iTRAQ-Based Proteomics Revealed the Bactericidal Mechanism of

Jul 26, 2016 - Sodium new houttuyfonate (SNH), an addition product of active ingredient houttuynin from the plant Houttuynia cordata Thunb., inhibits ...
0 downloads 0 Views 1MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

iTRAQ-Based Proteomics Revealed the Bactericidal Mechanism of Sodium New Houttuyfonate against Streptococcus pneumoniae Xiao-Yan Yang, Tianyuan Shi, Gaofei Du, Wanting Liu, XingFeng Yin, Xuesong SUN, Yunlong Pan, and Qing-Yu He J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b02147 • Publication Date (Web): 26 Jul 2016 Downloaded from http://pubs.acs.org on July 30, 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

iTRAQ-Based Proteomics Revealed the Bactericidal Mechanism of

2

Sodium New Houttuyfonate against Streptococcus pneumoniae

3

Xiao-Yan Yang1, 2, §, Tianyuan Shi2, §, Gaofei Du2, Wanting Liu2, Xing-Feng Yin2,

4

Xuesong Sun2, Yunlong Pan1,* and Qing-Yu He2,* 1

5 2

6

The First Affiliated Hospital of Jinan University, Guangzhou, 510632, China;

Key Laboratory of Functional Protein Research of Guangdong Higher Education

7

Institutes, Institute of Life and Health Engineering, College of Life Science and

8

Technology, Jinan University, Guangzhou 510632, China

9 10 11 12 13 14 15 16 17

§

Equal contributors.

18 19

*Correspondence should be addressed to: Prof. Qing-Yu He, Tel & Fax:

20

+86-20-85227039,

21

+86-20-38374151, E-mail: [email protected].

E-mail:

[email protected];

Prof.

Yunlong

Pan,

Tel:

22 1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 32

23

Abstract

24

Sodium New Houttuyfonate (SNH), an addition product of active ingredient

25

Houttuynin from the plant Houttuynia cordata Thunb, inhibits a variety of bacteria,

26

yet the mechanism by which it induces cell death has not been fully understood. In the

27

present study, we utilized iTRAQ-based quantitative proteomics to analyze the protein

28

alterations in Streptococcus pneumoniae (S. pneumoniae) in response to SNH

29

treatment. Numerous proteins related to the production of reactive oxygen species

30

(ROS) were found to be up-regulated by SNH, suggesting that ROS pathways may be

31

involved as analyzed via bioinformatics. As reported recently, cellular reactions

32

stimulated by ROS including superoxide anion (O2▪-), hydrogen peroxide (H2O2) and

33

hydroxyl radicals (OH▪) have been implicated as one of the mechanisms whereby

34

bactericidal antibiotics kill bacteria. We then validated that SNH killed S. pneumoniae

35

in a dose-dependent manner accompanied with the increasing level of H2O2. On the

36

other hand, addition of catalase, which can neutralize H2O2 in cells, showed a

37

significant recovery in bacterial survival. These results indicate that SNH indeed

38

induced H2O2 formation to contribute to the cell lethality, providing new insights into

39

the bactericidal mechanism of SNH and expanding our understanding of the common

40

mechanism of killing induced by bactericidal agents.

41

Keywords: Streptococcus pneumoniae, Sodium new houttuyfonate, iTRAQ, H2O2,

42

Bactericidal mechanism

43

2

ACS Paragon Plus Environment

Page 3 of 32

Journal of Agricultural and Food Chemistry

44

Introduction

45

One of the most common causes of community-acquired pneumoniae is the infection

46

of Streptococcus pneumoniae, a Gram-positive and catalase-negative bacterium 1. The

47

bacterium also causes many other serious diseases including meningitis, otitis media,

48

bacteremia and sepsis, especially among children, elderly people, patients with AIDS

49

and other immunocompromised individuals, posing a major threat to human health

50

worldwide 2. Due to the increasing incidence of antibiotic resistant clones and the

51

limitations of exiting vaccines, it is urgent to screen for novel drugs to combat

52

pneumococcal infection.

53

Reactive oxygen species (ROS) including superoxide anion (O2▪-), hydrogen

54

peroxide (H2O2) and hydroxyl radicals (OH▪), are produced as by-products of the use

55

of oxygen in fundamental enzymatic reactions and metabolism pathways 1. S.

56

pneumoniae can generate up to 2.0 mM H2O2 through the pyruvate oxidase (SpxB),

57

contributing to its virulence and deleterious effects on itself 1. The overproduced ROS

58

have numerous adverse effects on bacterial DNA damage, protein degradation and the

59

peroxidation of the cell membrane lipids 3. Recently, several studies pointed out that

60

ROS contribute to the cell death of bactericidal antibiotics 4-7.

61

Plant is the source of great interest for curing bacterial infection, by providing us 8

, andrographolide

9

berberine 10

62

with a set of antimicrobial agents, such as artesunate

63

and antimicrobial peptides 11. Houttuynin is the primary constituent in the volatile oil

64

of Houttuynia cordata Thunb, an important herbal medicine widely used as the

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

12

, antiviral

13

and antibacterial

14

Page 4 of 32

65

anti-inflammatory

agent for numerous years in

66

China. It was also used to prevent severe acute respiratory syndrome (SARS) in 2003

67

15

68

houttuyfonate (SH) and sodium new houttuyfonate (SNH), were synthesized and have

69

been approved by China Food and Drug Administration to be used in clinic for the

70

treatment of purulent skin infections, respiratory tract infections, including pneumonia

71

and acute or chronic bronchitis

72

effectively inhibit a variety of bacteria, including Staphylococcus aureus, Bacillus

73

subtilis and Pseudomonas aeruginosa

74

can,

75

Staphylococcus aureus (MRSA). However, despite its widespread and effective use,

76

the mechanism of SNH against bacteria remains unknown.

