AEM Accepted Manuscript Posted Online 24 April 2015 Appl. Environ. Microbiol. doi:10.1128/AEM.00648-15 Copyright © 2015, American Society for Microbiology. All Rights Reserved.

1

VirR-mediated resistance of Listeria monocytogenes against food antimicrobials and cross-

2

protection induced by exposure to organic acid salts

3 4

Jihun Kang1, Martin Wiedmann1, Kathryn J. Boor1, Teresa M. Bergholz1,2*

5 6 7 8

1

Department of Food Science, Cornell University, Ithaca, NY 14853

2

Department of Veterinary and Microbiological Sciences, North Dakota State University, Fargo,

ND 58108

9 10

Running title: Antimicrobial resistance mechanism in L. monocytogenes

11 12

*

Corresponding author

13

Mailing address:

14

PO 6050 Dept 7690

15

North Dakota State University

16

Fargo, ND 58108

17

Phone: 701-231-7692

18

Email: [email protected]

19

1

20

ABSTRACT

21

Formulation of ready-to-eat (RTE) foods with antimicrobial compounds constitutes an important

22

safety measure against foodborne pathogens such as Listeria monocytogenes. While the efficacy

23

of many commercially-available antimicrobial compounds has been demonstrated in a variety of

24

foods, current understanding of the resistance mechanisms employed by L. monocytogenes to

25

counteract these stresses is limited. In this study, we screened in-frame deletion mutants of two-

26

component system response regulators associated with cell envelope stress response for

27

increased sensitivity to commercially-available antimicrobial compounds (nisin, lauric arginate,

28

ε-polylysine, and chitosan). A virR deletion mutant showed increased sensitivity to all

29

antimicrobials and significantly greater loss of membrane integrity when exposed to nisin, lauric

30

arginate, and ε-polylysine (p < 0.05). The VirR-regulated operon, dltABCD, was shown to be the

31

key contributor to resistance against these antimicrobial compounds, whereas another VirR-

32

regulated gene mprF displayed antimicrobial-specific contribution to resistance. GUS reporter

33

fusion with the dlt promoter indicated that nisin does not specifically induce VirR-dependent

34

upregulation of dltABCD. Lastly, prior exposure of L. monocytogenes parent strain H7858 and

35

ΔvirR to 2% potassium lactate enhanced subsequent resistance against nisin and ε-polylysine (p

36

< 0.05). These data demonstrate that VirRS-mediated regulation of dltABCD is the major

37

resistance mechanism used by L. monocytogenes against cell envelope-damaging food

38

antimicrobials. Further, the potential for cross-protection induced by other food-related stresses

39

(e.g., organic acids) needs to be considered when applying these novel food antimicrobials as

40

hurdle strategy in RTE foods.

41

2

42

INTRODUCTION

43

Control of Listeria monocytogenes in ready-to-eat (RTE) foods is an important food safety goal

44

due to the high mortality rate associated with listeriosis, particularly in susceptible populations

45

such as pregnant women, the elderly, and those with a compromised immune system (1). L.

46

monocytogenes is of particular concern for those RTE foods that support growth of this pathogen

47

to high levels during refrigerated storage, which can potentially cause a life-threatening disease.

48

L. monocytogenes harbors a variety of stress-coping mechanisms that allow it to survive in

49

suboptimal environmental conditions associated with foods (e.g., acidic, osmotic, temperature

50

stress) (2). The ability of L. monocytogenes to tolerate and grow under such a wide range of

51

adverse conditions elevates the likelihood of foodborne transmission to a human host. Thus, a

52

multi-pronged approach (e.g., prevention of post-processing contamination and reformulation of

53

RTE foods with antimicrobials) to limit the number of L. monocytogenes in foods along the

54

farm-to-fork continuum is critical to reduce the potential for foodborne illnesses (3).

55

Natural antimicrobials are commonly applied to RTE foods to control foodborne pathogens such

56

as L. monocytogenes (4). Nisin (NIS) is one of the most widely used antimicrobials, which is a

57

bacteriocin naturally produced by Lactococcus lactis. Other antimicrobials that have been used

58

to control L. monocytogenes growth more recently include lauric arginate (LAE; derived from

59

lauric acid, L-arginine, and ethanol), ε-polylysine (EPL; produced by Streptomyces albulus), and

60

chitosan (CHI; derived from crustacean exoskeletons). The primary mode of action for NIS is

61

due to the binding to Lipid II (membrane-anchored cell-wall precursor) as a ‘docking molecule’

