Chronic Exposure to an Environmentally Relevant Triclosan

Feb 21, 2019 - Chronic Exposure to an Environmentally Relevant Triclosan Concentration Induces Persistent Triclosan Resistance but Reversible Antibiot...
0 downloads 0 Views 929KB Size
Subscriber access provided by MIDWESTERN UNIVERSITY

Ecotoxicology and Human Environmental Health

Chronic Exposure to an Environmentally Relevant Triclosan Concentration Induces Persistent Triclosan Resistance but Reversible Antibiotic Tolerance in Escherichia coli Mingzhu Li, Yuning He, Jing Sun, Jing Li, Junhong Bai, and Chengdong Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06763 • Publication Date (Web): 21 Feb 2019 Downloaded from http://pubs.acs.org on February 21, 2019

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 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 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.

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 31

Environmental Science & Technology

1

Chronic Exposure to an Environmentally Relevant Triclosan Concentration

2

Induces Persistent Triclosan Resistance but Reversible Antibiotic Tolerance

3

in Escherichia coli

4

Mingzhu Li, 1 Yuning He, 1 Jing Sun, 1 Jing Li, 1 Junhong Bai,2 Chengdong Zhang 2, *

5 6

1

College of Environmental Science and Engineering, Nankai University, Tianjin 300350, China

7 8

2

School of Environment, Beijing Normal University, Beijing 100875, China

9 10 11

Manuscript prepared for Environmental Science & Technology

12 13 14 15 16 17

* Corresponding author: (Phone/fax) 86-10-58802029.

18

E-mail address: [email protected].

19 20 21 22 23

Postal address: School of Environment Beijing Normal University No. 19, XinJieKouWai St., HaiDian District, Beijing, P. R. China 100875

1 ACS Paragon Plus Environment

Environmental Science & Technology

24

TOC of Art

25 26

2 ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

Environmental Science & Technology

27 28

ABSTRACT The major concern regarding the biocide triclosan (TCS) stems from its potential

29

coselection for antibiotic resistance. However, environmental impacts are often

30

investigated using high concentrations and acute exposure, while predicted releases are

31

typified by chronic low concentrations. Moreover, little information is available

32

regarding the reversibility of TCS and derived antibiotic resistance with diminishing TCS

33

usage. Here, the model Gram-negative bacterium Escherichia coli was exposed to 0.01

34

mg/L TCS continuously for more than 100 generations. The adapted cells gained

35

considerable resistance to TCS as indicated by a significant increase in the minimal

36

inhibitory concentration (MIC50) from 0.034 to 0.581 mg/L. This adaptive evolution was

37

attributed to overexpression and mutation of target genes (i.e., fabI) as evidenced by

38

transcriptomic and genomic analyses. However, only mild tolerance to various antibiotics

39

was observed, possibly due to reduced membrane permeability and biofilm formation.

40

After TCS exposure ceased, the adapted cells showed persistent resistance to TCS due to

41

inheritable genetic mutations, whereas their antibiotic tolerance declined over time. Our

42

results suggest that extensive use of TCS may promote the evolution and persistence of

43

TCS-resistant bacterial pathogens. Quantitative definition of the conditions under which

44

TCS selects for multidrug resistance in the environment is crucially needed.

3 ACS Paragon Plus Environment

Environmental Science & Technology

45 46

INTRODUCTION Triclosan [TCS; 5-chloro-2-(2,4-dichlorophenoxy)phenol] is widely used in

47

personal care products (e.g., toothpaste, soap and body wash) as an antimicrobial agent 1.

48

TCS is used in large quantities on a global scale, resulting in high residual levels in water.

49

For example, its concentration ranges from 2 × 10-7 to 4.78 × 10-4 mg/L in surface water,

50

and it can be detected in amounts as high as 8.62 × 10-2 mg/L and 5.37 × 10-3 mg/L,

51

respectively, in wastewater influent and effluent 2. In particular, most TCS preferentially

52

adsorbs onto carbon and lipid-rich sludge 3, 4. Therefore, the levels of TCS in digested

53

sewage sludge were reported to be as high as 133 mg/kg dry weight, with a mean

54

concentration of approximately 16 ± 65 mg/kg dry weight (± standard deviation (SD)) 1, 5.

55

Given the massive amount of TCS released into the environment, the

56

microorganisms living in contaminated environments may have evolved certain

57

mechanisms to tolerate/resist this biocide. For example, pathogenic bacteria, such as

58

Salmonella enterica 6, 7, Escherichia coli 8, 9 and Pseudomonas aeruginosa 10, 11, are

59

currently less susceptible to TCS than they were in the past. The mechanisms of TCS

60

resistance include target gene mutation 12, target gene overexpression 13, induction of

61

efflux pumps 14, reduced membrane permeability 15, and TCS transformation or

62

degradation 16. However, the typical concentrations of TCS used for these evolutionary

63

studies have ranged from 0.195 mg/L to 2.05 ×103 mg/L 17-20, which are much higher

64

than realistic environmental concentrations. The incubation times studied have generally

65

varied between 24 h and 20 days, which is a short period of time compared to years of

66

exposure in the environment. Moreover, beginning in September 2017, the FDA

67

prohibited the sale of "consumer antiseptic washes" containing TCS. Information 4 ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Environmental Science & Technology

68

regarding whether TCS resistance would decrease or disappear in response to suddenly

69

reduced concentrations is lacking.

70

TCS and its derivatives are considered a group of endocrine disruptors 1, 5, and an

71

emerging toxic outcome of concern is the potential link between TCS exposure and

72

coselection on antibiotic resistance. For instance, at a concentration of 0.2 mg/L, TCS

73

induces multidrug resistance in wild-type E. coli after 30 days of exposure 21. In this

74

study, TCS exposure-induced mutants demonstrated greater than 6-fold increases in the

75

MIC90 (the minimum inhibitory concentration that kills 90% of bacteria) against three

76

beta-lactam antibiotics, including ampicillin, cephalexin and amoxicillin. The

77

mechanisms that confer cross-resistance to antibiotics are induction of a multidrug efflux

78

pump and decreased influx or membrane permeability 16. Nevertheless, whether TCS can

79

directly trigger antibiotic resistance at environmentally relevant concentrations remains

80

under debate. Determining whether the antibiotic resistance associated with TCS

81

resistance will be reversible after the sudden decline in consumer usage is equally

82

important.

