S.H. Wu, Struc-tural and DNA-binding studies on the bovine antimicrobial peptide, indolicidin:evidence for multiple conformations involved in binding to membranes andDNA, Nucleic Acids Res. 33 (2005) 4053–4064.

[21] J.R. Lee, S.S. Lee, H.H. Jang, Y.M. Lee, J.H. Park, S.C. Park, J.C. Moon, S.K. Park, S.Y.Kim, S.Y. Lee, H.B. Chae, Y.J. Jung, W.Y. Kim, M.R. Shin, G.W. Cheong, M.G. Kim,K.R. Kang, K.O. Lee, D.J. Yun, S.Y. Lee, Heat-shock dependent oligomeric status al-ters the function of a plant-specific thioredoxin-like protein, AtTDX, Proc. Natl.Acad. Sci. U. S. A. 106 (14) (2009) 5978–5983.

[22] V. Dartois, J. Sanchez-Quesada, E. Cabezas, E. Chi, C. Dubbelde, C. Dunn, J. Granja,C. Gritzen, D. Weinberger, M.R. Ghadiri, T.R. Parr Jr., Systemic antibacterial activ-ity of novel synthetic cyclic peptides, Antimicrob. Agents Chemother. 49 (2005)3302–3310.

[23] E.A. Burstein, N.S. Vedenkina, M.N. Ivkova, Fluorescence and the location of tryp-tophan residues in protein molecules, Photochem. Photobiol. 18 (4) (1973)263–279.

[24] C. Subbalakshmi, N. Sitaram, Mechanism of antimicrobial action of indolicidin,FEMS Microbiol. Lett. 160 (1998) 91–96.

[25] W. Zheng, Z. Li, K. Skarstad, E. Crooke, Mutations in DnaA protein suppress thegrowth arrest of acidic phospholipid-deficient Escherichia coli cells, EMBO J. 20(5) (2001) 1164–1172.

[26] L. Yang, T.A. Harroun, T.M. Weiss, L. Ding, H.W. Huang, Barrel-stave model or to-roidal model? A case study on melittin pores, Biophys. J. 81 (2001) 1475–1485.

[27] R.S. Cantor, Size distribution of barrel-stave aggregates of membrane peptides:influence of the bilayer lateral pressure profile, Biophys. J. 82 (2002) 2520–2525.

[28] E. Gazit, W.J. Lee, P.T. Brey, Y. Shai, Mode of action of the antibacterial cecropinB2: a spectrofluorometric study, Biochemistry 33 (1994) 10681–10692.

[29] A.S. Ladokhin, M.E. Selsted, S.H. White, Sizing membrane pores in lipid vesicles byleakage of co-encapsulated markers: pore formation by melittin, Biophys. J. 72(1997) 1762–1766.

[30] K. Takeshima, A. Chikushi, K.K. Lee, S. Yonehara, K. Matsuzaki, Translocation ofanalogues of the antimicrobial peptides magainin and buforin across humancell membranes, J. Biol. Chem. 278 (2) (2003) 1310–1315.

[31] J.A. Killian, G. von Heijne, How proteins adapt to a membrane-water interface,Trends Biochem. Sci. 25 (2000) 429–434.

[32] E. Wallin, T. Tsukihara, S. Yoshikawa, G. von Heijne, A. Elofsson, Architecture ofhelix bundle membrane proteins: an analysis of cytochrome c oxidase from bo-vine mitochondria, Protein Sci. 6 (1997) 808–815.

[33] J.P. Segrest, H. De Loof, J.G. Dohlman, C.G. Brouillette, G.M. Anantharamaiah, Am-phipathic helix motif: classes and properties, Proteins 8 (1990) 103–117.

[34] M.R. de Planque, J.A. Kruijtzer, R.M. Liskamp, D. Marsh, D.V. Greathouse, R.E.Koeppe II, B. de Kruijff, J.A. Killian, Different membrane anchoring positions oftryptophan and lysine in synthetic transmembrane alpha-helical peptides, J.Biol. Chem. 274 (1999) 20839–20846.

[35] Ø. Rekdal, B.E. Haug, M. Kalaaji, H.N. Hunter, I. Lindin, I. Israelsson, T. Solstad, N.Yang, M. Brandl, D. Mantzilas, H.J. Vogel, Relative spatial positions of tryptophanand cationic residues in helical membrane-active peptides determine their cyto-toxicity, J. Biol. Chem. 287 (2012) 233–244.

[36] M. Dathe, T. Wieprecht, Structural features of helical antimicrobial peptides: theirpotential to modulate activity on model membranes and biological cells, Biochim.Biophys. Acta 1462 (1999) 71–87.

[37] A. Giangaspero, L. Sandri, A. Tossi, Amphipathic alpha helical antimicrobial pep-tides, Eur. J. Biochem. 268 (2001) 5589–5600.

[38] D.J. Schibli, R.F. Epand, H.J. Vogel, R.M. Epand, Tryptophan-rich antimicrobial pep-tides: comparative properties and membrane interactions, Biochem. Cell Biol. 80(2002) 667–677.

[39] R. Gopal, Y.J. Kim, C.H. Seo, K.S. Hahm, Y. Park, Reversed sequence enhances anti-microbial activity of a synthetic peptide, J. Pept. Sci. 17 (5) (2011) 329–334.

[40] C.B. Park, H.S. Kim, S.C. Kim, Mechanism of action of the antimicrobial peptidebuforin II: buforin II kills microorganisms by penetrating the cell membraneand inhibiting cellular functions, Biochem. Biophys. Res. Commun. 244 (1998)253–257.

[41] Y. Park, D.G. Lee, S.H. Jang, E.R. Woo, H.G. Jeong, C.H. Choi, K.S. Hahm, ALeu-Lys-rich antimicrobial peptide: activity and mechanism, Biochim. Biophys.Acta 1645 (2) (2003) 172–182.

[42] A. Yonezawa, J. Kuwahara, N. Fujii, Y. Sugiura, Binding of tachyplesin I to DNA re-vealed by footprinting analysis: significant contribution of secondary structure toDNA binding and implication for biological action, Biochemistry 31 (1992)2998–3004.

[43] H. Ulvatne, O. Samuelsen, H.H. Haukland, M. Kramer, L.H. Vorland, Lactoferricin Binhibits bacterial macromolecular synthesis in Escherichia coli and Bacillus subtilis,FEMS Microbiol. Lett. 237 (2004) 377–384.