. Because houttuynin is chemically unstable, its addition products, sodium

alone

and

synergizing

16, 17

. Previous studies indicated that SH and SNH

with

18, 19

. Recently, Lu et al reported

antibiotics,

combat

15

that SNH

Methicillin-Resistant

77

In the present study, we found that SNH killed S. pneumoniae in a

78

dose-dependent manner, then utilized iTRAQ-based quantitative proteomics to dissect

79

the global protein alteration of S. pneumoniae in response to short-term SNH exposure.

80

Bioinformatics analysis of the proteomic data suggested that SNH may induce H2O2

81

formation, and we detected that the H2O2 level indeed increased after SNH treatment.

82

Adding catalase to neutralize H2O2 greatly attenuated SNH-induced cell death. These

83

observations implicate that ROS were induced in S. pneumoniae by the presence of

84

SNH, contributing significantly to the cell death.

85

Materials and methods

4

ACS Paragon Plus Environment

Page 5 of 32

Journal of Agricultural and Food Chemistry

86

1. Bacterial strain and sodium new houttuyfonate

87

S. pneumoniae D39 strain NCTC 7466 was obtained from National Collection of

88

Type Culture (NCTC, London, UK). S. aureus strain ATCC 29213 was purchased

89

from American Type Culture Collection (ATCC). Sodium new houttuyfonate (SNH)

90

was purchased from Chembest company (Shanghai, China).

91

2. Growth conditions, MIC and MBC assays

92

For all experiments, S. pneumoniae was cultured in THY medium (Todd-Hewitt broth

93

(Oxiod, UK) with 0.5% yeast extract) or grown on Columbia agar (Difco, USA)

94

containing 5% sheep blood (Ruite, China) at 37°C in 5% CO2. S. aureus was grown in

95

tryptic soy broth at 37°C with shaking at 200 rpm. The minimum inhibitory

96

concentration (MIC) and minimum bactericidal concentration (MBC) of SNH were

97

measured by the broth micro-dilution method based on guidelines in the literature 20-22.

98

S. pneumoniae (5 × 106 CFU/mL) was incubated with SNH at concentrations of

99

twofold serial dilutions (range from 1.56 µM to 200 µM) in a 24 well microplate (1

100

mL per well) at 37°C in 5% CO2 for 24 h. The bacterial concentration in each well

101

was determined at 600 nm using spectrophotometry (Thermo, USA), the

102

concentration of SNH with no visible bacteria growth (OD600 < 0.1) was recorded as

103

the MIC

104

counting and the concentration of SNH corresponding to the well that produced a

105

99.99% kill relative to the starting inoculum was taken as the MBC

106

(Sigma, USA) was used as a positive control.

22

. Then the wells with no visible bacteria growth were taken into colony

20

. Ampicillin

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 32

107

3. Time–kill curve studies

108

Time-kill assays of S. pneumoniae treated by SNH (corresponding to 0, 1/8 × MIC,

109

1/4 × MIC, 1/2 × MIC, 1 × MIC and 2 × MIC) were performed according to the basic

110

microbiological techniques protocol

111

and viable bacteria were counted after 0, 1, 2, 3, 6, 12 and 24 h of incubation. Colony

112

counts were performed by 10-fold serial dilutions in 1×PBS then plating each onto

113

Columbia agar containing 5% sheep blood at 37°C in 5% CO2 for 24 h. The viable

114

counts were calculated to give CFU/mL, and time-kill curves were shown by plotting

115

log10 CFU/mL against time (h). Ampicillin was used as a positive control. Each

116

time–kill assay was performed in triplicate.

117

4. Determination of growth curves of S. pneumoniae

118

When bacteria were grown to an optical density (OD600) of approximately 0.3, 1 ×

119

MIC SNH was added. For H2O2 quenching, catalase (Beyotime, China) was added at

120

a concentration of 2 mg/mL at the same time as the SNH treatment. CFU/mL was

121

monitored every hour for 3 h after drug addition. For growth curve assay, S.

122

pneumoniae was incubated with SNH (corresponding to 0, 1/16 × MIC, 1/8 × MIC,

123

1/4 × MIC, 1/2 × MIC, 1 × MIC and 2 × MIC), OD600 was monitored every hour for

124

the first 4 h and every two hours for next 8 h after drug addition.

125

5. Cytotoxicity assay

126

The human bronchial epithelial cell line HBE was cultivated in DMEM media with

127

10% fetal bovine serum (FBS) at 37°C in 5% CO2. The cytotoxicity of SNH in HBE

20

. Bacteria were cultured at 37°C in 5% CO2,

6

ACS Paragon Plus Environment

Page 7 of 32

Journal of Agricultural and Food Chemistry

128

cells was assessed using LDH Cytotoxicity Assay Kit (Beyotime, China) according to

129

the manufacturer’s instructions. Briefly, 5 × 103 HBE cells were seeded into a sterile

130

96-well plate in triplicate and incubated at 37 °C with 5% CO2 for 12 h, followed by a

131

48 h incubation in dark with different amounts of SNH (0, 25, 50, 100 and 200 µM).