83

The goals of this project were (1) to determine whether prolonged exposure of the

84

model bacterium E. coli to an environmentally relevant concentration of TCS can trigger

85

TCS resistance and examine the associated mechanism and its hereditary stability, and

86

(2) to experimentally investigate cross-resistance/tolerance to antibiotics and the

87

reversibility of such cross-resistance/tolerance. The results indicated that chronic TCS

88

exposure may be linked to a significant increase in TCS resistance accompanied by a

89

moderate rise in drug tolerance. Once the stressor was removed, TCS resistance lasted

90

more than 20 generations, whereas stress-induced antibiotic tolerance eventually

5 ACS Paragon Plus Environment

Environmental Science & Technology

91

disappeared. Our study reveals that implementing environmental emission limits for TCS

92

would be an effective strategy to prevent preferential selection for antimicrobial

93

resistance.

6 ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

Environmental Science & Technology

94 95

MATERIALS AND METHODS Strains and the evolution experiment. TCS (purity ≥ 99%) and dimethyl

96

sulfoxide (DMSO) were purchased from Sigma-Aldrich, USA. TCS stock solution (100

97

mg/L) was prepared in DMSO. To prepare 0.01-mg/L TCS solution, 5 μL of stock

98

solution was added to 50 mL of Luria-Bertani (LB) medium 22, 23. Silicide glassware was

99

used throughout the experiment to prevent TCS adsorption. Escherichia coli K-12 strain

100

MP1 was obtained from the Agricultural Culture Collection of China (ACCC, Beijing).

101

E. coli was incubated in LB medium containing 0.01 mg/L TCS at 37°C under

102

continuous shaking (160 rpm). Every 24 h, 1 mL of bacterial suspension was transferred

103

to fresh medium containing the same amount of TCS. The process was repeated for 100

104

subculture cycles, and the TCS-adapted strain was termed ETCS (adapted cells). The

105

selected concentration of 0.01 mg/L in this study represents various degrees of

106

contamination in receiving waters, e.g., hotspots in surface water 21, the medians in

107

wastewater influent 24 and high residual levels in effluent 25.

108

Bacterial growth and MIC50 determination. Cells (both wild-type E. coli and

109

ETCS) were collected via centrifugation at 8000 ×g for 10 min at 4°C, washed with saline

110

(0.9% NaCl) three times (to remove residual TCS and metabolites), and then resuspended

111

in saline to achieve a concentration of 107-108 CFU/mL. One milliliter of cell suspension

112

was inoculated into 49 mL of sterilized LB medium in the presence or absence of 0.02

113

mg/L TCS, and the optical density was measured at 600 nm (OD600) with a UV-Vis

114

spectrophotometer (Cary 100 UV-Vis, Agilent Technologies, USA) over 12 h. The MIC50

115

was determined as the concentration of TCS that inhibited 50% of bacterial growth after

116

12 h of incubation.

7 ACS Paragon Plus Environment

Environmental Science & Technology

Ribonucleic acid (RNA) preparation, library construction and quantitative

117 118

real-time polymerase chain reaction (qRT-PCR) validation. Detailed descriptions of

119

RNA isolation, RNA sequencing, reverse transcription, and qRT-PCR can be found in the

120

Supporting Information (SI). The primers used for PCR were designed by Sangon

121

Biotech (Shanghai, China) and are listed in Table S1. RNA-sequencing data analysis. Clean reads were mapped to the Escherichia coli

122 123

K-12 MG1655 genome assembly using TopHat software 26. Fragments per kilobase of

124

exon model per million mapped reads (FPKMs) values for each gene and differentially

125

expressed genes (DEGs) were analyzed with Cufflinks v2.2.1. The DEGs between the

126

two samples were identified by considering both fold changes (log2FC > 1) and p-values

127

(p < 0.005). Hierarchical clustering analysis was performed using the FPKMs of DEGs of

128

E. coli, E. coli + TCS (wild-type cells exposed to 0.02 mg/L TCS) and ETCS + TCS

129

(adapted cells exposed to 0.02 mg/L TCS).

130

Deoxyribonucleic acid (DNA) extraction, whole-genome sequencing and data

131

processing. Genomic DNA was extracted using the FastDNATM SPIN Kit for Soil (MP,

132

USA). The Nextera XT DNA Sample Preparation Kit (Illumina, USA) was used to

133

prepare a whole-genome sequencing library, which was sequenced by Allwegene

134

BioTech Co. using a MiSeq instrument (Illumina). Two biological replicates were

135

performed in each group. Escherichia coli K-12 MG1655 (NCBI reference sequence

136

NC000913.3) was used as the reference genome sequence. Further details are given in the

137

SI.

138 139

Determination of the fatty acid composition of bacterial membranes. Fatty acids in bacterial membranes were extracted using the method reported by Yang et al. 27

8 ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Environmental Science & Technology

140

with some modifications. Fatty acid methyl ester was measured by gas chromatography

141

(7890A, Agilent Technologies) analysis on a device equipped with a flame ionization

142

detector and a capillary column HP-5 (30 m, 0.32 mm, 0.25 μm, Agilent Technologies);

143

see the SI for detailed descriptions. All treatments were performed in triplicate, and the

144

data are displayed as the means ± SD.

145

Biofilm characterization by confocal laser scanning microscopy. Cells were

146

grown statically in the presence of 0.02 mg/L TCS for 24 h. Biofilms on the bottoms of

147

culture dishes were stained with the LIVE/DEAD BacLight Bacterial Viability Kit

148

(L13152, ThermoFisher Scientific Inc.) 28 and were visualized by confocal laser scanning

149

microscopy (Zeiss, LSM880 with Airyscan, Germany). See the SI for details.

150

Antibiotic sensitivity tests. The MIC50 values to various antibiotics (i.e., penicillin,

151

kanamycin, gentamicin and ciprofloxacin) were determined as the concentration of the

152

corresponding antibiotic that inhibited 50% of bacterial growth (wild-type E. coli.).

153

Antibiotic sensitivity was measured by evaluating survival following antibiotic exposure.

154

Wild-type E. coli and ETCS were collected at the exponential phase of growth, and 1-mL

155

aliquots of cell suspension were added to 49 mL of sterilized LB medium containing 1 ×

156

MIC50, 2 × MIC50, and 4 × MIC50 concentrations of the corresponding antibiotics and

157

incubated at 37°C for 12 h. The optical density was measured at 600 nm (OD600). Each

158

strain was tested in triplicate, and sterilized LB medium was used as a blank control.