62

and subsequent aggregation of nisin molecules to induce pore formation in the bacterial

63

membrane, leading to dissipation of the proton motive force (5, 6). Likewise, the proposed

64

mechanism of action for LAE (7, 8), EPL (9-12), and CHI (13, 14) employ the common theme of 3

65

cell envelope disruption followed by concomitant disturbances in membrane-related cellular

66

functions, albeit exact molecular mechanism of action remains to be further elucidated. From a

67

bacterial perspective, the cell envelope not only provides structure to the cell, but also functions

68

as the primary barrier to exogenous aggressions as well as virulence modulator at the pathogen-

69

host interface (15). Thus, maintaining the integrity and function of the cell envelope under

70

fluctuating environmental conditions is critical for ensuring bacterial survival and transmission.

71

Sensing and managing of a cell envelope stress is typically facilitated through alternative sigma

72

factors and/or two-component systems (TCSs) (16). In the case of TCSs, sensing of a specific

73

input signal initiates autophosphorylation of a conserved histidine residue on a sensor histidine

74

kinase. The phosphoryl group is then transferred to an aspartic acid residue of the cognate

75

response regulator, which causes conformational change in the structure of the response

76

regulator, allowing it to function as a transcriptional activator (17). L. monocytogenes harbors 15

77

two-component systems and an orphan response regulator (RR) (18). Of these, 4 TCSs (liaRS,

78

lisRK, cesRK, and, virRS) have been reported to play a central role in modulating cell envelope

79

stress response (19). A number of genes previously implicated in L. monocytogenes nisin

80

resistance are also TCSs such as liaRS (20-22), lisRK (23), virRS (24, 25), as well as genes that

81

are part of TCS regulons including, telA (26), mprF (27), anrAB (24), dltABCD (28), and

82

lmo2229 (21, 22). As an abundance of evidence indicate that TCSs play a central role in the cell

83

envelope stress response and are induced by cationic antimicrobial peptides (CAMPs) and other

84

cell wall-acting antibiotics (16, 19, 29), we sought to determine whether these TCSs also have a

85

role in resistance against antimicrobials that may be used in foods to control L. monocytogenes.

86

Better understanding of resistance mechanisms used by L. monocytogenes against these

4

87

membrane-damaging agents may provide further insight on effective hurdle strategy as well as

88

development of inhibitors targeting these TCSs to improve the pathogen control in foods.

89 90

MATERIALS AND METHODS Bacterial strains, mutant construction, antimicrobials, and growth conditions. All L.

91

monocytogenes strains used in this study are listed in Table 1. Non-polar deletion mutants were

92

constructed from the parent strain H7858 using the splicing-by-overlap-extension (SOE) method

93

(30). All deletions mutants were confirmed by PCR and subsequent sequencing of the

94

chromosomal copy of the deletion allele. The stock antimicrobial concentrations were 500 µg/ml

95

(NIS), 1000 µg/ml (LAE), 50000 µg/ml (EPL), and 50000 µg/ml (CHI) and were prepared in

96

sterile H2O except CHI which was prepared in 1% (v/v) acetic acid as previously described (31).

97

L. monocytogenes strains were maintained in Brain Heart Infusion (BHI) broth at -80oC with 15%

98

glycerol. Prior to each experiment, strains were streaked onto BHI agar and incubated at 37oC for

99

24 h. A single colony was used to inoculate 5 ml BHI, followed by incubation at 37oC for 16 h

100

with shaking (230 rpm). Overnight cultures were transferred into 10 ml BHI (1:100) and

101

incubated until log phase (OD600 = 0.2-0.3) at 7oC.

102 103

Minimum Inhibitory Concentration (MIC) determination. The MIC experiment was

104

conducted to determine if the mutants had increased sensitivity to the antimicrobial effects (i.e.,

105

bactericidal and/or bacteriostatic) of selected compounds. On the day of the experiment,

106

antimicrobial stock solutions were diluted in sterile H2O and 100 µl was loaded into a sterile 96-

107

well flat-bottom polystyrene microtiter plate (Corning Inc., Corning, NY). Log phase cultures of

108

L. monocytogenes H7858 and its isogenic response regulator mutants grown at 7oC were diluted 5

109

1:10 in double strength (2X) BHI and 100 µl was loaded into corresponding wells using a multi-

110

channel pipette. The microtiter plate was covered with an adhesive film (Breathe-Easy®,

111

Diversified Biotech, Dedham, MA) to prevent contamination and the optical density (OD600) was

112

measured immediately with the Synergy H1 microplate reader (BioTek, Winooski, VT). The

113

microplate was stored at 7oC for 7 d and the OD600 was measured again. The MICs were defined

114

as the lowest concentration that completely inhibited growth (OD600 increase ≤ 0.05) after 7 d

115

incubation at 7oC (32, 33). MIC determination was based on 3 biological replicates for all strains.

