Antibacterial Effects of a Cell-Penetrating Peptide Isolated from Kefir

Mar 22, 2016 - E-mail: [email protected]., *(Q.H.) Phone: (848) 932-5514. Fax: (732) 932-6776. E-mail: ... Kefir is a traditional fermented milk...
0 downloads 4 Views 4MB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

Antibacterial effects of a cell-penetrating peptide isolated from kefir Jianyin Miao, Haoxian Guo, Feilong Chen, Lichao Zhao, Liping He, Yangwen Ou, Manman Huang, Yi Zhang, Baoyan Guo, Yong Cao, and Qingrong Huang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b00730 • Publication Date (Web): 22 Mar 2016 Downloaded from http://pubs.acs.org on March 25, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32

Journal of Agricultural and Food Chemistry

1

Antibacterial effects of a cell-penetrating peptide isolated from kefir

2

Jianyin Miao†,‡,§,#, Haoxian Guo†,§, Feilong Chen†,‡, Lichao Zhao†,‡, Liping He†,‡, Yangwen

3

Ou#, Manman Huang†, Yi Zhang†, Baoyan Guo†, Yong Cao†,‡, *, Qingrong Huang¶,*

4 5



6

Republic of China

7



8

Guangzhou 510642, People’s Republic of China

9

§

10



11

08901, USA

12

#

13

410000, People’s Republic of China

College of Food Science, South China Agricultural University, Guangzhou 510642, People’s

Guangdong Provence Engineering Research Center for Bioactive Natural Products,

Qinzhou University, Qinzhou 535000, People’s Republic of China Department of Food Science, Rutgers University, 65 Dudley Road, New Brunswick, NJ

School of Pharmacy, Hunan University of Chinese Medicine, Changsha

14

15

Corresponding Authors

16

* Telephone: +86-20-85286234. Fax: +86-20-85286234. E-mail: [email protected].

17

* Telephone: 848-932-5514. Fax: 732-932-6776. E-mail: [email protected].

18 19

20

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

21

ABSTRACT: Kefir is a traditional fermented milk beverage used throughout the

22

world for centuries. A cell-penetrating peptide F3 was isolated from kefir by

23

Sephadex G-50 gel filtration, DEAE-52 ion exchange and reverse-phase high

24

performance liquid chromatography. F3 was determined to be a low molecular weight

25

peptide containing one leucine and one tyrosine with two phosphate radicals. This

26

peptide displayed antimicrobial activity across a broad spectrum of organisms

27

including several Gram-positive and Gram-negative bacteria as well as fungi, with

28

minimal inhibitory concentration (MIC) values ranging from 125 to 500 µg/mL.

29

Cellular penetration and accumulation of F3 were determined by confocal laser

30

scanning microscopy. The peptide was able to penetrate the cellular membrane of

31

Escherichia coli and Staphylococcus aureus. Changes in cell morphology were

32

observed by scanning electron microscopy (SEM). The results indicate that peptide F3

33

may be a good candidate for use as an effective biological preservative in agriculture

34

and food industry.

35 36

KEYWORDS: Kefir, antimicrobial peptide, purification, identification, antibacterial

37

effects

38 39 40 41 42

1

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

Journal of Agricultural and Food Chemistry

43

INTRODUCTION

44

Foodborne illness caused by microbial contamination results in large economic

45

losses in food industry each year.1-3 Although traditional heat sterilization can

46

effectively inhibit microbial growth in foods, the obtained sterile food products are

47

always partly decreased in the quality of sensory, nutrition and biological function.

48

Compared with traditional heat sterilization technology, the synthetic preservatives

49

can fully retain food nutrients and the original flavor, however consumers are

50

increasingly concerned about the safety of the synthetic chemicals used as

51

preservatives in food. Therefore, development of effective, natural antimicrobial

52

substances for food preservation and food safety is in great demand in food industry.

53

Extensive research has investigated the potential application of natural antimicrobial

54

agents in food preservation, such as plant essential oils and their constituents,4,5

55

animal origin antimicrobial agents,6,7 and microbial origin antimicrobial agents.8,9

56

Among these natural antimicrobial agents, antimicrobial peptides have been under

57

consideration as promising candidates because they provide nutrition, no special

58

flavor, and have special mechanism of action compared with traditional antibiotics.10

59

Nisin, a bacteriocin produced by Lactococcus lactis subsp. lactis was the first natural

60

antimicrobial peptide approved for food use and is the only antimicrobial peptide

61

widely employed as a food preservative in over 40 countries.11,12

62

Kefir is a traditional, natural fermented milk beverage fermented with a starter kefir

63

grains, which mainly composed of three types of microorganisms: lactic acid bacteria,

64

acetic acid bacteria, and yeasts.13,14 Kefir has been used for centuries in many 2

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

65

countries. As a traditional safe fermented milk, numerous studies have focused on the

66

biological functions of kefir in recent years. Kefir is believed to have a number of

67

beneficial properties,15 including the clinical treatment of metabolic diseases,

68

hypertension, ischemic heart disease (IHD), and allergies.16 Among these properties,

69

the antimicrobial activity displayed by kefir has attracted more attention.17-20 The

70

substances in kefir that may be responsible for its antimicrobial activity include lactic

71

acid, volatile acids, and inorganic compounds and antimicrobial peptides produced by

72

lactic acid bacteria.17 Some studies reported the antimicrobial activity of kefir,

73

however the exact substance compositions were not clear, because the antibacterial

74

effect were contributed by fermentation broth,18,19,21 a mixture made up of many

75

different substances ( such as polysaccharides, proteins, peptides, lactic acid or acetic

76

acid). Therefore, purification and obtaining pure antimicrobial substance are key

77

processes in researching the antimicrobial activity of kefir, which are preconditions to

78

identify the chemical structure and to further study the antibacterial effects and

79

antibacterial mechanisms.