159

Reversibility of TCS resistance and antibiotic tolerance. The reversibility of

160

resistance/tolerance to TCS and antibiotics in adapted ETCS was determined by

161

subculturing in TCS-free medium. Briefly, after 100 subcultures, ETCS species were

162

collected, washed and re-suspended in saline. One milliliter of microbial suspension was

9 ACS Paragon Plus Environment

Environmental Science & Technology

163

inoculated into 49 mL of fresh LB medium without TCS, followed by 24 h of incubation.

164

The process was repeated for 20 additional cycles, and changes in the MIC50 to TCS was

165

monitored. Alterations in antibiotic tolerance were also evaluated every 5 subcultures by

166

monitoring the OD600 after 12 h of incubation in the presence of 25 mg/L penicillin. After

167

TCS exposure was ceased, the biofilm formation of ETCS was visualized by confocal laser

168

scanning microscopy.

169

Statistical analysis. At least three replicates were run per treatment. All data were

170

expressed as the means ± SD, and statistical differences were determined by an

171

independent t test or one-way analysis of variance (ANOVA). A p-value less than 0.05

172

(*) or 0.01 (**) indicated a significant difference.

10 ACS Paragon Plus Environment

Page 10 of 31

Page 11 of 31

173

Environmental Science & Technology

RESULTS AND DISCUSSION

174

Development of TCS resistance. Following exposure to 0.01 mg/L TCS for more

175

than 100 generations (> 100 days), E. coli developed high resistance to TCS as indicated

176

by the loss of growth inhibition at 0.02 mg/L TCS (Figure 1a) and marked augmentation

177

of the MIC50 from the original 0.033 mg/L to 0.562 mg/L (Figure 1b). The observed

178

resistance degree may not be as pronounced as that reported previously 8, 18, 19, which was

179

possibly due to the fluctuating culture conditions (e.g., the TCS concentration, exposure

180

time, and bacterial strain) under different treatments. In particular, we did not select the

181

potential mutant strain for this purpose. Instead, the susceptibility of the overall

182

population was assessed; thus, the tolerance profile can be maintained undisturbed.

183

Meanwhile, we did not observe any obvious morphological changes in the bacteria after

184

chronic exposure (Figure S1).

185

In general, the resistance evolution model presumes a trade-off between resistance

186

and fitness costs in the adapted populations in the absence of pollutant 29. However, no

187

detrimental effect on fitness was observed in the adapted population (as indicated by no

188

growth reduction in the adapted cells in the absence of TCS, as shown in Figure 1a),

189

indicating that the major resistance mechanism was related to alteration of specific target

190

genes rather than stress-induced global physiological changes. Thus, the molecular

191

mechanisms responsible for TCS resistance were determined by transcriptional analysis.

192

Illumina sequencing data of the tested samples are shown in Table S2 and Table S3. The

193

fold changes of the DEGs are displayed in the SI (Excel file 1). The transcriptomic data

194

were selectively validated via qRT-PCR analysis (Figure 2a). Upon exposure to 0.02

195

mg/L TCS, the total number of DEGs in adapted cells was notably lower than that in

11 ACS Paragon Plus Environment

Environmental Science & Technology

196

wild-type cells using E. coli alone as a biotic control (Figure 2b). The response pattern of

197

the adapted cells was grouped together with that of the untreated wild-type cells and was

198

distinct from the acute response in wild-type cells (Figure 2c). Collectively, the TCS

199

resistance mechanism was mostly attributed to alterations of target genes.

200

Chronic exposure to TCS triggers differential expression and mutation of

201

target genes. We focused on modifications of target gene expression in adapted cells.

202

TCS targets enoyl-acyl carrier protein reductases, blocking bacterial type II fatty acid

203

synthesis 30. Reportedly, fabI, which catalyzes the final step in each fatty acid elongation

204

cycle, is an important target of TCS 31. Kyoto Encyclopedia of Genes and Genomes

205

(KEGG) analysis verified that fatty acid-related pathways (i.e., synthesis, metabolism,

206

and degradation) were differentially enriched in adapted cells relative to wild-type E. coli

207

upon TCS exposure (Figure 3a). The KEGG enrichment analysis of DEGs between ETCS

208

and E. coli is displayed in the SI (Excel file 2). A summarized sketch of the key genes

209

involved in fatty acid synthesis in Figure 3b (quantified via qRT-PCR analysis) shows

210

differential expression levels of selected genes that are influenced by chronic TCS

211

exposure. Among them, fabI was significantly upregulated (log2 FC > 2.8), corroborating

212

the overexpression of target genes (Figure 3b and Figure 2a). Overexpression of fabI in

213

TCS-resistant bacteria has been previously observed for Staphylococcus aureus 32, 33 and

214

E. coli 8, 34.

215

Moreover, to reveal the genetic changes involved in TCS resistance, we conducted

216

whole-genome sequencing on ETCS and wild-type E. coli. TCS inhibits bacterial fatty acid

217

synthesis by binding enoyl reductase; therefore, missense mutations in target genes may

218

reduce TCS efficacy by altering the structures of the target proteins, i.e., fabI, fabB, fabD

12 ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31

Environmental Science & Technology

219

and fabZ (Figure 4). Mutation in fabI after TCS exposure was also reported for S. aureus

220

following ten passages with a gradient of TCS exposure concentrations 17 and for E. coli

221

with exposure to 0.2 mg/L TCS for 30 days 21. Notably, in different studies, the mutation

222

patterns may not be identical even for the same target gene due to dissimilar evolutionary

223

mechanisms. Other mechanisms conferring multidrug resistance, such as changes in the

224

outer membrane and biofilm formation, may also actively participate in TCS resistance

225

and are discussed below.

226

TCS treatment induces mild multi-antibiotic tolerance. TCS is generally

227

thought to induce coselection for antibiotic resistance; thus, we tested the susceptibility of

228

TCS-resistant bacteria to various types of antibiotics (i.e., beta-lactam, aminoglycoside,

229

and quinolone compounds). TCS-resistant cells showed decreased susceptibility to

230

antibiotics, i.e., penicillin, gentamicin, kanamycin, and ciprofloxacin, to various extents

231

(Figure 5). However, the tolerance degree was not specifically relevant to any class of

232

antibiotic (i.e., beta-lactam, aminoglycoside, and quinolone). The increases in antibiotic

233

cross-tolerance were less strong than those previously reported for TCS-resistant bacteria

234

18, 20, 21.

235

of wild-type E. coli (Figure S2). In studies performed on E. coli and P. aeruginosa, the

236

MIC50 to chloramphenicol and tetracycline increased 10-fold following TCS exposure 16.