116 117

Membrane integrity assay. The assay utilizes the nucleic acid stains SYTO 9 (Green

118

fluorescent) and propidium iodide (Red fluorescent). SYTO 9 can penetrate cells with intact as

119

well as compromised membranes, whereas propidium iodide can only penetrate those with

120

compromised membranes. The ratio of the green to red fluorescence thus can be used as an

121

indicator of the relative membrane integrity under antimicrobial stress. Prior to each experiment,

122

log phase cultures of L. monocytogenes H7858 and ΔvirR grown at 7oC were aliquoted into 1.5

123

ml tubes and stock solutions of NIS, LAE, EPL, and CHI were transferred 1:1 (500 µl

124

antimicrobial into 500 µl cells) into tubes at the previously determined MIC concentration for

125

ΔvirR except for NIS which was tested at 0.5 µg/ml. Higher concentrations of NIS resulted in the

126

saturation of the propidium iodide penetration, possibly due to the strong pore-forming activity

127

of NIS. Following incubation at 7oC for 24 h, cells (1 ml) were harvested by centrifugation

128

(6,800 X g for 5 min), resuspended in 1 ml 0.85% NaCl, and the assay was carried out using

129

these cells according to the manufacturer’s protocol (Live/Dead BacLight Bacterial Viability kit,

130

Molecular Probes, Inc., Eugene, OR). Washed cells (100 µl) were loaded into a 96-well flat-

131

bottom polystyrene microtiter plate (Corning Inc.), mixed with 100 µl 2X staining solution 6

132

(mixture of SYTO 9 and propidium iodide), and incubated at room temperature for 15 min in the

133

dark. The fluorescence intensity of each well was measured with a Synergy H1 microplate reader

134

(BioTek); SYTO 9 intensity was measured with the excitation wavelength of 485 nm and

135

emission wavelength of 530 nm while propidium iodide intensity was measured with the

136

excitation wavelength of 485 nm and emission wavelength of 630 nm. The raw fluorescence

137

readings were used to quantify the relative dye ratios for each treatment. The relative dye ratios

138

were calculated as the ratio of SYTO 9 to propidium iodide and divided by the fluorescence ratio

139

of the untreated parent strain (for the parent strain relative dye ratio) or by the fluorescence ratio

140

of untreated ΔvirR (for ΔvirR relative dye ratio). The membrane integrity assay was replicated

141

three times for all strain and treatment combinations.

142 143

Survival of H7858, ∆virR, ∆dltA, ∆mprF, and ∆anrAB exposed to NIS, LAE, EPL,

144

and CHI. The relative survival experiment was performed to determine the contributions of each

145

VirR regulon member to resistance to each antimicrobial compound. Cultures of L.

146

monocytogenes parent strain H7858 and in-frame deletion mutants of virR and VirR-regulated

147

genes dltA, mprF, and anrAB were grown to log phase at 7oC as described above. These cultures

148

were diluted in phosphate-buffered saline (PBS) and enumerated on BHI agar plates using a

149

spiral plater (Autoplate 4000, Spiral Biotech, Inc., Norwood, MA) to obtain the initial cell counts

150

(T0). The antimicrobial solutions (100 µl) were directly added to cultures to achieve the final

151

concentration of 0.5 µg/ml NIS, 20 µg/ml LAE, 500 µg/ml EPL, and 50 µg/ml CHI. These

152

antimicrobial concentrations were selected in order to specifically measure bacterial inactivation

153

rather than growth inhibition (i.e., MIC), and antimicrobial concentrations different than those

154

used in the MIC assay were needed to observe differences between the parent strain and the 7

155

mutants. Cells treated with antimicrobials were incubated at 7oC for 24 h and viable cells were

156

enumerated on BHI agar plates. The BHI agar plates were incubated at 37oC overnight and

157

colonies were counted using the Q-count colony counter (Spiral Biotech, Inc.). Viable cell counts

158

were transformed to log10CFU/ml and the log reduction in cell counts was determined as the

159

difference in log10CFU/ml between T24 and T0 for each strain. The survival experiment was

160

replicated three times for all strain and treatment combinations.