80

In the present study, antimicrobial peptide F3 was purified from kefir by a

81

three-step purification procedure. The chemical structure of F3 was identified by

82

MALDI–TOF MS, NMR experiments and X-ray fluorescence analysis. The effect of

83

F3 on the integrity of the bacterial membrane was investigated by the outer membrane

84

/inner membrane permeability assay and confocal laser scanning microscopy assay.

85

The effect of F3 on the bacterial morphology was analyzed using scanning electron

86

microscope (SEM). Our results provided fundamental information on the purification, 3

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Journal of Agricultural and Food Chemistry

87

identification and antimicrobial activity of the natural antimicrobial peptide F3 from

88

kefir.

89

MATERIALS AND METHODS

90

Preparation of the kefir fermentation broth. Tibetan kefir grains were obtained

91

from the laboratory of food microorganisms in the College of Food Science at South

92

China Agricultural University (Guangzhou, China). Kefir grains (2%, m/v) were

93

added to sterilized milk and kept hermetically at 37 °C for 24 h. After fermentation,

94

the kefir grains were filtered. The fermentation broth was obtained to analyze its

95

antibacterial activity. The broth was centrifuged at 3500 g for 20 min and the

96

supernatant was filtered through a 0.22-µm millipore filter. The cell-free supernatant

97

was used for subsequent purification process.

98

Screening for antimicrobial activity components. To confirm the antimicrobial

99

activity of each purified fraction from the cell-free supernatant, the agar spot-test 22

100

method

was applied where the diameter of the inhibitory zone was used as an

101

indicator of antimicrobial activity. Escherichia coli (E. coli) and Staphylococcus

102

aureus (S. aureus) were used as test microorganisms.

103

Purification of fraction with the highest antimicrobial activity. All the eluted

104

fractions were recorded at 214 nm during the purification. The cell-free supernatant

105

was initially purified via a Sephadex G-50 gel filtration column (70 cm × 1.6 cm) at

106

20 °C. The eluent used was doubly distilled water with the flow rate of 1.5 mL/min. 4

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

107

The eluted fractions (SP-1, SP-2 and SP-3) were then collected, concentrated, and

108

lyophilized before the antimicrobial activity assay.

109

After purification by gel filtration chromatography, the dried fraction displaying the

110

highest antimicrobial activity was redissolved in doubly distilled water and loaded

111

onto a DEAE-52 ion exchange column (40 cm × 2.6 cm) at 20 °C. The eluent gradient

112

used was doubly distilled water (0-3 min), 5.0 mM NH4OAc (4-10 min), 10 mM

113

NH4OAc (10-15 min), 20 mM NH4OAc (16-25 min), and 100 mM NH4OAc (26-60

114

min). The flow rate was 1.5 mL/min. The eluted fractions (W-1 and Y-2) were

115

collected, concentrated, and lyophilized for the antimicrobial assay.

116

The fraction obtained from ion exchange chromatography displaying the highest

117

antimicrobial activity was redissolved in doubly distilled water and loaded onto a

118

reversed-phase (RP) Shim-pack PRC-ODS (K) column (250 mm × 30 mm, 15µm,

119

Shimadzu, Kyoto, Japan). Solvent A was 0.1% trifluoroacetic acid in double-distilled

120

water, and solvent B was 100% methanol. A linear gradient of 5%–20% solvent B was

121

applied at a flow rate of 1 mL/min for 80 min. Eluted peaks (F0, F1, F2 and F3) were

122

collected, concentrated, and lyophilized for use in the antimicrobial activity assay. The

123

highest antimicrobial activity fraction was then analyzed on a C18 column (300 mm ×

124

3.9 mm, 4 µm) (Waters, Boston, MA) using a 5% methanol isocratic elution at a flow

125

rate of 1 mL/min to determine the purity.

126

Identification of the fraction with the highest antimicrobial activity. Ninhydrin

127

test was applied to determine whether the fraction with the highest antimicrobial

128

activity contains peptide or protein.23 One drop of the fraction was added onto a silica 5

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

Journal of Agricultural and Food Chemistry

129

gel plate and dried using a hair drier. A ninhydrin solution (0.25% ninhydrin dissolved

130

in ethanol) was then sprayed on the silica gel plate. The silica gel plate was then kept

131

at 100 °C for 5 min. A color change to purple indicates the test material may contain a

132

peptide or protein. Molecular weight of the fraction displaying the largest

133

antimicrobial activity isolated from RP-HPLC was determined using an ABI 4800

134

MALDI-TOF-MS (Shimadzu, Kyoto, Japan). 1H and

135

AV-600 spectrometer (Bruker Spectrospin AG, Rheinstetten, Germany) at room

136

temperature (1H NMR, 400 MHz;

137

were measured in ppm. The fraction with the highest antimicrobial activity (6 mg)

138

was dissolved in 0.6 ml D2O in a 5 mm NMR tube. The element compositions were

139

then characterized with an S4 PIONEER X-ray fluorescence spectrometer (Bruker,

140

Karlsruhe, Germany).24

141

Determination of the antimicrobial spectrum and minimum inhibitory

142

concentration (MIC). The spectrum of antimicrobial activity and minimum

143

inhibitory concentration (MIC) of the fraction displaying the largest antimicrobial

144

activity was determined using a microdilution technique.20,25 The bacterial strains

145

tested were Escherichia coli (ATCC 25922), Staphylococcus aureus (ATCC 63589),

146

Salmonella enterica (CMCC 9812), Shigella dysenteriae (CMCC(B)50071) and

147

Bacillus thuringiensis (CMCC 9812), and the fungi strains were Aspergillus niger

148

(ACCC 30005), Aspergillus flavus (CGMCC 3. 2890), Rhizopus nigricans (AS3.4997)

149

and Penicillium glaucum (STL 3501). Samples were dissolved in sterilized water and

150

diluted to an initial concentration of 2 mg/mL.