237

This discrepancy was attributed to the low dose of TCS utilized in our study (0.01 mg/L);

238

at this concentration, TCS did not trigger significant reactive oxygen species (ROS)

239

generation or cell growth inhibition (Figure S3). Thus, antibiotic-resistant mutations via

240

the ROS stress response system 35 were less effective, which can be verified by the

241

unaltered mutation frequency 36 (data not shown, see the SI for the mutagenesis assay).

For example, the MIC50 of penicillin for ETCS was only 1.76-fold greater than that

13 ACS Paragon Plus Environment

Environmental Science & Technology

242

Alteration of the membrane structure and biofilm formation were responsible

243

for antibiotic tolerance. Fatty acids are important components of the phospholipids that

244

form the phospholipid bilayers out of which cell membranes are constructed. Thus,

245

interference with fatty acid synthesis by TCS results in alterations in the membrane

246

structure (Table 1) that diminish antibiotic uptake. The ratio of saturated to unsaturated

247

fatty acids rose from 0.64 ± 0.05 to 0.74 ± 0.04 (p < 0.05), indicating a decrease in

248

membrane permeability. Similar fatty acid compositional changes in the cell membrane

249

have been observed in E. coli, Bacillus subtilis and other bacteria subjected to various

250

environmental stresses, e.g., nanoparticles 37, 38 and environmental pollutants 39.

251

However, the degree of alteration in the relative proportions of fatty acids in this study

252

may not be as high as that reported in acute toxicity tests 37, 38. One possible explanation

253

is that alteration of the membrane structure is an adaptive strategy adopted in evolved

254

bacteria rather than an acute detrimental effect. Inhibition of membrane activity (i.e.,

255

intercalation into the bacterial cell membrane) due to acute TCS exposure has been

256

described previously 40, 41. However, this may be the first report of a protective response

257

from cells via alteration of the membrane structure with chronic TCS exposure.

258

Interestingly, TCS exposure induced notable biofilm formation in adapted cells, as

259

shown in Figure 6. Sublethal concentrations of chemicals (e.g., biocidal, antibiotic, and

260

quorum-sensing molecules) can promote biofilm development, which is an adaptive

261

strategy of bacteria to cope with various environmental stimuli 42, 43. Biofilm-associated

262

cells show low susceptibility to TCS due to limited diffusion 15, which may also be an

263

important mechanism underlying the observed antibiotic tolerance.

14 ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31

264

Environmental Science & Technology

Meanwhile, efflux is another common resistance strategy, and exposure to TCS has

265

been reported to effectively upregulate efflux pump genes in a few studies 19, 44. E. coli

266

harbors a large number of efflux pumps with a wide spectrum of substrates, and the most

267

relevant efflux pump for multidrug transporters is acrAB-tolC (the resistance nodulation

268

and cell division (RND) efflux pump) 45. However, we did not detect significant

269

upregulation of the acrAB-tolC pump based on our transcriptomic analysis. In contrast,

270

the pathway encoding ATP Binding Cassette (ABC) transporters (another type of efflux

271

pump) was enriched in adaptive cells according to the KEGG analysis (Figure 3a). The

272

enhanced expression of the ABC transporter-coding genes cysPU, ydcVTUS and gltI was

273

validated by qRT-PCR analysis (Figure 2a). Since only a few members of this family

274

have been identified as multidrug exporters in prokaryotes 46, we tentatively infer that the

275

contribution of the induction of efflux pumps to antibiotic tolerance in this study is

276

limited. Collectively, at environmentally relevant concentrations, TCS appears to affect

277

only the target genes and may not induce overall oxidative stress, thus avoiding triggering

278

overexpression of a specific efflux pump.

279

Persistent TCS resistance and reversible antibiotic tolerance. Whether reduced

280

TCS use will influence the evolved resistant strains to regain TCS susceptibility and lose

281

their derived antibiotic resistance/tolerance traits over time is still unclear. The MIC50 of

282

ETCS to TCS decreased slightly after stopping exposure for 20 generations (Figure 7a) but

283

was still greater than 16-fold higher than that of wild-type E. coli. The hereditary stability

284

assay suggested the ETCS can stably transfer resistance to their offspring. This effect was

285

attributed to inheritable resistance via genetic mutations at target sites because

15 ACS Paragon Plus Environment

Environmental Science & Technology

286

transcriptional regulation alone (i.e., overexpression of target genes in response to stress)

287

cannot induce transgenerational inheritance of acquired traits 47, 48.

288

However, the tolerance of ETCS towards penicillin dropped markedly as a function

289

of time, and after 20 generations, the evolved tolerance approached the wild-type level

290

(Figure 7b). Simultaneously, the level of biofilm formation returned to normal (similar to

291

that of wild-type E. coli) (Figure 7c), which further strengthened our conclusion that

292

antibiotic tolerance was largely attributed to biofilm formation. Little information is

293

currently available regarding the reversibility of TCS exposure-induced antibiotic

294

resistance in adapted bacteria and related mechanisms. Some sporadic data suggest that

295

most antibiotic resistance mechanisms are associated with a fitness cost, and that the

296

magnitude of the fitness cost determines the stability of resistance 49. Accordingly, the

297

lack of fitness cost in the current study implies that at the 0.01-mg/L concentration, TCS

298

triggered antibiotic tolerance mostly via functional adjustments (i.e., alteration of

299

membrane permeability and biofilm formation), which were unstable and not heritable.

300

Notably, the environmental concentration of TCS will drop gradually (not sharply) due to

301

continuous usage in other products, and how the attenuation effect influences the

302

dissipation of antibiotic resistance is of interest.

303

ENVIRONMENTAL IMPLICATIONS

304

This study demonstrated that chronic exposure of E. coli to an environmentally

305

relevant TCS dose induced transgenerational TCS resistance but non-heritable multidrug

306

tolerance. This work reveals that reduced release of TCS may be an efficient approach to

307

mitigate the propagation of antibiotic resistance. Future studies are needed to determine

308

the threshold concentration of TCS and the breakthrough time point that trigger

16 ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31

Environmental Science & Technology

309

pathogenic evolution. Identifying the behavior of horizontal gene transfer (another means

310

of acquiring antibiotic resistance) in the presence of low-dose TCS is equally important.

311

This information will be critical for accurate quantitative predictions regarding the impact

312

of TCS on the emergence and spread of resistant bacteria in the environment.