161 162

Cytochrome c binding assay. Cytochrome c binding assay was performed as previously

163

described (34-36) to assess whether deletion of virR and VirR-regulated genes (i.e., dltA and

164

mprF) leads to altered cell envelope charge. Briefly, cells grown to log phase at 7oC were

165

collected by centrifugation at 4,000 rpm and washed twice with 20 mM MOPS (3-(N-

166

morpholino)propanesulfonic acid) buffer (pH 7). The washed cells were adjusted to a

167

concentration of 108 CFU/ml in the same buffer and cytochrome c (Sigma-Aldrich, St. Louis,

168

MO) was added to a final concentration of 50 µg/ml. After 10 min incubation at room

169

temperature, the cell suspension was centrifuged at 13,000 rpm for 5 min and the supernatant

170

absorbance was measured at 410 nm. The absorbance of cytochrome c in the absence of cells

171

was compared to the absorbance of cytochrome c with cells (i.e., maximum unbound cytochrome

172

c) as a measure of relative cytochrome c binding using the following equation:

% cytochrome bound = 100 −

OD OD

,

× 100

,

173

GUS (β-glucuronidase) activity assay. The Pdlt-GUS transcriptional fusion was

174

constructed to assay for the activation of VirR upon exposure to NIS (Table 1). The quantitative

8

175

determination of GUS activity directed from VirR-dependent dlt promoter was conducted as

176

previously described (37). Cells were grown to log phase at 7oC in BHI and were then exposed to

177

a sublethal concentration (100 ng/ml) of NIS or sterile H2O for 12 h at 7oC. Cells (1 ml) were

178

harvested prior to NIS exposure and following 12 h incubation by centrifugation and washed

179

with 1 ml ABlight buffer (60 mM K2HPO4, 40 mM KH2PO4, 0.1 M NaCl), followed by

180

resuspension in 1 ml ABlight buffer. Aliquotes (100 µl) were taken from each sample and

181

enumerated on BHI agar plates supplemented with 10 µg/ml chloramphenicol. The remaining

182

cells were frozen at -80oC. Prior to GUS measurements, frozen cells were thawed and 135 µl of

183

CellLytic B reagent (Sigma-Aldrich) was added to lyse cells for 10 min at room temperature.

184

Lysed cells were loaded into 96-well flat-bottomed black polystyrene plates (Corning Inc.) in

185

duplicate and 20 µl of MUG (Sigma-Aldrich) was added for a final concentration of 0.4 mg/ml.

186

For a parallel set of cells, the same volume of H2O was added instead of MUG to measure the

187

background fluorescence from the cells in the absence of the substrate. The enzymatic

188

fluorescent byproduct 4-methylumbelliferone (MU; Sigma-Aldrich) was used to generate a

189

standard curve from which respective GUS activity was inferred. The enzymatic reaction was

190

stopped after 30 min by the addition of the Stop solution (1 M Na2CO3). The fluorescence was

191

measured in the Synergy H1 plate reader (BioTek) with the excitation wavelength of 365 nm and

192

the emission wavelength of 460 nm. The background fluorescence from the cells without MUG

193

was subtracted from the corresponding wells and the MU standard curve was used to calculate

194

GUS activity, reported as nM MU/log CFU/min.

195 196

Organic acid-induced cross-protection against NIS, LAE, EPL, and CHI. L.

197

monocytogenes strains H7858 and ΔvirR grown to log phase at 7oC were transferred (1 ml) to 9 9

198

ml BHI, BHI + potassium lactate (PL), and sodium diacetate (SD) for the final concentration of 2%

199

PL and 0.14% SD. BHI containing no acid (CTRL) and acid-supplemented BHI were all

200

adjusted to pH 6.0 to separate the effects of pH from that of the acid salts. The initial pH of the

201

BHI treatment media were 7.35 ± 0.01, 7.35 ± 0.01, 6.65 ± 0.06 for plain BHI, BHI + PL, and

202

BHI + SD, respectively. Since the salts of organic acids were used, the direct impact on the pH

203

of BHI media was not as profound as using lactic acid or acetic acid. L. monocytogenes strains

204

were exposed to acid conditions for 8 h at 7oC prior to antimicrobial addition at the final

205

concentration of 1 µg/ml NIS, 20 µg/ml LAE, 1000 µg/ml EPL, 100 µg/ml CHI. Acid-exposed

206

cells were enumerated prior to antimicrobial challenge and following 24 h incubation at 7oC in

207

the presence of indicated antimicrobial concentrations as described above. To examine organic

208

acid-induced cross protection, antimicrobial concentrations sufficient to cause at least 1 log

209

reduction of the parent strain were tested for each antimicrobial.