132

LDH releasing agent was added at 1 hour before the end of the incubation as the

133

positive control. After incubation at 37 °C for 1 h, the supernatant of each well was

134

collected and incubated with LDH working buffer, then the concentration of LDH in

135

each well was detected in a microplate reader at 490 nm.

136

6. Protein sample preparation

137

S. pneumoniae was grown to an optical density (OD600) of approximately 0.3, then

138

exposed to 1/4 × MIC (6.25 µM) of SNH, and samples were taken at 0, 1 and 2 h after

139

addition of the SNH. Bacteria were harvested using centrifugation at 8,000 g for 10

140

min at 4 °C, and washed three times with 1 × PBS, then lysed in SDS lysis buffer by

141

intermittent sonication. The concentrations of the cellular proteins were measured by

142

BCA Protein Assay Kit (Thermo, USA).

143

7. iTRAQ Labeling, SCX fractionation and LC-ESI-MS/MS analysis

144

Proteins from each sample (200 µg) were processed and subjected to iTRAQ (isobaric

145

tags for relative and absolute quantitation) labeling using the AB SCIEX iTRAQ

146

reagents Multiplex kit according to the manufacturers’ instructions. Briefly, proteins

147

were first reduced and alkylated, then digested with trypsin using FASP (filter aided

148

sample preparation) method and finally labeled with 114 (SNH treated 0 h), 116 (SNH

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 32

149

treated 1 h) and 117 (SNH treated 2 h) isobaric tags. The labeled peptides were

150

incubated at room temperature for 2 h and then pooled and dried using vacuum

151

centrifugation. The iTRAQ-labeled peptide mixture was dissolved in 100 µL of high

152

pH buffer A (20 mM NH4CHO2 containing 4% (v/v) acetonitrile, pH 10.0) and

153

separated using reverse-phase UPLC with an Ultremex SCX column (4.6 mm × 250

154

mm, 5 µm, Phenomenex, USA). In brief, iTRAQ-labeled peptide mixture was eluted

155

with a gradient of 5-60% acetonitrile in 20 mM ammonium formate (pH 10.0) for 65

156

min at a flow-rate of 0.8 mL/min. The peptide elution was monitored at 214 nm. After

157

5 minutes, the eluted peptides were collected every minute, pooled into ten fractions

158

and then lyophilized.

159

Dried fractions from the high pH reverse-phase separations were resuspended

160

with buffer A (5% acetonitrile, 0.1% formic acid) and detected using an AB SCIEX

161

Triple-TOF 5600 mass spectrometer (AB SCIEX, USA) coupled with a Nanospray III

162

source and a pulled quartz tip. The identification parameters were used according to

163

previous report 23.

164

8. Downstream processing and analysis of proteomics data

165

The acquired raw data files (.wiff) of each fraction were combined to perform

166

searching against S. pneumoniae D39 protein database, and the proteins were

167

identified and quantified using ProteinPilot™ Software 4.5. The search parameters

168

were set as follows: Sample Type, iTRAQ 4 plex (Peptide Labeled); Cys. Alkylation,

169

Iodoacetic acid; Digestion, Trypsin; Instrument, Triple-TOF 5600; ID Focus,

8

ACS Paragon Plus Environment

Page 9 of 32

Journal of Agricultural and Food Chemistry

170

Biological modifications; Database, S. pneumoniae D39_.fasta; Search Effort,

171

Thorough; Detected Protein Threshold [Unused ProtScore (Conf)] >1.30 (95.0%).

172

The protein-level quantitative dataset of 1087 proteins was produced by removing

173

proteins with only one unique peptides and those not quantified in all of the three

174

replicate experiments. A protein defined as differentially expressed protein (DEP)

175

should match with the criteria: the fold change had to be greater than 1.50 or less than

176

0.67 with p < 0.05 in SNH treated 1 h or 2 h from biological triplicate experiments.

177

Volcano plot and protein hierarchical analysis were performed using Matlab.

178

Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways enrichment

179

analysis was performed on the lists of DEPs to identify over-represented biological

180

pathways using Blast2GO program, and visualized with Cytoscape.

181

9. Real-Time quantitative PCR (RT-qPCR)

182

RNA was prepared from S. pneumoniae D39 treated with 1/4 × MIC (6.25 µM) of

183

SNH for 0, 1 h and 2 h. Total RNA was obtained using TRIZOL method according to

184

previous protocol

185

value of gyrB amplified from the corresponding sample and calculated using the

186

2−∆∆Ct method 25. The primer sequences are listed in Table S1.

187

10. Concentration measurement of hydrogen peroxide (H2O2)

188

To measure H2O2 levels, bacteria were removed by centrifugation at 8,000 g for 5 min

189

at 4 °C, then supernatants were filtered through a 0.22 µm filter and used for analysis

190

in accordance with the Amplex Red hydrogen peroxide/peroxidase assay kit

24

. The fold changes of selected genes were normalized to the Ct

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 32

191

(Invitrogen, USA). The final value was normalized to give µM/106 CFU.

192

11. Statistical analysis

193

Data were analyzed using two-tailed unpaired Student t-test, and expressed as mean ±

194

SEM. Statistical analysis was conducted using GraphPad Prism 5.0. For all

195

comparisons, p < 0.05 was considered as significant.