13

13

C NMR were recorded on an

C NMR, 100 MHz). Chemical shifts (δ scale)

6

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 32

151

Membrane permeability assay. E. coli and S. aureus were chosen to assess the cell

152

membrane damaging ability of fraction peptide F3. The outer membrane permeability

153

assay for E. coli was performed using a synergistic growth inhibition assay in the

154

presence of erythromycin and F3. E. coli was cultured to logarithmic phase, washed,

155

and re-suspended in LB broth at 106 CFU/mL. 0.5 MIC (62.5 µg/mL) of F3 with

156

different concentrations of erythromycin (1 µg/mL, 2 µg/mL, 4 µg/mL, 7 µg/mL, 13

157

µg/mL, or 25 µg/mL) were incubated with the cells at 37°C for 10 h.26 Synergistic

158

growth inhibition was monitored by detecting a decrease in absorbance at 630 nm.

159

Permeability of the E. coli inner cell membrane and the S. aureus cell membrane

160

was

determined

by

detection

of

β-Galactosidase

161

O-nitrophenyl-β-D-galactopyranoside

162

compound, as a substrate.27 E. coli and S. aureus were cultured to logarithmic phase,

163

washed, and re-suspended in 10 mM sodium phosphate buffer (pH 7.4). ONPG (1.5

164

mM) was dissolved in the same buffer and F3 (final concentration of F3 was 1 MIC)

165

was incubated with E. coli or S. aureus. The degree of permeability was determined

166

every 30 min by monitoring the hydrolysis of ONPG to o-nitrophenol at 405 nm. A

167

cell suspension without F3 was used as control.

168

Confocal laser Scanning Microscopy. Fluorescein isothiocyanate (FITC)-labeled

169

peptide F3 was prepared as described previously.28 An E. coli cell suspension grown

170

until the exponential phase (108 CFU/mL) was mixed with FITC- labeled F3 to a final

171

concentration equal to 1MIC (125 µg/mL). Samples were kept in the dark at 37 °C. At

172

time points of 10, 30, and 180 min, cells were washed with PBS buffer three times

(ONPG),

a

activity

non-membrane

7

ACS Paragon Plus Environment

using

permeative

Page 9 of 32

Journal of Agricultural and Food Chemistry

173

and observed using a confocal laser-scanning microscope (Zeiss, Berlin, Germany).

174

Examination of bacterial membrane damage by scanning electron microscopy

175

(SEM). The test strain was grown to logarithmic phase in LB broth. Cell suspension

176

(106 CFU/mL) was incubated with F3 (final concentration of 1 MIC) at 37°C for

177

various time periods (0.5, 1, and 4 h). After pelleted by centrifugation at 3,000 g for 5

178

min, cell morphology was observed under an XL30 ESEM scanning electron

179

microscope (Philips, Eindhoven, Netherlands) according to a previously published

180

method.29

181

RESULTS AND DISCUSSION

182

Purification and molecular weight determination of the fraction with the highest

183

antimicrobial activity. The results from the purification process are shown in Figure

184

1 and Table 1. The highest antimicrobial activity fraction was obtained by a three-step

185

purification procedure (Figure 1A, B and C). The most active peak, F3, which eluted

186

at 48.02 min during the third step of the three-step purification procedure (Figure 1C)

187

exhibited the highest antimicrobial activity against E.coli and S. aureus with

188

diameters measured to be 7.4±0.18 mm and 8.1±0.41 mm, respectively. This active

189

peak (F3) was then loaded on a reversed-phase (RP) C18 column to evaluate its purity.

190

A single peak was found at 19.24 min (Figure 1D), highlighting the high purity of F3.

191

Fraction F3 was then collected and subjected to MALDI–TOF MS to identify its

192

molecular mass (Figure 1E). The m/z ratio of the major peak was found to be 453.1.

193

The more accurate mass of peptide F3 determined by high resolution mass 8

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 32

194

spectrometry was 453.16861 Da, which is similar to the result of 453.18 Da by

195

MALDI–TOF MS. Compared to previously reported antimicrobial peptides isolated

196

from kefir and other materials, such as bacST8KF (3.5 kDa),17 bacteriocin F1

197

(2113.842 kDa),20 sakacin C2 (2113.842 kDa),30 bacteriocin A5-11A (5206 Da) and

198

A5-11B (5218 Da),31 fraction F3 did not match with any of these known antimicrobial

199

peptides.