210 211

Statistical analysis. The MIC of the mutants was compared to the MIC of the parent

212

strain using the ‘Interval’ package in R-Studio v 0.98.1103, which performs logrank tests for

213

interval-censored data based on non-parametric maximum likelihood estimation of the survival

214

distributions (38). For statistical analysis of the membrane integrity assay, VirR regulon mutant

215

survival assay, and cytochrome c binding assay, one-way analysis of variance (ANOVA) was

216

carried out for each antimicrobial using the relative dye ratio, decrease in log10CFU/ml, and

217

relative binding, respectively, as the response variable. The strain genotype and replicate were

218

modeled as fixed effects. GUS activity assay was analyzed with a three-way ANOVA model

219

using GUS activity as the response variable and the strain genotype, NIS concentration, assay

220

time, and their interaction as fixed effects. Similar two-way ANOVA model with the decrease in 10

221

log10CFU/ml as the response variable and the strain genotype, acid treatment, and their

222

interaction as fixed effects was used to assess the cross-protection induced by organic acids. The

223

Tukey multiple comparison procedure was applied to all ANOVA results. Adjusted p values of


331

0.05) whereas both PL and SD had significant protective effect against subsequent LAE

332

exposure for ΔvirR with a mean decrease in cell density of 3.2 ± 0.0 and 4.3 ± 0.2 log10CFU/ml,

333

respectively, compared to 4.7 ± 0.1 log10CFU/ml without acid exposure (p < 0.05). Organic acid-

334

induced cross-protection against EPL was similar to NIS; prior exposure to PL increased EPL

15

335

resistance (p < 0.05) whereas SD had no significant effect on subsequent EPL resistance (p >

336

0.05) for both the parent strain and ΔvirR. The potential cross-protective effect of organic acids

337

against CHI was less pronounced. Prior exposure to PL had no effect on the parent strain and

338

significant (p < 0.05) but numerically limited (0.3 log10CFU/ml) cross-protective effect on ΔvirR.

339

Prior exposure to SD slightly increased sensitivity of the parent strain and ΔvirR (by 0.41 ± 0.23

340

and 0.22 ± 0.03 log10CFU/ml, respectively) to CHI but this increased sensitivity was also

341

numerically limited.

342 343

DISCUSSION

344

The TCS response regulator VirR plays a key role in resistance to a range of food

345

antimicrobials. In this study, we have identified VirR as a critical transcriptional regulator for

346

resistance against antimicrobials used in foods. Reduced membrane integrity for ΔvirR under

347

NIS, LAE, and EPL stress suggests that the absence of VirR renders cells more susceptible to

348

membrane-perturbing effects of these compounds, likely due to increased binding. While ΔvirR

349

exhibited increased sensitivity to CHI stress, the loss of membrane integrity for ΔvirR was

350

similar to the parent strain under CHI stress. This observation is consistent with a previous report

351

on CHI mechanism of action in which teichoic acids were proposed to be the main molecular

352

target of CHI, followed by subsequent disruption of membrane-related functions (13).

353

In L. monocytogenes EGD, VirR has been reported to regulate 12 genes, and along with VirS, its

354

cognate sensor kinase, this signal transduction system has been implicated as a critical virulence

355

determinant in L. monocytogenes through modulating the interaction between L. monocytogenes

356

cells and components of the innate immune system (25). Further, the VirRS regulon is composed 16

357

of genes with previously ascribed roles in resistance against animal- and plant-derived cationic

358

antimicrobial peptides (CAMP). While VirRS has been studied extensively in regard to

359

virulence-related functions in the host and resistance against therapeutic antibiotics, relatively

360

little is known about its role in resistance against antimicrobials used in foods. Our findings of

361

the universal contribution of VirR toward antimicrobial resistance relevant to food preservation

362

further highlight the important functional role of VirR and its regulon in pathogen survival and

363

transmission in different environments (e.g., foods vs human host). Since the contribution of

364

VirR in resistance to antimicrobials used in foods was assessed in a laboratory medium, further

365

examination in appropriate food systems is warranted. Moreover, the evaluation of potential

366

growth phase-dependent effects on VirR-mediated antimicrobial resistance may be of importance

367

as there may be overlapping mechanisms that could contribute to resistance due to the activation

368

of the stress response in stationary phase (46).