196

Results

197

1. Bactericidal activity of SNH against S. pneumoniae

198

The bacteriostatic and bactericidal activities of SNH against S. pneumoniae were

199

investigated using MIC and MBC. We measured that the MIC value of SNH against S.

200

pneumoniae is 25 µM, and MBC value also is 25 µM. According to previous reports 26,

201

the bactericidal activity was defined as the MBC value ≤ 4 × MIC, the value (both

202

MIC and MBC equals 25 µM) indicated that SNH is bactericidal. The time-kill curves

203

of SNH and ampicillin for S. pneumoniae are displayed in Figure 1; ampicillin was

204

used as control in this study as it is well-known broad-spectrum bactericidal that

205

inhibits cell wall synthesis. The result confirmed the bactericidal nature of SNH, as 1

206

× MIC, or mostly 2 × MIC was sufficient to kill 99.99% bacteria with 1 h of

207

incubation. In contrast, the time for 1 × MIC or 2 × MIC of ampicillin to kill 99.99%

208

bacteria is 6 h. These results indicated that the bactericidal rate of SNH is faster than

209

ampicillin.

210

2. No toxicity of SNH to host cells

10

ACS Paragon Plus Environment

Page 11 of 32

Journal of Agricultural and Food Chemistry

211

While good bactericidal activity is basic for a potential antimicrobics, SNH must also

212

display low or no toxicity towards host cells. Hence, the cytotoxicity of SNH against

213

human bronchial epithelial cell line HBE was determined by LDH assay. As shown in

214

Figure 2B, SNH at 200 µM (corresponding to 8 × MIC of SNH against S. pneumoniae)

215

exhibited no toxicity towards HBE cells over a 24 h period, suggesting that SNH is

216

selectively toxic against S. pneumoniae.

217

3. iTRAQ-based proteomic analysis on S. pneumoniae in response to SNH

218

In order to investigate the bactericidal mechanism of SNH against S. pneumoniae,

219

proteomic analyses were conducted using iTRAQ-based LC-MS/MS approach to

220

obtain large scale data sets in three biological replicates. According to the time-kill

221

curves (Fig. 1A) and growth curves (Fig. 2A), 1/4 × MIC doses of SNH were applied

222

to treat S. pneumoniae, with a consideration to minimize bacterial cell death caused by

223

SNH and thus to observe real changes in the abundance of proteins. Bacteria were

224

harvested at 0, 1 h and 2 h after treatment with SNH (6.25 µM) so that they had

225

sufficient time to produce proteins to counteract the inflicted damage while

226

minimizing the secondary unspecific effects.

227

A total of 1087 proteins (Table S2) had quantitative values available from

228

triplicate biological experiments and were used for downstream analyses. The volcano

229

plot for the 1087 proteins after SNH-treated 1 h and 2 h are shown in Figures 3A&B.

230

The proteins with the fold changes greater than 1.50 or less than 0.67 with p < 0.05

231

were considered as DEPs. As shown in Figure 3C, 135 proteins (100 up-regulations

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 32

232

and 35 down-regulations) were found significantly changed in the samples with

233

SNH-1 h treatment, while more DEPs (159 proteins including 82 up-regulations and

234

77 down-regulations) were identified in the samples with SNH-2 h treatment. Only 75

235

DEPs were overlapped in the two groups, with 60 and 84 proteins uniquely and

236

differentially expressed in SNH-1 h or -2 h treatments, respectively (Fig. 3D). In order

237

to comprehensively analyze the protein changes in response to the SNH incubation,

238

all the 219 proteins (Table S3) that belong to the union set of DEPs in both treatment

239

groups were subjected to bioinformatics analysis.

240

4. Classification of the identified DEPs

241

To gain insights into the biological pathways of the 219 DEPs, hierarchical clustering

242

of these proteins was performed using Matlab to group the data into those proteins

243

increasing in abundance following SNH treatment and those decreasing in abundance

244

(Fig. 4A). KEGG pathways enrichment analysis indicated that the proteins generally

245

down-regulated in SNH treatment are enriched in pathways associated with ribosome,

246

fatty acid biosynthesis and RNA polymerase, while the up-regulated proteins are

247

enriched for propanoate metabolism, oxidative phosphorylation, other glycan

248

degradation, bacterial secretion system, pyruvate metabolism and pyrimidine

249

metabolism.

250

As shown in Figure 4, 27 ribosomal proteins responsible for the protein

251

biosynthesis (Fig. 4B), three RNA polymerases (RpoB, RpoC, RpoZ) (Fig. 4D), and

252

eight proteins (AccB, AccC, AccD, FabD, FabF, FabG, FabH and FabK) involved in

12

ACS Paragon Plus Environment

Page 13 of 32

Journal of Agricultural and Food Chemistry

253

fatty acid biosynthesis (Fig. 4E) decreased their expression following SNH treatment.

254

These data suggest that RNA, protein and fatty acid biosynthesis were impaired in the

255

bacterium after being exposed to SNH. Two possibilities may lead to this result. One

256

is that SNH may directly inhibit the RNA, protein and fatty acid biosynthesis, and the

257

other is that SNH may stimulate the production of ROS and indirectly interfere the

258

RNA, protein and fatty acid biosynthesis, then eventually induced cell death.