313 314

Supporting Information Availability: Supporting information is available free of charge

315

on the Internet at…

316

The detailed procedures for RNA preparation and library construction,

317

transcriptomic analysis, qRT-PCR validation, DNA extraction, whole-genome

318

sequencing and data processing, preparation of membrane samples for fatty acid

319

composition determination, transmission electron microscopy analysis of bacterial

320

morphology, measurement of intracellular ROS and the mutagenesis assay are described.

321

The primers used in the qRT-PCR analysis and their target classification are provided

322

(Table S1); Sequencing and assembly statistics for the transcriptomic data (Table S2);

323

The distributions of genes at differentially expressed levels (Table S3); Summary of the

324

total numbers of differentially expressed genes in ETCS + TCS compared to those of E.

325

coli + TCS (Excel file 1); KEGG enrichment analysis of differentially expressed genes

326

between ETCS and E. coli (upregulated) (Excel file 2); TEM images showing no obvious

327

morphological change in adapted ETCS compared with wild-type E. coli (Figure S1);

328

MIC50 determination of E. coli and ETCS towards penicillin (Figure S2); No obvious

329

adverse effect of 0.01 mg/L TCS on bacterial growth and intracellular ROS production

330

(Figure S3).

17 ACS Paragon Plus Environment

Environmental Science & Technology

331

Acknowledgments. This project was supported by the National Natural Science

332

Foundation of China (Grant 21777077). Partial support was also provided by the

333

Fundamental Research Funds for the Central Universities and the Interdiscipline

334

Research Funds of Beijing Normal University.

335 336

Notes― The authors declare no competing financial interests.

337

18 ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31

Environmental Science & Technology

338

REFERENCES

339

1.

340

the United States. Environ. Sci. Technol. 2014, 48 (7), 3603-3611.

341

2.

342

cancer risk. Int. J. Env. Res. Pub. He. 2014, 11 (2), 2209-2217.

343

3.

344

McNamara, P. J. Chronic exposure to triclosan sustains microbial community shifts and

345

alters antibiotic resistance gene levels in anaerobic digesters. Environ. Sci-proc. Imp.

346

2016, 18 (8), 1060-1067.

347

4.

348

induced changes in the antibiotic resistome of anaerobic digesters. Environ. Pollut. 2018,

349

241, 1182-1190.

350

5.

351

five parabens, triclosan, triclocarban and its transformation products in sewage sludge

352

from China. J. Hazard. Mater. 2019, 365, 502-510.

353

6.

354

susceptibility to triclosan in Salmonella enterica isolates from animals and humans and

355

association with multiple drug resistance. Int. J. Antimicrob. Ag. 2010, 36 (3), 247-251.

356

7.

357

Webb, C.; Iddles, R.; Maxwell, A.; Piddock, L. J. V. Quinolone-resistant gyrase mutants

358

demonstrate decreased susceptibility to triclosan. J. Antimicrob. Chemoth. 2017, 72 (10),

359

2755-2763.

360

8.

361

Transcriptomic analysis of triclosan-susceptible and -tolerant Escherichia coli O157: H19

362

in response to triclosan exposure. Microb. Drug Resist. 2014, 20 (2), 91-103.

363

9.

364

Duffy, G.; Fanning, S.; Nally, J. E.; Burgess, C. M. Proteomic and phenotypic analysis of

365

triclosan tolerant verocytotoxigenic Escherichia coli O157:H19. J. Proteomics 2013, 80,

366

78-90.

367

10.

368

triclosan may play a role in Pseudomonas aeruginosa antibiotic resistance in

Halden, R. U. On the need and speed of regulating triclosan and triclocarban in Dinwiddie, M. T.; Terry, P. D.; Chen, J. Recent evidence regarding triclosan and Carey, D. E.; Zitomer, D. H.; Kappell, A. D.; Choi, M. J.; Hristova, K. R.;

Fujimoto, M.; Carey, D. E.; McNamara, P. J. Metagenomics reveal triclosan-

Chen, J.; Meng, X. Z.; Bergman, A.; Halden, R. U. Nationwide reconnaissance of

Copitch, J. L.; Whitehead, R. N.; Webber, M. A. Prevalence of decreased

Webber, M. A.; Buckner, M. M. C.; Redgrave, L. S.; Ifill, G.; Mitchenall, L. A.;

Lenahan, M.; Sheridan, A.; Morris, D.; Duffy, G.; Fanning, S.; Burgess, C. M.

Sheridan, A.; Lenahan, M.; Condell, O.; Bonilla, S. R.; Sergeant, K.; Renaut, J.;

D'Arezzo, S.; Lanini, S.; Puro, V.; Ippolito, G.; Visca, P. High-level tolerance to

19 ACS Paragon Plus Environment

Environmental Science & Technology

369

immunocompromised hosts: evidence from outbreak investigation. BMC Res. Notes

370

2012, 5, 43-43.

371

11.

372

Pseudomonas aeruginosa PAO1 is due to FabV, a triclosan-resistant enoyl-acyl carrier

373

protein reductase. Antimicrob. Agents Ch. 2010, 54 (2), 689-698.

374

12.

375

Mechanism of triclosan inhibition of bacterial fatty acid synthesis. J. Biol. Chem. 1999,

376

274 (16), 11110-11114.

377

13.

378

Fotland, T. O.; Naterstad, K.; Kruse, H. Triclosan and antimicrobial resistance in

379

bacteria: an overview. Microb. Drug Resist. 2006, 12 (2), 83-90.

380

14.

381

across two membranes. Mol. Microbiol. 2000, 37 (2), 219-225.

382

15.

383

L.; Yaron, S. Effect of triclosan on Salmonella typhimurium at different growth stages

384

and in biofilms. Fems Microbiol. Lett. 2007, 267 (2), 200-206.

385

16.

386

resistance in the environment. Front. Microbiol. 2015, 5, 1-11.

387

17.

388

McBain, A. J. Fatty acid supplementation reverses the small colony variant phenotype in

389

triclosan-adapted Staphylococcus aureus: genetic, proteomic and phenotypic analyses.

390

Sci. Rep. 2018, 8 (1), 3876.

391

18.

392

in Salmonella enterica serovar Typhimurium. J. Med. Microbiol. 2009, 58 (4), 436-441.

393

19.

394

enterica and Escherichia coli O157 and cross-resistance to antimicrobial agents. J. Clin.

395

Microbiol. 2004, 42 (1), 73-78.

396

20.