369

VirR-regulated dltABCD is the main contributor to resistance against bactericidal

370

effects of food antimicrobials. Members of the VirR regulon (25) such as dltABCD (28, 35, 47),

371

mprF (27, 48, 49), and anrAB (24) are conserved in many Gram positive bacteria and are

372

important for conferring resistant to a variety of antimicrobial compounds including NIS and

373

other cationic peptides of animal, plant, and microbial origin (39). The relative contribution of

374

VirR to resistance to each antimicrobial is summarized in Figure 6. The side-by-side comparison

375

of these mutants against the parent strain and ΔvirR indicated that DltABCD is the major

376

contributor of VirR-mediated resistance, as ΔdltA was as sensitive as ΔvirR in the presence of all

377

4 antimicrobials tested here. A possible explanation for this observation is that the increased

378

electropositive charge facilitated by DltABCD results in electrostatic repulsion of cationic

379

antimicrobials, thus preventing antimicrobials from binding to their targets (35, 39, 50). Recently, 17

380

Saar-Dover et al. has further refined this notion with a newly proposed mechanism of Dlt-

381

mediated increase in cell wall density (51). This resistance mechanism based on cell-wall

382

thickening has also been suggested from transcriptomic analysis of a NIS-resistant Lactococcus

383

lactis strain (50).

384

MprF shares a functional similarity with DltABCD in that the protein alters cell envelope charge

385

by modifying the membrane lipid phosphatidylglycerol with L-lysine (27, 39). The relative

386

resistance contribution by MprF was highly varied among antimicrobials. The relative

387

contribution of MprF to NIS resistance was less, compared to VirR, consistent with a previous

388

report, which showed that the MIC of ΔmprF under NIS stress was 2-fold higher compared to

389

ΔvirR (24). Our results also demonstrated a minor effect of mprF deletion on LAE and CHI

390

resistance. In Gram positive bacteria, the cell envelope is composed of the phospholipid bilayer

391

shrouded by multiple layers of cross-linked murein which are further decorated with cell wall

392

glycopolymers such as teichoic acids (52, 53). Since MprF is responsible for reduced anionicity

393

of the membrane (as opposed to the cell wall), we hypothesize that the presence of DltABCD in

394

ΔmprF imparts a sufficient ‘shielding effect’ which can reduce local concentration of

395

antimicrobials in the vicinity of cell envelopes. Interestingly, ΔmprF was shown to be more

396

susceptible to EPL compared to ΔvirR or ΔdltA. The greater sensitivity of ΔmprF compared to

397

ΔvirR suggests a possible co-regulation of mprF by other regulators, under EPL stress, despite its

398

initial classification into the VirR regulon (25).

399

VirRS signal transduction pathway is not induced by NIS. We also sought to

400

determine whether the presence of NIS specifically induces VirRS-dependent transcriptional

401

regulation of dltABCD. In the absence of VirR or VirS, dltABCD promoter activity was minimal

402

or negligible, consistent with its reported dependence on VirRS (25). While there was an obvious 18

403

increase in GUS activity during 12 h incubation at 7oC, the increase in GUS activity due to NIS

404

was not significant, indicating that NIS only weakly induces or does not specifically induce

405

VirR-dependent regulation of dltABCD. Although NIS has been shown to induce dlt operon

406

expression in B. subtilis (54) and Clostridium difficile (55), it was shown to be weakly induced in

407

these studies. Further, no significant up-regulation of dltA was observed in spontaneous leucocin-

408

or pediocin-resistant mutants of L. monocytogenes, even though a higher D-alanine content in

409

teichoic acids was observed in these strains (36). Thus, the VirR-dependent increase in dlt

410

transcription observed here appears to be the result of a growth-dependent effect rather than a

411

specific induction by NIS. Growth phase-dependent regulation of dltABCD and mprF has also

412

been shown in Staphylococcus aureus in which the expression of these genes is upregulated

413

during the exponential phase whereas the opposite occurs during the stationary phase (34).

414

Pre-exposure to potassium lactate increases subsequent resistance to nisin and ε-

415

polylysine. L. monocytogenes typically encounters multiple stresses in a food environment and

416

adaptation to a stress condition may lead to cross-protection against a subsequent stress. For

417

example, osmotic stress and acid stress have been reported to induce cross protection against NIS

418

in L. monocytogenes (20, 42). Nevertheless, this potential cross-protective effect has not been

419

assessed for other relevant food antimicrobial compounds. Further, the evidence of VirR-regulon

420

induction under acid stress prompted us to hypothesize that a potential cross-protective effect

421

would be VirR-mediated (43-45). Our results indicated that exposure to 2% PL increased

422

subsequent NIS and EPL resistance of both the parent strain and ΔvirR. NIS resistance has

423

previously been associated with alterations in cytoplasmic membrane fatty acid compositions,

424

reflecting more rigid membrane fluidity (42, 56, 57). Further, alterations in membrane fatty acid

425

profiles (e.g., higher branched to straight chain fatty acid ratio) during growth at 10oC were 19

426

associated with more fluid membrane and resulted in increased NIS sensitivity in L.