259

Moreover, seven proteins involved in pyruvate metabolism were up-regulated

260

after SNH treatment (Fig. 4E), which is compatible with the second possibility. In

261

particular, pyruvate oxidase (SpxB), an enzyme converting pyruvate to hydrogen

262

peroxide (H2O2) 27, was up-regulated by 2.56 folds and 3.52 folds in SNH-1 h and -2 h

263

treatment, respectively. The up-regulation of spxB in mRNA level was further

264

validated by RT-qPCR analysis (Fig. 5A). Also, non-heme iron-containing ferritin

265

(SPD_1402, homologous with Dpr) involved in hydrogen peroxide stress defense

266

(Table S3 & Fig. 5A) and phosphocarrier protein HPr (PstH) that transports phosphate

267

were increased after SNH treatments. In addition, we observed the significant

268

up-regulation of chaperonin proteins involved in the regulation of misfolded proteins

269

(DnaJ, GroS, GrpE) (Table S3). All these observations suggest that ROS production

270

was stimulated by SNH to induce cell death.

271

5. H2O2 formation induced by SNH contributing to cell death

272

To support the hypothesis that SNH may simulate the production of high ROS and

273

indirectly interfere the RNA, protein and fatty acid biosynthesis, then eventually lead

28

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 32

274

to cell death, we measured the levels of H2O2 in the supernatants from various SNH

275

treatments (corresponding to 0, 1/8 × MIC, 1/4 × MIC, 1/2 × MIC and 1 × MIC). As

276

expected, we observed cellular death with 1/4 × MIC, 1/2 × MIC and 1 × MIC SNH

277

(Fig. 1A), accompanied with the increase of H2O2 production (Fig. 5B). Furthermore,

278

the concentration of H2O2 in supernatant at 2 h incubation with SNH was more than

279

that with 1 h incubation. These results indicate that SNH induced H2O2 formation in

280

time- and concentration-dependent manners, which are in accord with the time-kill

281

curves.

282

To further explore the biological significance of H2O2 contribution to cell death,

283

we exploited catalase that can neutralize H2O2 to the experiment. We found that S.

284

pneumoniae treated simultaneously with SNH and catalase (2 mg/mL) showed a

285

significant increase in cell survival, near to the survival level of no SNH treatments

286

(Fig. 5C). This increase in bacterial survival suggests that catalase almost eliminated

287

the SNH bactericidal effect, indicating that H2O2 was indeed an important mediator in

288

the bactericidal action of SNH. These results provided evidence that H2O2 plays a

289

significant role in SNH-induced cell death.

290

Discussion

291

SNH inhibits a variety of bacteria, yet its anti-bacterial mechanism remains unknown.

292

In this study, we investigated the bactericidal activity and mechanism of SNH using S.

293

pneumoniae as a model. The MIC and MBC assays combined with time-kill curves

294

showed that SNH killed S. pneumoniae in a dose-dependent manner. We then

14

ACS Paragon Plus Environment

Page 15 of 32

Journal of Agricultural and Food Chemistry

295

employed iTRAQ-based quantitative proteomics analysis to demonstrate that SNH

296

stimulated the expression of the proteins involved in H2O2 formation pathway (Fig. 6),

297

including pyruvate metabolism related proteins (Pta, SpxB, Ldh, PflB) and

298

phosphocarrier protein PstH. Meanwhile, proteins involved in protective responses to

299

ROS, such as non-heme iron-containing ferritin (SPD_1402) and chaperone systems

300

proteins DnaJ, GroS and GrpE, were activated by SNH treatment. On the other hand,

301

SNH suppressed 27 ribosomal proteins responsible for the protein biosynthesis, three

302

RNA polymerases (RpoB, RpoC, RpoZ) involved in RNA biosynthesis and eight

303

proteins (AccB, AccC, AccD, FabD, FabF, FabG, FabH and FabK) involved in fatty

304

acid biosynthesis (Fig. 6).

305

Importantly, the content of H2O2 was greatly increased in the culture supernatant

306

of the bacterium exposed to SNH, while the H2O2 neutralizing enzyme catalase can

307

greatly attenuate the killing by SNH (Fig. 5). Based on these observations, we

308

concluded that SNH induced the formation of H2O2 via SpxB participation in pyruvate

309

metabolism, then elevated H2O2 to result in the damage to RNA, protein and cell

310

membrane lipids and ultimately contributed to the cell death of bacteria (Fig. 6).

311

Interestingly, previous in vitro work has shown that SNH demonstrated

312

antibacterial activity against 103 clinical strains of MRSA, functioning in synergetic

313

effects with oxacillin or netilmicin 15. Here, we also detected the antibacterial activity

314

of SNH against S. aureus by measuring the level of H2O2 in the culture supernatant

315

after SNH treatment. As shown in Figure S1, SNH stimulated the H2O2 production

316

and catalase also greatly decreased the inhibition effect of SNH on S. aureus, a result 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 32

317

in agreement with the observation in S. pneumoniae. In addition, SNH analogue SH

318

showed the antibacterial activity against biofilm formation and alginate production of

319

P. aeruginosa 29, echoing the report that ROS production contributes to the killing of

320

biofilm subpopulations in P. aeruginosa exposed to ciprofloxacin

321

results, it seems reasonable to conclude that SNH utilizes a common mechanism of

322

bacteria-killing by inducing the production of ROS responsible for cell lethality.