200

Identification of the highest antimicrobial activity fraction. The ninhydrin test is a

201

classic method used to confirm and characterize amino acids, peptides, and

202

proteins.32-35 This test was used to determine whether F3, the fraction with the highest

203

antimicrobial activity, was composed of amino acids, peptides or proteins. The single

204

purple spot (positive color change) on the silica gel plate corresponding to fraction F3

205

confirmed that it was composed of amino acids, peptides, or proteins. With the

206

ninhydrin test and molecular weight determination, we proposed that F3 is a low

207

molecular weight peptide. 1

208

H and

13

C NMR spectra were then used to identify the chemical structure of F3

209

(Figure 2). The proton peaks at 6.9 and 7.1 ppm (each 2H, d, J = 6.6 Hz) and at 6.5–

210

7.5 ppm in the 1H NMR spectra (Figure 2A) were attributed to the proton signals of

211

the benzene ring of tyrosine. Peaks at 157.7, 133.7, 118.7 and 120.2 ppm in the

212

NMR spectra (Figure 2B) were attributed to the carbon signals of the benzene ring of

213

tyrosine. This confirmed the presence of a tyrosine residue in F3. Peaks at 3-4 ppm in

214

the 1H NMR spectra (Figure 2A) and peaks at 23.8, 24.9, 27.0, 42.6 and 56.3 in the

215

13

13

C

C NMR spectra (Figure 2B) corresponded to methyl and methylene groups of 9

ACS Paragon Plus Environment

Page 11 of 32

Journal of Agricultural and Food Chemistry

216

leucine, therefore F3 contains a leucine. Considering the determined molecular weight

217

of 452.1 Da, F3 should contain other post-translational modification groups except for

218

tyrosine and leucine.

219

The elemental composition of F3 was measured by X-ray fluorescence. This

220

experiment showed that, except for O, C, and N, F3 contains other elements including

221

P (2.107%), Sr (0.001%), Cl (0.031%), Si (0.066% SiO2), Al (0.01%), Na (0.173%),

222

Ca (0.009%), Ni (0.002%), Zn (0.001%), and Ga (0.001%). After eliminating system

223

errors caused by equipment, we confirmed that F3 contained the phosphorus element.

224

In the present study, no signals for other amino acids were found in the 1H and

225

NMR spectra aside from tyrosine and leucine. With the molecular weight, NMR

226

experiments and X-ray fluorescence analysis in hand, we concluded that there are two

227

phosphate radicals on F3. The proposed chemical structure of F3 contains two amino

228

acids (leucyl-tyrosine or tyrosyl-leucine) with two phosphate radicals connected to the

229

benzene ring of tyrosine. Based on the analysis by Edman degradation, the tentative

230

molecular formula of F3 is considered to be tyrosyl-leucine with two phosphate

231

radicals connected to the benzene ring of tyrosine (Figure 3). Several reported short

232

cationic antibacterial peptides, such as some dipeptides or tripeptides, were

233

surprisingly active compared to many far larger size antibacterial peptides, especially

234

against multi-resistant strains.36, 37 Post-translational modifications have been found in

235

several antimicrobial peptides,38 and some modifications are more complex and

236

extensive.39,40 The biological role of post-translational modifications probably make

237

the peptides resistance to endogenous proteolytic enzymes, and increase their activity 10

ACS Paragon Plus Environment

13

C

Journal of Agricultural and Food Chemistry

238

stability. 38 The proposed chemical structure showed that F3 is a dipeptide with two

239

phosphate radicals connected to the C-terminal, which indicated F3 is a novel

240

antimicrobial peptide and worthy of further study.

241

Antimicrobial spectrum and minimum inhibitory concentration (MIC). The

242

microdilution technique was used to further assess the antimicrobial activity of F3.

243

Table 2 shows that F3 displayed antimicrobial activity against several bacteria and

244

fungi with MIC values ranging from 125 to 500 µg/mL. E. coli and Bacillus

245

thuringiensis exhibited the highest sensitivity to F3 with a MIC value of 125 µg/mL.

246

Although F3 was initially screened with E. coli, a Gram-negative bacterial strain and

247

S. aureus, a Gram-positive bacterial after purification, antimicrobial activity against

248

several fungi was also observed. These results were similar to an antimicrobial

249

peptide reported in a previous study from our group.20 Antimicrobial peptides showing

250

antimicrobial activity against Gram-positive and Gram-negative bacteria, and fungi

251

were not common in previous reports, since the majority of antimicrobial peptides

252

only preserved antibacterial activity against Gram-positive bacteria.30,41 Our results

253

suggested that F3 was an antimicrobial peptide with a broad antimicrobial spectrum.

254

Effect of the peptide F3 on the permeability of bacterial membranes. Unlike

255

Gram-positive bateria, Gram-negative bacteria possess an outer membrane attached to

256

the peptidoglycan layer. This membrane acts as a selective permeability barrier,

257

reinforcing the shape of the cell and providing a protective barrier against harmful

258

agents in the external environment.42 Erythromycin is less effective against 11

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Journal of Agricultural and Food Chemistry

259

Gram-negative bacteria because of its poor permeation of the outer membrane,

260

especially at the lower concentrations, while it can quickly cross through a damaged

261

outer membrane and exert detrimental effects on cells.43 In our previous experiment,

262

we found that 0.5 MIC of F3 did not cause any significant inhibition to E. coli growth,

263

therefore we chose erythromycin as a probe in conjunction with 0.5 MIC F3. This

264

would allow us to determine whether F3 has the penetration enhancing ability. The

265

results are shown in Figure 4. In the absence of F3, erythromycin showed

266

dose-dependent inhibition on E. coli growth. When a low concentration of

267

erythromycin and 0.5 MIC F3 were added to the cell culture, an enhanced inhibitory

268

effect was observed on the growth of E. coli. This result indicated that F3 increased

269

the permeability of the outer membrane of E. coli to allow erythromycin to enter cells

270

and kill the bacteria at a lower concentration. Several reported antimicrobial

271

substances also exhibit similar outer membrane permeability activity in E. coli.26,44

272

O-nitrophenyl-β-D-galactopyranoside (ONPG) is an analog of lactose which can

273

be hydrolyzed into a yellow product, ο-nitrophenol, by β-Galactosidase, a hydrolase

274

inside the cytoplasm.45 β-Galactosidase can be released from cells when the

275

cytoplasmic membrane is permeable. Therefore, ONPG was chosen as a probe to

276

determine the permeability of the cytoplasmic membrane. As shown in Figure 4,

277

β-Galactosidase leakage was observed from E. coli (Figure 4B) and S. aureus (Figure

278

4C) in the presence of F3 in a time-dependent manner. No significant change of

279

β-Galactosidase activity was detected in the culture medium composed of control

280

cells. These results suggested that F3 had the ability to cause permeability of the 12

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

281

cytoplasmic membrane of E. coli and S. aureus.