427

monocytogenes (58). Acid adaptation in L. monocytogenes also alters membrane fatty acid

428

composition that is suggestive of more rigid membrane fluidity, which may act as a defense

429

mechanism by restricting the diffusion of acids across the membrane (59, 60). Thus, the

430

increased rigidity of the cytoplasmic membrane during adaptation to potassium lactate could at

431

least partially increase subsequent resistance against NIS.

432

Transcriptional profiling of L. monocytogenes during adaptation to lactic, acetic, hydrochloric

433

acid has also previously indicated a consistent pattern of alterations in cell membrane including

434

down-regulation of genes involved in branched chain fatty acid synthesis (43). Compared to SD,

435

PL was shown to have a stronger antilisterial activity as well as more pronounced down-

436

regulation of branched-chain fatty acid-related genes (43) even when the concentration of

437

undissociated form of SD was higher (3.7 mmol/L for SD compared to 1.4 mmol/L for PL),

438

which determines the primary inhibitory effects (60). In our study, the estimated concentration of

439

undisscoiated form of PL and SD were 1.1 mmol/L and 0.54 mmol/L, respectively using the

440

Henderson-Hasselbalch equation. Thus, relatively stronger effect of PL can be expected

441

compared to SD, which may explain profound PL-induced cross-protective effect. As the mode

442

of action for EPL also involves cell membrane disruption (9, 11, 12), it is reasonable to speculate

443

that the membrane fatty acid modification is linked with PL-induced cross-protection against

444

EPL.

445

The pre-exposure of L. monocytogenes to PL induced cross-protection against LAE only for

446

ΔvirR but not the parent strain. Consistent with the observation that dltABCD is the major

447

resistance determinant for LAE, it may be the case that the regulation of cell wall glycopolymers

448

in the parent strain provides protection against LAE whereas in the absence of VirR-dependent 20

449

regulation of cell wall structures, acid-induced changes in cytoplasmic membrane affords

450

additional protection. In the case of CHI, relatively minor effects of either acid on subsequent

451

CHI resistance is consistent with a previous report that the primary molecular target of CHI is

452

likely teichoic acids with subsequent extraction of membrane lipids (13). It is worth noting that

453

the increased sensitivity of ΔvirR under CHI stress (100 µg/ml) was not as evident compared to

454

the survival experiment with 50 µg/ml CHI (Fig. 2). This could be due to the concentration-

455

dependent saturation of the bactericidal effect of CHI, upon which there is no additional lethality

456

even with increasing concentrations (61).

457

Conclusions. This study shows a critical role of the L. monocytogenes two-component

458

system response regulator VirR in resistance against antimicrobials intended for food

459

preservation. Improved understanding of resistance mechanisms may facilitate the development

460

of effective hurdle strategies utilizing different molecular targets (33). Additionally, the potential

461

for cross-protection induced by a food-relevant stress (e.g., organic acids) should be taken into

462

consideration to avoid unintended over-estimation of preservative strength in a food system (20,

463

42). A mechanistic understanding of the interaction between antimicrobials and the target

464

pathogen also facilitates better delineation of how other food constituents may enhance (57) or

465

decrease (62) the antimicrobial efficacy. Lastly, identification and characterization of TCSs with

466

respect to stress tolerance and virulence have potential implication in development of better

467

control strategies for the pathogen as novel antimicrobial agents may be developed to target these

468

molecular targets (63, 64). In light of recent advances in high-throughput sequencing

469

technologies and decreasing cost, future efforts in understanding global changes in the

470

transcriptome of L. monocytogenes under antimicrobial challenge can be expected to provide

471

more comprehensive view of bacterial stress response and resistance mechanisms (65). 21

472

ACKNOWLEDGEMENTS

473

This work was supported by New York Sea Grant R/SHH-16, funded under award

474

NA07OAR4170010 from the National Sea Grant College Program of the U.S. Department of

475

Commerce’s National Oceanic and Atmospheric Administration, to the Research Foundation of

476

State University of New York, and by Agriculture and Food Research Initiative grant 2010-

477

65201-20575 from the U.S. Department of Agriculture, National Institute of Food and

478

Agriculture, Food Safety Program.