30

. With these

323

It is somewhat surprising that the bactericidal rate of SNH is better than

324

ampicillin. As shown in time-kill curves (Fig. 1), 1-2 × MIC SNH killed almost all the

325

bacteria in 1 hour, while the equivalent concentration of ampicillin needed 6 hours to

326

kill the bacteria. Bacterial reproduction is very rapid, the generation of most bacteria

327

is about 20-30 min; the shorter time of killing means more effective for the drug to

328

control the infection. Interestingly, the bacteria could recover from the sixth hour post

329

the application of 1/2 × MIC SNH; a possible reason is that some bacteria could not

330

be eliminated by the low concentration of SNH and then became resistant to the drug.

331

However, this phenomenon was not observed during ampicillin application,

332

implicating that the antibacterial mechanisms between SNH and ampicillin may be

333

different. These results provide helpful information for guiding the use of an

334

appropriate concentration of SNH in disease treatment, and also suggest that SNH

335

may be combined with existing antibiotics as an effective strategy to combat bacteria.

336

Bactericidal drug design has generally fallen into three groups: inhibition of DNA

337

replication, protein synthesis and cell-wall turnover, which has yielded great progress

338

in antimicrobial therapy 31. However, the rise of antibiotics resistant bacteria has made 16

ACS Paragon Plus Environment

Page 17 of 32

Journal of Agricultural and Food Chemistry

339

it critical for us to develop novel and more effective strategies to kill bacteria. In

340

recent years, the formation of ROS represents one of the major mechanisms to

341

generate cell death. Our results indicated that SNH possesses strong bactericidal

342

activity against S. pneumoniae based on the generation of H2O2. Moreover,

343

proteomics coupled with bioinformatics analysis may reveal drug targets for

344

stimulating ROS formation and thus open-up new avenues for developing new classes

345

of bactericidal drugs.

346

347

ASSOCIATED CONTENT

348

Supporting Information

349

Supporting Information Available: Figure S1, Tables S1, S2 and S3. This material is

350

available free of charge via the Internet at http://pubs.acs.org.

351 352

Funding

353

This work was supported by the National Natural Science Foundation of China

354

(21271086, to Q.-Y. H.), China Postdoctoral Science Foundation (153150, to X.-Y. Y.),

355

and the Open Fund of the First Affiliated Hospital, Jinan University.

356

Notes

357

The authors declare no competing financial interest.

358

359

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 32

360

361 362

18

ACS Paragon Plus Environment

Page 19 of 32

Journal of Agricultural and Food Chemistry

363

References

364

1.

365

reactive oxygen species: an unusual approach to living with radicals. Trends Microbiol 2013, 21,

366

187-195.

367

2.

368

pneumonia. Lancet 2009, 374, 1543-1556.

369

3.

Imlay, J. A., Pathways of oxidative damage. Annu Rev Microbiol 2003, 57, 395-418.

370

4.

Kohanski, M. A.; Dwyer, D. J.; Hayete, B.; Lawrence, C. A.; Collins, J. J., A common

371

mechanism of cellular death induced by bactericidal antibiotics. Cell 2007, 130, 797-810.

372

5.

373

Agents Chemother 2009, 53, 1395-1402.

374

6.

375

D. J.; Khalil, A. S.; Collins, J. J., Antibiotic efficacy is linked to bacterial cellular respiration. Proc

376

Natl Acad Sci U S A 2015, 112, 8173-8180.

377

7.

378

Activity of Quinolone Antibiotics. Angew Chem Int Ed Engl 2016, 55, 2397-2400.

379

8.

380

Bojang, K.; Olaosebikan, R.; Anunobi, N.; Maitland, K.; Kivaya, E.; Agbenyega, T.; Nguah, S. B.;

381

Evans, J.; Gesase, S.; Kahabuka, C.; Mtove, G.; Nadjm, B.; Deen, J.; Mwanga-Amumpaire, J.;

382

Nansumba, M.; Karema, C.; Umulisa, N.; Uwimana, A.; Mokuolu, O. A.; Adedoyin, O. T.;

Yesilkaya, H.; Andisi, V. F.; Andrew, P. W.; Bijlsma, J. J., Streptococcus pneumoniae and

van der Poll, T.; Opal, S. M., Pathogenesis, treatment, and prevention of pneumococcal

Wang, X.; Zhao, X., Contribution of oxidative damage to antimicrobial lethality. Antimicrob

Lobritz, M. A.; Belenky, P.; Porter, C. B.; Gutierrez, A.; Yang, J. H.; Schwarz, E. G.; Dwyer,

Kottur, J.; Nair, D. T., Reactive Oxygen Species Play an Important Role in the Bactericidal

Dondorp, A. M.; Fanello, C. I.; Hendriksen, I. C.; Gomes, E.; Seni, A.; Chhaganlal, K. D.;

19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 32

383

Johnson, W. B.; Tshefu, A. K.; Onyamboko, M. A.; Sakulthaew, T.; Ngum, W. P.; Silamut, K.;

384

Stepniewska, K.; Woodrow, C. J.; Bethell, D.; Wills, B.; Oneko, M.; Peto, T. E.; von Seidlein, L.;

385

Day, N. P.; White, N. J.; group, A., Artesunate versus quinine in the treatment of severe falciparum

386

malaria in African children (AQUAMAT): an open-label, randomised trial. Lancet 2010, 376,

387

1647-1657.

388

9.