282

Internalization of FITC-labeled F3 in Escherichia coli cells. The results of the cell

283

membrane permeability test described above indicated that F3 was able to increase

284

bacterial membrane permeability. To determine the localization of F3, FITC-labeled

285

F3 was incubated with E. coli cells and was visualized by confocal laser scanning

286

microscopy. From the observed results showed in Figure 5, F3 was found to localize

287

in the cytoplasm of bacterial cells in a time dependent manner between 10 and 180

288

minutes. This further confirms the membrane permeability activity of F3.

289

Examination of morphologic changes in cells by scanning electron microscopy.

290

The surface morphology of E. coli and S. aureus in the presence of F3 were evaluated

291

using SEM. As shown in Figure 6, untreated E. coli (Figure 6A) and S. aureus (Figure

292

6E) cells showed a normal, smooth, intact surface. After treatment with F3 for 2 h, E.

293

coli (Figure 6D) and isolated S. aureus (Figure 6H) cells showed a significant change

294

in morphology indicated by the presence of wrinkles and deformation on cellular

295

surfaces. These results indicated that F3 damages the cellular integrity and caused

296

significant deformation in cell morphology. The ultrastructure changes observed on

297

cell surfaces caused by F3 were similar to other reported results.46-47

298

The present study demonstrates that peptide F3 is a novel dipeptide, especially

299

with the post-translational modification of phosphate radicals, providing a new

300

reference in the field of short peptide antibiotics. The special chemical structure of F3

301

may contribute to its broad antimicrobial spectrum against Gram-positive bacteria, 13

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Journal of Agricultural and Food Chemistry

302

Gram-negative bacteria and fungi, which is definitely worth further exploring. In

303

addition, the membrane-penetrating activity of F3 provides further insight into the

304

interaction between antimicrobial peptides and bacterial membranes. F3 has the

305

potential to serve as a valuable natural food preservative in food industry.

306

AUTHOR INFORMATION

307

Corresponding Author

308

* Telephone: +86-20-85286234. Fax: +86-20-85286234. E-mail:

309

[email protected].

310

* Telephone: 848-932-5514. Fax: 732-932-6776. E-mail: [email protected].

311

Funding

312

The authors would like to express their gratitude to the National Natural Science

313

Foundation of China (No. 31171768) and the Scientific Research Project of

314

Guangdong Province Office of Education (No. 2013gjhz0003) for the financial

315

support.

316

Notes

317

The authors declare no competing financial interest.

318

319

320

321

322

14

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 32

323

REFERENCES

324

(1) Scallan, E.; Hoekstra, R. M.; Angulo, F. J.; Tauxe, R. V.; Widdowson, M. A.; Roy,

325

S. L.; Jones, J. L.; Griffin, P. M., Foodborne illness acquired in the United States-

326

major pathogens. Emerg. Infect. Dis. 2011, 17,1-21.

327

(2) Lanzas, C.; Lu, Z.; Gröhn, Y. T., Mathematical modeling of the transmission and

328

control of foodborne pathogens and antimicrobial resistance at preharvest. Foodborne

329

Pathog. Dis. 2011, 8, 1-10.

330

(3) Thomas, M. K.; Murray, R.; Flockhart, L.; Pintar, K.; Pollari, F.; Fazil, A.; Nesbitt,

331

A.; Marshall, B., Estimates of the burden of foodborne illness in Canada for 30

332

specified pathogens and unspecified agents, circa 2006. Foodborne Pathog. Dis.2013,

333

10, 639-648.

334

(4) Burt, S. A.; Reinders, R. D., Antibacterial activity of selected plant essential oils

335

against Escherichia coli O157: H7. Lett. Appl. Microbiol. 2003, 36, 162-167.

336

(5) Alzoreky, N.; Nakahara, K., Antibacterial activity of extracts from some edible

337

plants commonly consumed in Asia. Int. J. Food Microbiol. 2003, 80, 223-230.

338

(6) Borch, E.; Wallentin, C.; Rosén, M.; Björck, L., Antibacterial effect of the

339

lactoperoxidase/thiocyanate/hydrogen

340

Campylobacter isolated from poultry. J. Food Prot. 1989, 52, 638-641.

341

(7) Selsted, M. E.; Tang, Y. Q.; Morris, W.; McGuire, P.; Novotny, M.; Smith, W.;

342

Henschen, A.; Cullor, J., Purification, primary structures, and antibacterial activities

343

of beta-defensins, a new family of antimicrobial peptides from bovine neutrophils. J.

344

Biol. Chem. 1993, 268, 6641-6648.

peroxide

system

15

ACS Paragon Plus Environment

against

strains

of

Page 17 of 32

Journal of Agricultural and Food Chemistry

345

(8) Bhunia, A.; Johnson, M.; Ray, B., Purification, characterization and antimicrobial

346

spectrum of a bacteriocin produced by Pediococcus acidilactici. J. Appl. Bacteriol.