397

Schweizer, H. P. Cross-resistance between triclosan and antibiotics in Pseudomonas

398

aeruginosa is mediated by multidrug efflux pumps: exposure of a susceptible mutant

Zhu, L.; Lin, J.; Ma, J.; Cronan, J. E.; Wang, H. Triclosan resistance of

Heath, R. J.; Rubin, J. R.; Holland, D. R.; Zhang, E. L.; Snow, M. E.; Rock, C. O.

Yazdankhah, S. P.; Scheie, A. A.; Hoiby, E. A.; Lunestad, B. T.; Heir, E.;

Zgurskaya, H. I.; Nikaido, H. Multidrug resistance mechanisms: drug efflux Tabak, M.; Scher, K.; Hartog, E.; Romling, U.; Matthews, K. R.; Chikindas, M.

Carey, D. E.; McNamara, P. J. The impact of triclosan on the spread of antibiotic Bazaid, A. S.; Forbes, S.; Humphreys, G. J.; Ledder, R. G.; O'Cualain, R.;

Birosova, L.; Mikulasova, M. Development of triclosan and antibiotic resistance Braoudaki, M.; Hilton, A. C. Adaptive resistance to biocides in Salmonella

Chuanchuen, R.; Beinlich, K.; Hoang, T. T.; Becher, A.; Karkhoff, S. R. R.;

20 ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

Environmental Science & Technology

399

strain to triclosan selects nfxB mutants overexpressing MexCD-OprJ. Antimicrob. Agents

400

Ch. 2001, 45 (2), 428-432.

401

21.

402

Non-antibiotic antimicrobial triclosan induces multiple antibiotic resistance through

403

genetic mutation. Environ. Int. 2018, 118, 257-265.

404

22.

405

triclosan, and methyl triclosan on thyroid hormone action and stress in frog and

406

mammalian culture systems. Environ. Sci. Technol. 2011, 45 (12), 5395-5402.

407

23.

408

toxicogenomic assessment of triclosan in human HepG2 cells using genome-wide

409

CRISPR-Cas9 screening. Environ. Sci. Technol. 2016, 50 (19), 10682-10692.

410

24.

411

different microalgal species in aquatic environment. J. Hazard. Mater. 2018, 342, 643-

412

650.

413

25.

414

of triclosan and triclocarban in selected wastewater treatment plants in Gauteng Province,

415

South Africa. Emerg. Contam. 2017, 3 (3), 107-114.

416

26.

417

TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions

418

and gene fusions. Genome Biol. 2013, 14 (4), R36.

419

27.

420

with various diameters on bacterial cellular membranes: cytotoxicity and adaptive

421

mechanisms. Chemosphere 2017, 185, 162-170.

422

28.

423

Alternating current influences anaerobic electroactive biofilm activity. Environ. Sci.

424

Technol. 2016, 50 (17), 9169-9176.

425

29.

426

is determined by a trade-off between treatment efficacy and relative fitness. B. Math.

427

Biol. 2004, 66 (4), 825-840.

Lu, J.; Jin, M.; Son Hoang, N.; Mao, L.; Li, J.; Coin, L. J. M.; Yuan, Z.; Guo, J.

Hinther, A.; Bromba, C. M.; Wulff, J. E.; Helbing, C. C. Effects of triclocarban,

Xia, P.; Zhang, X.; Xie, Y.; Guan, M.; Villeneuve, D. L.; Yu, H. Functional

Wang, S.; Poon, K.; Cai, Z. Removal and metabolism of triclosan by three

Lehutso, R. F.; Daso, A. P.; Okonkwo, J. O. Occurrence and environmental levels

Kim, D.; Pertea, G.; Trapnell, C.; Pimentel, H.; Kelley, R.; Salzberg, S. L.

Yang, F.; Jiang, Q.; Xie, W.; Zhang, Y. Effects of multi-walled carbon nanotubes

Wang, X.; Zhou, L.; Lu, L.; Lobo, F. L.; Li, N.; Wang, H.; Park, J.; Ren, Z. J.

Hall, R. J.; Gubbins, S.; Gilligan, C. A. Invasion of drug and pesticide resistance

21 ACS Paragon Plus Environment

Environmental Science & Technology

428

30.

Levy, C. W.; Roujeinikova, A.; Sedelnikova, S.; Baker, P. J.; Stuitje, A. R.;

429

Slabas, A. R.; Rice, D. W.; Rafferty, J. B. Molecular basis of triclosan activity. Nature

430

1999, 398 (6726), 383-384.

431

31.

432

economical assay to screen for triclosan binding to FabI. J. Biomol. Screen. 2016, 21 (4),

433

391-398.

434

32.

435

DeWolf, W. E.; Huang, J. Z.; McDevitt, D.; Miller, W. H.; Seefeld, M. A.; Newlander,

436

K. A.; Jakas, D. R.; Head, M. S.; Payne, D. J. Defining and combating the mechanisms of

437

triclosan resistance in clinical isolates of Staphylococcus aureus. Antimicrob. Agents Ch.

438

2002, 46 (11), 3343-3347.

439

33.

440

discovery. Bba-mol. Cell Biol. L. 2017, 1862 (11), 1300-1309.

441

34.

442

resistant Escherichia coli. J. Antimicrob. Chemoth. 2010, 65 (6), 1171-1177.

443

35.

444

the evolution of antibiotic resistance: evidence of mutation bias and its medicinal

445

implications. J. Biomol. Struct. Dyn. 2013, 31 (7), 729-733.

446

36.

447

Rose, J.; Achouak, W. Aged TiO2-based nanocomposite used in sunscreens produces

448

singlet oxygen under long-wave UV and sensitizes Escherichia coli to cadmium.

449

Environ. Sci. Technol. 2014, 48 (9), 5245-5253.

450

37.

451

membranes of bacteria: implication for an adaptive mechanism to the toxicity of carbon

452

nanotubes. Environ. Sci. Technol. 2014, 48 (7), 4086-4095.

453

38.

454

fullerene water suspension on bacterial phospholipids and membrane phase behavior.

455

Environ. Sci. Technol. 2007, 41 (7), 2636-2642.

456

39.

457

of five pseudomonad archetypes grown with and without toluene. Appl. Microbiol. Biot.

458

2000, 54 (3), 382-389.