389

deoxyandrographolide and neoandrographolide in the Chinese herb Andrographis paniculata by

390

micellar electrokinetic capillary chromatography. Phytochem Anal 2002, 13, 222-227.

391

10. Peng, L.; Kang, S.; Yin, Z.; Jia, R.; Song, X.; Li, L.; Li, Z.; Zou, Y.; Liang, X.; Li, L.; He, C.;

392

Ye, G.; Yin, L.; Shi, F.; Lv, C.; Jing, B., Antibacterial activity and mechanism of berberine against

393

Streptococcus agalactiae. Int J Clin Exp Pathol 2015, 8, 5217-5223.

394

11. Miao, J.; Chen, F.; Duan, S.; Gao, X.; Liu, G.; Chen, Y.; Dixon, W.; Xiao, H.; Cao, Y.,

395

iTRAQ-Based Quantitative Proteomic Analysis of the Antimicrobial Mechanism of Peptide F1

396

against Escherichia coli. J Argic Food Chem 2015, 63, 7190-7197.

397

12. Lu, H. M.; Liang, Y. Z.; Yi, L. Z.; Wu, X. J., Anti-inflammatory effect of Houttuynia cordata

398

injection. J Ethnopharmacol 2006, 104, 245-249.

399

13. Hayashi, K.; Kamiya, M.; Hayashi, T., Virucidal effects of the steam distillate from

400

Houttuynia cordata and its components on HSV-1, influenza virus, and HIV. Planta Med 1995, 61,

401

237-241.

402

14. Lu, H.; Wu, X.; Liang, Y.; Zhang, J., Variation in chemical composition and antibacterial

403

activities of essential oils from two species of Houttuynia THUNB. Chem Pharm Bull 2006, 54,

Zhao, J.; Yang, G.; Liu, H.; Wang, D.; Song, X.; Chen, Y., Determination of andrographolide,

20

ACS Paragon Plus Environment

Page 21 of 32

Journal of Agricultural and Food Chemistry

404

936-940.

405

15. Lu, X.; Yang, X.; Li, X.; Lu, Y.; Ren, Z.; Zhao, L.; Hu, X.; Jiang, J.; You, X., In vitro activity

406

of sodium new houttuyfonate alone and in combination with oxacillin or netilmicin against

407

methicillin-resistant Staphylococcus aureus. PloS one 2013, 8, e68053.

408

16. Wang, D.; Yu, Q.; Eikstadt, P.; Hammond, D.; Feng, Y.; Chen, N., Studies on adjuvanticity of

409

sodium houttuyfonate and its mechanism. Int Immunopharmacol 2002, 2, 1411-1418.

410

17. Liang, L., Observation of sodium new houttuyfonat injection in the treatment of children

411

acute upper respiratory infection. Shenzhen Integra Tradit Chin West Med 2006, 16, 36-38.

412

18. Ye, X. L.; Li, X. G.; Yuan, L. J.; He, H. M., Relationship between the antibacterial and

413

immunological activities of houttuyfonate homologues and their surface activities. J Asian Nat

414

Prod Res 2006, 8, 327-334.

415

19. Wu, D.; Huang, W.; Duan, Q.; Li, F.; Cheng, H., Sodium houttuyfonate affects production of

416

N-acyl homoserine lactone and quorum sensing-regulated genes expression in Pseudomonas

417

aeruginosa. Front Microbiol 2014, 5, 635 (1-9).

418

20. Motyl M, D. K., Barrett J et al., Current Protocols in Pharmacology. New York: John Wiley &

419

Sons 2005, 13A, 1-22.

420

21. Wiegand, I.; Hilpert, K.; Hancock, R. E., Agar and broth dilution methods to determine the

421

minimal inhibitory concentration (MIC) of antimicrobial substances. Nat Protoc 2008, 3, 163-175.

422

22. Clinical and Laboratory Standards Institute, Performance Standards for Antimicrobial

423

Susceptibility Testing: Twenty-second Informational Supplement M100-S22. CLSI, Wayne, PA,

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 32

424

USA. 2012.

425

23. Yin, S.; Xue, J.; Sun, H.; Wen, B.; Wang, Q.; Perkins, G.; Zhao, H. W.; Ellisman, M. H.;

426

Hsiao, Y. H.; Yin, L.; Xie, Y.; Hou, G.; Zi, J.; Lin, L.; Haddad, G. G.; Zhou, D.; Liu, S.,

427

Quantitative evaluation of the mitochondrial proteomes of Drosophila melanogaster adapted to

428

extreme oxygen conditions. PloS one 2013, 8, e74011.

429

24. Yang, X. Y.; Sun, B.; Zhang, L.; Li, N.; Han, J.; Zhang, J.; Sun, X.; He, Q. Y., Chemical

430

interference with iron transport systems to suppress bacterial growth of Streptococcus pneumoniae.

431

PloS one 2014, 9, e105953.

432

25. Livak, K. J.; Schmittgen, T. D., Analysis of relative gene expression data using real-time

433

quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 2001, 25, 402-408.

434

26. Pankey, G. A.; Sabath, L. D., Clinical relevance of bacteriostatic versus bactericidal

435

mechanisms of action in the treatment of Gram-positive bacterial infections. Clin Infect Dis 2004,

436

38, 864-870.