347

1988, 65, 261-268.

348

(9) Miao, J.; Zhou, J.; Liu, G.; Chen, F.; Chen, Y.; Gao, X.; Dixon, W.; Song, M.;

349

Xiao, H.; Cao, Y., Membrane disruption and DNA binding of Staphylococcus aureus

350

cell induced by a novel antimicrobial peptide produced by Lactobacillus paracasei

351

subsp. tolerans FX-6. Food Control 2016, 59, 609-613.

352

(10) Zasloff, M., Antimicrobial peptides of multicellular organisms. Nature 2002, 415,

353

389-395.

354

(11) Cleveland, J.; Montville, T. J.; Nes, I. F.; Chikindas, M. L., Bacteriocins: safe,

355

natural antimicrobials for food preservation. Int. J. Food Microbiol. 2001, 71, 1-20.

356

(12) da Silva Malheiros, P.; Daroit, D. J.; Brandelli, A., Food applications of

357

liposome-encapsulated antimicrobial peptides. Trends Food Sci. Tech. 2010, 21,

358

284-292.

359

(13) Abraham, A. G.; de Antoni, G. L., Characterization of kefir grains grown in cows'

360

milk and in soya milk. J. Dairy Res. 1999, 66, 327-333.

361

(14) Garrote, G. L.; Abraham, A. G.; de Antoni G. L., Chemical and microbiological

362

characterisation of kefir grains. J. Dairy Res. 2001, 68, 639-652.

363

(15) Farnworth, E. R., Kefir: from folklore to regulatory approval. J. Nutr. Funct. Med.

364

Foods 1999, 1, 57-68.

365

(16) St-Onge, M.-P.; Farnworth, E. R.; Savard, T.; Chabot, D.; Mafu, A.; Jones, P. J.,

366

Kefir consumption does not alter plasma lipid levels or cholesterol fractional synthesis 16

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

367

rates relative to milk in hyperlipidemic men: a randomized controlled trial. BMC

368

Complem. Altern. Med. 2002, 2, 1-7.

369

(17) Powell, J.; Witthuhn, R.; Todorov, S.; Dicks, L., Characterization of bacteriocin

370

ST8KF produced by a kefir isolate Lactobacillus plantarum ST8KF. Int.Dairy J. 2007,

371

17, 190-198.

372

(18) Rodrigues, K. L.; Caputo, L. R. G.; Carvalho, J. C. T.; Evangelista, J.;

373

Schneedorf, J. M., Antimicrobial and healing activity of kefir and kefiran extract. Int.

374

J. Antimicrob.Ag. 2005, 25, 404-408.

375

(19) Yüksekdağ, Z.; Beyatli, Y.; Aslim, B., Determination of some characteristics

376

coccoid forms of lactic acid bacteria isolated from Turkish kefirs with natural

377

probiotic. LWT-Food Sci. Technol. 2004, 37, 663-667.

378

(20) Miao, J.; Guo, H.; Ou, Y.; Liu, G.; Fang, X.; Liao, Z.; Ke, C.; Chen, Y.; Zhao, L.;

379

Cao, Y., Purification and characterization of bacteriocin F1, a novel bacteriocin

380

produced by Lactobacillus paracasei subsp. tolerans FX-6 from Tibetan kefir, a

381

traditional fermented milk from Tibet, China. Food Control 2014, 42, 48-53.

382

(21) Santos, A.; San Mauro, M.; Sanchez, A.; Torres, J.; Marquina, D., The

383

antimicrobial properties of different strains of Lactobacillus spp. isolated from kefir.

384

Syst. Appl. Microbiol. 2003, 26, 434-437.

385

(22) Van Reenen, C.; Dicks, L.; Chikindas, M., Isolation, purification and partial

386

characterization of plantaricin 423, a bacteriocin produced by Lactobacillus

387

plantarum. J. Appl. Microbiol. 1998, 84, 1131-1137.

388

(23) Patowary, K.; Saikia, R. R.; Kalita, M. C.; Deka, S., Degradation of polyaromatic 17

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

Journal of Agricultural and Food Chemistry

389

hydrocarbons

employing

biosurfactant-producing

Bacillus

390

Ann.Microbiol. 2015, 65, 225-234.

391

(24) Liu, Y.; Zhang, L.; Cheng, L.; Xu, Y.; Liu, Y., Design, fabrication, and

392

characterization of silicon nitride particle-reinforced silicon nitride matrix composites

393

by chemical vapor infiltration. Int. J. Appl. Ceram. Tec. 2010, 7, 63-70.

394

(25) Rahman, M. M.; Gray, A. I., A benzoisofuranone derivative and carbazole

395

alkaloids from Murraya koenigii and their antimicrobial activity. Phytochemistry 2005,

396

66, 1601-1606.

397

(26) Hao, G.; Shi, Y. H.; Tang, Y. L.; Le, G. W., The membrane action mechanism of

398

analogs of the antimicrobial peptide buforin 2. Peptides 2009, 30, 1421-1427.

399

(27) Pellegrini, A.; von Fellenberg, R., Design of synthetic bactericidal peptides

400

derived from the bactericidal domain P 18–39 of aprotinin. BBA-Protein Struct. M.

401

1999, 1433, 122-131.

402

(28) Helmerhorst, E. J.; Breeuwer, P.; van‘t Hof, W.; Walgreen-Weterings, E.; Oomen,

403

L. C.; Veerman, E. C.; Amerongen, A. V. N.; Abee, T., The cellular target of histatin 5

404

on Candida albicans is the energized mitochondrion. J. Biol. Chem.1999, 274,

405

7286-7291.