Demissie, R. D.; Kabre, P.; Tuntland, M. L.; Fung, L. W. M. An efficient and

Fan, F.; Yan, K.; Wallis, N. G.; Reed, S.; Moore, T. D.; Rittenhouse, S. F.;

Yao, J.; Rock, C. O. Bacterial fatty acid metabolism in modern antibiotic Yu, B. J.; Kim, J. A.; Pan, J. Signature gene expression profile of triclosanWang, Z.; Xiong, M.; Fu, L.; Zhang, H. Oxidative DNA damage is important to

Santaella, C.; Allainmat, B.; Simonet, F.; Chaneac, C.; Labille, J.; Auffan, M.;

Zhu, B.; Xia, X.; Xia, N.; Zhang, S.; Guo, X. Modification of fatty acids in

Fang, J.; Lyon, D. Y.; Wiesner, M. R.; Dong, J.; Alvarez, P. J. J. Effect of a

Fang, J.; Barcelona, M. J.; Alvarez, P. J. J. Phospholipid compositional changes

22 ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31

Environmental Science & Technology

459

40.

Fang, J.; Stingley, R. L.; Beland, F. A.; Harrouk, W.; Lumpkins, D. L.; Howard,

460

P. Occurrence, efficacy, metabolism, and toxicity of triclosan. J. Environ. Sci. Heal. C.

461

2010, 28 (3), 147-171.

462

41.

463

effects of the antibacterial agent triclosan. Arch. Biochem. Biophys. 2001, 390 (1), 128-

464

136.

465

42.

466

Wang, L. Graphene oxide-silver nanocomposites modulate biofilm formation and

467

extracellular polymeric substance (EPS) production. Nanoscale 2018, 10 (41), 19603-

468

19611.

469

43.

470

in the development of E. coli biofilm. Environ. Sci. Technol. 2010, 44 (12), 4583-4589.

471

44.

472

resistance. J. Appl. Microbiol. 2002, 92, 65S-71S.

473

45.

474

Luisi, B. F. Multidrug efflux pumps: structure, function and regulation. Nat. Rev.

475

Microbiol. 2018, 16 (9), 523-539.

476

46.

477

B.; Martinez, J. L. Multidrug efflux pumps as main players in intrinsic and acquired

478

resistance to antimicrobials. Drug Resist. Update. 2016, 28, 13-27.

479

47.

480

byproducts leads to increase of antibiotic resistance in Pseudomonas aeruginosa.

481

Environ. Sci. Technol. 2014, 48 (14), 8188-8195.

482

48.

483

induce transgenerationally inheritable survival advantages via germline-to-soma

484

communication in Caenorhabditis elegans. Nat. Commun. 2017, 8, 14031.

485

49.

486

reverse resistance? Nat. Rev. Microbiol. 2010, 8 (4), 260-271.

Villalain, J.; Mateo, C. R.; Aranda, F. J.; Shapiro, S.; Micol, V. Membranotropic

Liu, S.; Cao, S.; Guo, J.; Luo, L.; Zhou, Y.; Lin, C.; Shi, J.; Fan, C.; Lv, M.;

Rodrigues, D. F.; Elimelech, M. Toxic effects of single-walled carbon nanotubes Levy, S. B. Active efflux, a common mechanism for biocide and antibiotic Du, D.; Xuan, W.; Neuberger, A.; Van, V. H. W.; Pos, K. M.; Piddock, L. J. V.;

Sara, H. A.; Blanco, P.; Manuel, A. R.; Corona, F.; Reales C, J. A.; Sanchez, M.

Lv, L.; Jiang, T.; Zhang, S.; Yu, X. Exposure to mutagenic disinfection

Kishimoto, S.; Uno, M.; Okabe, E.; Nono, M.; Nishida, E. Environmental stresses

Andersson, D. I.; Hughes, D. Antibiotic resistance and its cost: is it possible to

487

23 ACS Paragon Plus Environment

Environmental Science & Technology

(b)

0.8

E. coli E. coli + 0.02 mg/L TCS ETCS ETCS + 0.02 mg/L TCS

0.6

OD600

Percentage of OD600 (Relative to E. coli alone)

(a)

0.4 0.2 0.0 0

488 489 490 491 492 493 494

2

4

6

8

10

Page 24 of 31

120

E. coli ETCS

100 80 60

MIC50 = 0.033 mg/L

40 20

MIC50 = 0.562 mg/L

0 0.0 0.04 0.08 0.12 0.2 0.4 0.6 0.8 1.0 1.2

12

Time (h)

TCS concentration (mg/L)

Figure 1. ETCS developed considerable resistance to TCS. a) Growth curves of E. coli and ETCS in the presence or absence of 0.02 mg/L TCS; b) The MIC50 of E. coli and ETCS towards TCS. Cells were incubated in LB medium containing various concentrations of TCS for 12 h. The MIC50 values were calculated by fitting the data using the fourparameter logistic curve in Sigma Plot v10.0. Error bars, which are smaller than the symbols in some cases, represent the SD of triplicate experiments.

495

24 ACS Paragon Plus Environment

Page 25 of 31

Environmental Science & Technology

12

Log2 fold change

10

5

RNA-seq data qRT-PCR data

qRT-PCR data

(a)

8 6

R2 = 0.8483 4

3

2

2

4

4

6

10

8

RNA-seq data

2 0 atoB

fabI

atoA

atoE

Fatty acid metabolism

torC

torD

Quorum sensing

gadB

gmr

ydcV

ydcT

Two-component system

Gene name

(b)

ydcS

ydcU

cysP

cysU

ABC transporters

(c) 1

E. coli + TCS

0.5

ETCS + TCS

0 -0.5 -1

869

161

128

E. coli + TCS ETCS+ TCS E. coli

496 497 498 499 500 501 502 503 504 505 506 507

Figure 2. Transcriptomic analysis revealed different response patterns of E. coli and ETCS upon exposure to 0.02 mg/TCS for 5 h. (a) Validation of RNA sequencing data via qRTPCR analysis of selected genes. The inset shows the correlation analysis between RNAseq and qRT-PCR results from the same RNA samples. The Spearman correlation coefficient expresses the strength of the relationship between the RNA-seq and qRT-PCR variables. (b) Venn diagram comparing the number and overlap of DEGs in E. coli and ETCS in response to 0.02 mg/L TCS. (c) Dendrogram and unsupervised hierarchical clustering heat map (using Euclidean distance) of gene expression based on the log ratio FPKM data. Untreated E. coli was used as a biotic control. The qRT-PCR results of three biological replicates are shown as the mean values ± SD. RNA sequencing analysis was performed using two biological replicates.