437

27. Spellerberg, B.; Cundell, D. R.; Sandros, J.; Pearce, B. J.; Idanpaan-Heikkila, I.; Rosenow, C.;

438

Masure, H. R., Pyruvate oxidase, as a determinant of virulence in Streptococcus pneumoniae. Mol

439

Microbiol 1996, 19, 803-813.

440

28. Tsou, C. C.; Chiang-Ni, C.; Lin, Y. S.; Chuang, W. J.; Lin, M. T.; Liu, C. C.; Wu, J. J., An

441

iron-binding protein, Dpr, decreases hydrogen peroxide stress and protects Streptococcus pyogenes

442

against multiple stresses. Infect Immun 2008, 76, 4038-4045.

443

29. Wu, D. Q.; Cheng, H.; Duan, Q.; Huang, W., Sodium houttuyfonate inhibits biofilm

22

ACS Paragon Plus Environment

Page 23 of 32

Journal of Agricultural and Food Chemistry

444

formation and alginate biosynthesis-associated gene expression in a clinical strain of. Exp Ther

445

Med 2015, 10, 753-758.

446

30. Jensen, P. O.; Briales, A.; Brochmann, R. P.; Wang, H.; Kragh, K. N.; Kolpen, M.; Hempel,

447

C.; Bjarnsholt, T.; Hoiby, N.; Ciofu, O., Formation of hydroxyl radicals contributes to the

448

bactericidal activity of ciprofloxacin against Pseudomonas aeruginosa biofilms. Pathog Dis 2014,

449

70, 440-443.

450

31. Walsh, C., Molecular mechanisms that confer antibacterial drug resistance. Nature 2000, 406,

451

775-781.

452

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 32

453

Figure captions

454

Figure 1. Time-kill curves of SNH (A) and ampicillin (B) against S. pneumoniae.

455

Figure 2. Effects of SNH on S. pneumoniae and human bronchial epithelial cell line

456

HBE. (A) SNH inhibited the growth of S. pneumoniae in a dose-dependent manner.

457

The growth curve measurements were taken every hour for the first 4 h and every two

458

hours for next 8 h after drug addition at OD600. (B) SNH exhibited low toxicity

459

towards host cells. HBE cells were treated with the various concentrations of SNH,

460

and the cytotoxicity was determined using LDH Cytotoxicity Assay Kit. The data

461

shown represent the mean of three experiments; error bars indicate SEM.

462

Figure 3. Proteomic changes in S. pneumoniae in response to the treatment with SNH.

463

(A) Volcano plots of DEPs in SNH treatment for 1 h. (B) Volcano plots of DEPs in

464

SNH treatment for 2 h. (C) Numbers of DEPs in the SNH-1 h and -2 h treatments. (D)

465

Venn diagram of the number of DEPs in the SNH-1 h and -2 h treatments,

466

respectively.

467

Figure 4. Classification of 219 proteins that belong to the union set of DEPs in both

468

SNH-1 h and -2 h treatments. (A) Hierarchical cluster analysis was conducted using

469

Matlab for the 219 proteins. The color indicates relative fold changes (red =

470

up-regulated, blue = down-regulated). DEPs were then subjected to KEGG pathways

471

analysis using Blast2GO program, significantly enriched pathways (p < 0.05) were

472

summarized in the tables shown. (B-G) The pathways enriched using Blast2GO,

473

visualized with Cytoscape. Red represents the proteins up-regulated, blue represents 24

ACS Paragon Plus Environment

Page 25 of 32

Journal of Agricultural and Food Chemistry

474

the proteins down-regulated after SNH treatments.

475

Figure 5. SNH induced H2O2 formation that contributed to cell death. (A) RT-qPCR

476

analysis of selected genes, spxB, atpA, spd_1402, fabK and fabG in S. pneumoniae

477

with SNH treatments for 1 and 2 h versus 0 h (control). The relative gene expression

478

was calculated with gyrB as the reference gene. All results represent the relative

479

expression level with SNH treatments for 1 and 2 h versus 0 h (control) (**p < 0.01,

480

***p < 0.001), shown as the means value (± SEM) from three independent

481

experiments. (B) H2O2 production quantified in the culture supernatants at 0, 1 and 2

482

h treatments with different concentrations of SNH (0, 1/8 × MIC, 1/4 × MIC, 1/2 ×

483

MIC and 1 × MIC). Results show mean ± SEM for three experiments. **p < 0.01,

484

***p < 0.001, unpaired Student's t-test. (C) Log change in CFU/mL following

485

exposure to no SNH as the control (circle), 1 × MIC SNH (cross) and 1 × MIC SNH

486

with 2 mg/mL catalase (diamond). Results show mean ± SEM for three experiments.

487

Figure 6. Proposed mechanism of SNH-induced cell death in S. pneumoniae. SNH

488

induced the formation of H2O2 via SpxB participation in pyruvate metabolism, then

489

elevated H2O2 to result in the damage to RNA, protein and cell membrane lipids and

490

finally contributed to cell death.

491 492

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 32

493

Figure 1

26

ACS Paragon Plus Environment

Page 27 of 32

Journal of Agricultural and Food Chemistry

Figure 2

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 32

Figure 3

28

ACS Paragon Plus Environment

Page 29 of 32

Journal of Agricultural and Food Chemistry

Figure 4

.

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 32

Figure 5

30

ACS Paragon Plus Environment

Page 31 of 32

Journal of Agricultural and Food Chemistry

Figure 6

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 32

TOC Graphic

32

ACS Paragon Plus Environment