406

(29) Li, L.; Shi, Y.; Cheng, X.; Xia, S.; Cheserek, M. J.; Le, G., A cell-penetrating

407

peptide analogue, P7, exerts antimicrobial activity against Escherichia coli

408

ATCC25922 via penetrating cell membrane and targeting intracellular DNA. Food

409

Chem. 2015, 166, 231-239.

410

(30) Gao, Y.; Jia, S.; Gao, Q.; Tan, Z., A novel bacteriocin with a broad inhibitory 18

ACS Paragon Plus Environment

pumilus

KS2.

Journal of Agricultural and Food Chemistry

411

spectrum produced by Lactobacillus sake C2, isolated from traditional Chinese

412

fermented cabbage. Food Control 2010, 21, 76-81.

413

(31) Batdorj, B.; Dalgalarrondo, M.; Choiset, Y.; Pedroche, J.; Metro, F.; Prévost, H.;

414

Chobert, J. M.; Haertlé, T., Purification and characterization of two bacteriocins

415

produced by lactic acid bacteria isolated from Mongolian airag. J. Appl. Microbiol.

416

2006, 101, 837-848.

417

(32) Friedman, M., Applications of the ninhydrin reaction for analysis of amino acids,

418

peptides, and proteins to agricultural and biomedical sciences. J. Agric. Food. Chem.

419

2004, 52, 385-406.

420

(33) De, S.; Das, D. C.; Mandal, T.; Das, M., Analysis of phytochemical profile of

421

Cardanthera difformis Druce whole plant extract with antimicrobial properties. Int.

422

J.Rec. Sci. Res. 2015, 6, 4564-4567.

423

(34) Rathore, R.; Rahal, A.; Mandil, R.; Prakash, A.; Garg, S., Comparison of the

424

antiinflammatory activity of plant extracts from Cimicifuga racemosa and Mimosa

425

pudica in a rat model. Aust. Vet. Pract. 2012, 42, 275-278.

426

(35) Sharma, V.; Agarwal, A.; Chaudhary, U.; Singh, M., Phytochemical investigation

427

of various extracts of leaves and stems of Achyranthes aspera Linn. Int. J.

428

Pharm .Pharm. Sci. 2013, 5, 317-20.

429

(36) Strøm, M. B.; Haug, B. E.; Skar, M. L.; Stensen, W.; Stiberg, T.; Svendsen, J.

430

S., The pharmacophore of short cationic antibacterial peptides. J. Med. Chem. 2003,

431

46: 1567-1570.

432

(37) Rajendra Narayan, M.; Anshupriya, S.; Pritha, P.; 19

ACS Paragon Plus Environment

Prasanta Kumar, D.,

Page 20 of 32

Page 21 of 32

Journal of Agricultural and Food Chemistry

433

Antimicrobial activity, biocompatibility and hydrogelation ability of dipeptide-based

434

amphiphiles. Org. Biomol. Chem. 2008, 7, 94-102.

435

(38) Sperstad, S. V.; Haug, T.; Blencke, H. M.; Styrvold, O. B.; Li, C.; Stensvåg, K.,

436

Antimicrobial peptides from marine invertebrates: challenges and perspectives in

437

marine antimicrobial peptide discovery. Biotechnol. Adv. 2011, 29, 519-530.

438

(39) Lee, I.H., Cho, Y., Lehrer, R.I., Styelins, broad-spectrum antimicrobial peptides

439

from the solitary tunicate, styela clava. Comp. Biochem. Phys. B. 1997, 118: 515-521.

440

(40) Noga, E.J.; Stone, K.L.; Wood, A.; Gordon, W.L.; Robinette, D., Primary

441

structure and cellular localization of callinectin, an antimicrobial peptide from the

442

blue crab. Dev. Comp. Immunol., 2011, 35: 409-415.

443

(41) Messaoudi, S.; Kergourlay, G.; Dalgalarrondo, M.; Choiset, Y.; Ferchichi, M.;

444

Prévost, H.; Pilet, M. F.; Chobert, J.M.; Manai, M.; Dousset, X., Purification and

445

characterization of a new bacteriocin active against Campylobacter produced by

446

Lactobacillus salivarius SMXD51. Food Microbiol. 2012, 32, 129-134.

447

(42)Amro, N. A.; Kotra, L. P.; Wadu-Mesthrige, K.; Bulychev, A.; Mobashery, S.; Liu,

448

G.-y., High-resolution atomic force microscopy studies of the Escherichia coli outer

449

membrane: structural basis for permeability. Langmuir 2000, 16, 2789-2796.

450

(43) Vaara, M.; Porro, M., Group of peptides that act synergistically with hydrophobic

451

antibiotics against Gram-negative enteric bacteria. Antimicrob. Agents Chemother.

452

1996, 40, 1801-1805.

453

(44) da Silva, I. F.; de Oliveira, R. G.; Soares, I. M.; da Costa Alvim, T.; Ascêncio, S.

454

D.; de Oliveira Martins, D. T., Evaluation of acute toxicity, antibacterial activity, and 20

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

455

mode of action of the hydroethanolic extract of Piper umbellatum L. J.

456

Ethnopharmacol. 2014, 151, 137-143.

457

(45) Marri, L.; Dallai, R.; Marchini, D., The novel antibacterial peptide ceratotoxin A

458

alters permeability of the inner and outer membrane of Escherichia coli K-12. Curr.

459

Microbiol. 1996, 33, 40-43.