508

25 ACS Paragon Plus Environment

Environmental Science & Technology

509 510 511 512 513 514 515 516

Figure 3. Functional enrichment analysis shows that pathways related to fatty acids are significantly impacted by TCS exposure. (a) KEGG enrichment analysis of DEGs (upregulated) in ETCS compared with E. coli based on transcriptomic sequencing data. (b) Quantification of DEGs relevant to fatty acid synthesis in ETCS compared with that in E. coli via qRT-PCR analysis. For both transcriptomic and qRT-PCR analyses, wild-type E. coli and adapted ETCS were exposed to 0.02 mg/L TCS in LB medium for 5 h, and E. coli alone was used as a biotic control.

26 ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

Environmental Science & Technology

(b)

(a)

fabD fabI

ETCS1 ETCS2 fabB

fabZ

Gene Changes Site Annotation fabD T → G 84115 F275V fabD C → T 84165 T291T fabI G → A 312262 N227N fabI A → C 312376 A189A fabI A → G 312478 N155N fabI T → C 312520 L141L fabI A → G 312616 T109T fabI T → C 312738 T69A fabB G → T 277083 T384T fabB A → G 277407 G276G fabB G → A 277440 D265D fabB C → A 277893 V114V fabB A → G 277899 F112F fabZ A → G 21509 F101F fabZ G → A 21530 A94A

517 518 519 520 521 522 523 524 525

Figure 4. The whole-genome sequencing analysis showing genetic mutations in target sites. (a) Genetic mutations annotated for adapted ETCS. Sequencing coverage for each clone (two biological replicates, ETCS1 and ETCS2, under 0.01-mg/L TCS exposure for 100 subcultures) is plotted according to color on concentric tracks. Each genetic change related to fatty acid synthesis, represented by colored dots, is marked at the appropriate genomic position. The outer circle represents the 4.8-Mb E. coli reference genome. (b) Table showing the list of corresponding gene mutations and annotations in ETCS according to Fig. 4a.

526

27 ACS Paragon Plus Environment

Environmental Science & Technology

100

60

Growth inhibition (% of E. coli alone) 528 529 530 531 532 533 534 535

**

40 20 1 × MIC50 2 × MIC50

Penicillin

120 100

E. coli ETCS

(d) *

**

60

*

40 20 1 × MIC50 2 × MIC50

120

Gentamicin

**

80

*

60 40 20

1 × MIC50 2 × MIC50

4 × MIC50

Kanamycin

120

E. coli ETCS

100

*

80

**

60 40 20 0

4 × MIC50

E. coli ETCS

100

0

4 × MIC50

80

0

527

**

80

0

(c)

(b) E. coli ETCS

Growth inhibition (% of E. coli alone)

120

Growth inhibition (% of E. coli alone)

Growth inhibition (% of E. coli alone)

(a)

Page 28 of 31

1 × MIC50 2 × MIC50

4 × MIC50

Ciprofloxacin

Figure 5. TCS-resistant bacteria (ETCS) exhibit less susceptibility to various types of antibiotics compared to E. coli. The inhibition rate of bacterial growth in the presence of various concentrations of (a) penicillin; (b) kanamycin; (c) gentamicin and (d) ciprofloxacin are shown. Bacteria were exposed to various concentrations of antibiotics at 37 °C for 12 h in LB medium. The MIC50 refers to the antibiotic concentration that inhibits 50% of wild-type E. coli growth. All experiments were performed at least in triplicate, and the results are expressed as the mean values ± SD. * and ** denote p < 0.05 and p < 0.01 compared to E. coli under the same treatment, respectively.

536

28 ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

PI

Merge

ETCS

E. coli

SYTO 9

537 538 539 540 541

Figure 6. Confocal laser scanning microscopy images comparing biofilm formation by E. coli and ETCS. E. coli and ETCS were exposed to 0.02 mg/L TCS in LB medium at 37 ºC for 24 h. Live cells were stained with SYTO 9 dye (green), and dead cells were stained with propidium iodide (PI) (red). The scale bar in each image is 20 μm.

542

29 ACS Paragon Plus Environment

Environmental Science & Technology

(b)

25

Stop TCS exposure

20

**

**

4 2 E. coli

50

100

Generation

SYTO 9

60

40 30

**

20

**

**

10 0

120

Stop TCS exposure

50

E. coli 100 105 110 115 120

Generation

PI

Merge

543

ETCS

after stopping exposure

(c)

**

15

Growth inhibition (% of E. coli alone)

Fold change of MIC50 (Relative to E. coli)

(a)

Page 30 of 31

544 545 546 547 548 549 550 551 552

Figure 7. ETCS exhibits irreversible TCS resistance and reversible penicillin tolerance after stopping TCS exposure. (a) Fold changes in the MIC50 to TCS; (b) changes in susceptibility to 25 mg/L penicillin; and (c) changes in biofilm formation via confocal laser scanning microscopy following subculture for an additional 20 passages without TCS exposure. The scale bar is 20 μm. After stopping exposure to TCS for the indicated number of generations, the bacteria were incubated with serial dilutions of TCS or 25 mg/L penicillin for 12 h in LB medium. All experiments were performed at least in triplicate, and the results are expressed as the mean values ± SD. ** denotes p < 0.01 relative to the MIC50 of E. coli and the inhibition rate of 25 mg/L penicillin for E. coli.

553

30 ACS Paragon Plus Environment

Page 31 of 31

554

Environmental Science & Technology

Table 1. Fatty acid profiles of the cell membranes in E. coli and ETCS. % of total fatty acids E. coli ETCS C6:0 3.028 ± 0.58 2.57 ± 0.36 C8:0 1.78 ± 0.27 1.46 ± 0.21 a C10:0 ND 1.09 ± 0.35 C11:0 7.67 ± 0.95 7.36 ± 1.19 C13:0 4.32 ± 0.43 3.94 ± 0.52 C14:1 4.06 ± 0.3 4.89 ± 0.3 C15:1 ND 0.73 ± 0.63 C15:0 4.31 ± 0.4 9.53 ± 1.09 C16:1 42.07 ± 1.05 40.07 ± 0.56 C16:0 3.43 ± 0.47 4.53 ± 0.18 C18:2 15.08 ± 1.41 11.90 ± 0.97 C18:0 14.24 ± 1.01 11.91 ± 1.96 b SFA 38.79 ± 2.18 42.41 ± 1.63 c UFA 61.21 ± 2.18 57.59 ± 1.63 SFA/UFA 0.64 ± 0.05 0.74 ± 0.04 The value is the average of three replicates ± SD. a Not detected. b Saturated fatty acids. c Unsaturated fatty acids. Fatty acids

555 556 557 558 559 560 561

31 ACS Paragon Plus Environment