460

(46) Mine, Y.; Ma, F.; Lauriau, S., Antimicrobial peptides released by enzymatic

461

hydrolysis of hen egg white lysozyme. J. Agr. Food Chem. 2004, 52, 1088-1094.

462

(47) Lee, N. K.; Jin Han, E.; Jun Han, K.; Paik, H. D., Antimicrobial effect of

463

bacteriocin KU24 produced by Lactococcus lactis KU24 against methicillin resistant

464

Staphylococcus aureus. J. Food Sci. 2013, 78, M465-M469.

465 466 467 468 469 470 471 472 473 474 475 476 21

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

Journal of Agricultural and Food Chemistry

477

FIGURE CAPTIONS

478

Figure 1. Purification of peptide F3. A. Analysis of active fractions in the supernatant;

479

B. Analysis of active fraction SP-3 obtained from A; C. Analysis of active fraction

480

W-1 obtained from B; D. Analysis of active fraction F3 obtained from C; E.

481

MALDI-TOF MS analysis of F3.

482

Figure 2. A. 1H, and B. 13C NMR spectra of F3.

483

Figure 3. The tentative chemical structure of F3.

484

Figure 4. Effects of F3 on the membrane permeability of E. coli and S. aureus cells.

485

A. Effects on the outer membrane permeability of Escherichia coli cells. ( )

486

Erythromycin (1 µg/mL, 2 µg/mL, 4 µg/mL, 7 µg/mL, 13 µg/mL, or 25 µg/mL); (

487

0.5 MIC F3 and erythromycin (1 µg/mL, 2 µg/mL, 4 µg/mL, 7 µg/mL, 13 µg/mL, or

488

25 µg/mL); and ( ) sterile water. B. Effects on the inner membrane permeability of E.

489

coli cells. (

490

of S. aureus cells. ( ) 1 MIC F3; (×) control. The mean and standard deviation of

491

triplicate values are shown.

492

Figure 5. E. coli cells treated with FITC-conjugated F3. A. 10 min; B. 30 min; and C.

493

180 min.

494

Figure 6. Scanning electron microscopy observation in E.coli (A, B, C, and D) and S.

495

aureus (E, F, G, and H) treated with 1 MIC of F3 for 0 h (A and E), 0.5 h (B and F), 1

496

h (C and G) and 2 h (D and H).

)

) 1 MIC F3; (×) control. C. Effects on the cell membrane permeability

497

22

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 32

Tables Table 1 Antimicrobial activity of each purified fraction.

Purified fraction

Concentration(mg/mL)

Inhibition zone diameter (mm) Staphylococcus Escherichia coli aureus ATCC 25922 ATCC 63589 – – 3.5±0.21 3.3±0.23

SP-1 SP-2

300 300

SP-3

300

6.9±0.23*

6.5±0.24*

W-1 Y-1 F0 F1 F2

300 300 50 50 50

8.5±0.32** 6.6±0.22 – 4.6±0.17 3.2±0.37

7.2±0.27** 5.8±0.23 – 4.1±0.28 2.9±0.25

F3

50

7.4±0.18***

8.1±0.41***

–: No inhibition zone recorded. * indicates statistical significance between SP-1, SP-2and SP-3 (p < 0.05, n = 3). ** indicates statistical significance between W-1and Y-1 (p < 0.05, n = 3). *** indicates statistical significance between F0, F1, F2 and F3 (p < 0.05, n = 3). The mean and standard deviation of triplicate values are shown.

23

ACS Paragon Plus Environment

Page 25 of 32

Journal of Agricultural and Food Chemistry

Table 2 Minimum inhibitory concentrations (MIC) of the peptide F3.

Microorganism

MIC values (µg/mL) F3

a

Escherichia coli ATCC 25922

125±1.29 a

Staphylococcus aureus ATCC 63589

500±3.23 b

Salmonella enterica CMCC 9812

500±4.14 b

Shigella dysenteriae CMCC(B)50071

500±1.93 b

Bacillus thuringiensis CMCC 9812

125±2.21 a

Aspergillus niger ACCC 30005

500±3.43 b

Aspergillus flavus CGMCC 3. 2890

500±4.73 b

Rhizopus nigricans AS3.4997

500±4.11 b

Penicillium glaucum STL 3501

500±0.23 b

Differences were considered significant at probability levels of p< 0.05, indicated

by different letters. The mean and standard deviation of triplicate values are shown.

24

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 32

Figure 1.

A

1.6 1.4

W-1

1.0

SP-2

Abs at 214nm

1.2

Abs at 214nm

B

1.2

SP-3

1.0 0.8 0.6

0.8

0.6

0.4

SP-1 0.4

Y-1

0.2 0.2

0.0

0.0 0

20

40

60

80

100

120

140

160

0

180

20

40

60

80

Fraction number

Fraction number

F1

C

D F3

F0 F3 F2

E

25

ACS Paragon Plus Environment

Page 27 of 32

Journal of Agricultural and Food Chemistry

Figure 2.

A

B

26

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3.

27

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

Journal of Agricultural and Food Chemistry

Figure 4.

B

OD(405 nm)

A

0.2 0.15 0.1 0.05 0 0

1

2 Time(h)

C

28

ACS Paragon Plus Environment

3

4

Journal of Agricultural and Food Chemistry

Page 30 of 32

Figure 5.

A

B

29

ACS Paragon Plus Environment

C

Page 31 of 32

Journal of Agricultural and Food Chemistry

Figure 6. B

C

E

F

G

H

30

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table of Contents Graphic

31

ACS Paragon Plus Environment

Page 32 of 32