479

We would like to thank Barbara Bowen and Rebecca Schmidt for constructing the

480

deletion mutants used in this study, and Matthew Stasiewicz for assistance with the analysis of

481

MIC data.

482

22

REFERENCES

483 484

1.

Vazquez-Boland JA, Kuhn M, Berche P, Chakraborty T, Dominguez-Bernal G,

485

Goebel W, Gonzalez-Zorn B, Wehland J, Kreft J. 2001. Listeria pathogenesis and

486

molecular virulence determinants. Clin. Microbiol. Rev. 14:584-640.

487

2.

survive. Int. J. Food Microbiol. 113:1-15.

488 489

Gandhi M, Chikindas ML. 2007. Listeria: A foodborne pathogen that knows how to

3.

ILSI Research foundation/Risk Science Institute. 2005. Achieving continuous

490

improvement in reductions in foodborne listeriosis--a risk-based approach. J. Food Prot.

491

68:1932-1994.

492

4.

review of food science and technology 3:381-403.

493 494

5.

6.

7.

Pattanayaiying R, A HK, Cutter CN. 2014. Effect of lauric arginate, nisin Z, and a combination against several food-related bacteria. Int. J. Food Microbiol. 188:135-146.

499 500

Breukink E, de Kruijff B. 2006. Lipid II as a target for antibiotics. Nat. Rev. Drug. Discov. 5:321-332.

497 498

Breukink E, de Kruijff B. 1999. The lantibiotic nisin, a special case or not? Biochim. Biophys. Acta. 1462:223-234.

495 496

Juneja VK, Dwivedi HP, Yan X. 2012. Novel natural food antimicrobials. Annual

8.

Rodriguez E, Seguer J, Rocabayera X, Manresa A. 2004. Cellular effects of

501

monohydrochloride of L-arginine, N-lauroyl ethylester (LAE) on exposure to Salmonella

502

Typhimurium and Staphylococcus aureus. J. Appl. Microbiol. 96:903-912.

503 504

9.

Shima S, Matsuoka H, Iwamoto T, Sakai H. 1984. Antimicrobial action of epsilonpoly-L-lysine. J Antibiot. 37:1449-1455.

23

505

10.

Liu H, Pei H, Han Z, Feng G, Li D. 2015. The antimicrobial effects and synergistic

506

antibacterial mechanism of the combination of ε-Polylysine and nisin against Bacillus

507

subtilis. Food Control 47:444-450.

508

11.

Li YQ, Han Q, Feng JL, Tian WL, Mo HZ. 2014. Antibacterial characteristics and

509

mechanisms of epsilon-poly-lysine against Escherichia coli and Staphylococcus aureus.

510

Food Control. 43:22-27.

511

12.

Ye R, Xu H, Wan C, Peng S, Wang L, Xu H, Aguilar ZP, Xiong Y, Zeng Z, Wei H.

512

2013. Antibacterial activity and mechanism of action of ε-poly-l-lysine. Biochem.

513

Biophys. Res. Commun. 439:148-153.

514

13.

chitosan as an antibacterial compound. Appl. Environ. Microbiol. 74:3764-3773.

515 516

14.

15.

16.

17.

Hoch JA. 2000. Two-component and phosphorelay signal transduction. Curr. Opin. Microbiol. 3:165-170.

523 524

Jordan S, Hutchings MI, Mascher T. 2008. Cell envelope stress response in Grampositive bacteria. FEMS Microbiol. Rev. 32:107-146.

521 522

Carvalho F, Sousa S, Cabanes D. 2014. How Listeria monocytogenes organizes its surface for virulence. Front. Cell. Infect. Microbiol. 4:48.

519 520

Rabea EI, Badawy ME, Stevens CV, Smagghe G, Steurbaut W. 2003. Chitosan as antimicrobial agent: applications and mode of action. Biomacromolecules 4:1457-1465.

517 518

Raafat D, von Bargen K, Haas A, Sahl HG. 2008. Insights into the mode of action of

18.

Glaser P, Frangeul L, Buchrieser C, Rusniok C, Amend A, Baquero F, Berche P,

525

Bloecker H, Brandt P, Chakraborty T, Charbit A, Chetouani F, Couve E, de

526

Daruvar A, Dehoux P, Domann E, Dominguez-Bernal G, Duchaud E, Durant L,

527

Dussurget O, Entian KD, Fsihi H, Garcia-del Portillo F, Garrido P, Gautier L,

24

528

Goebel W, Gomez-Lopez N, Hain T, Hauf J, Jackson D, Jones LM, Kaerst U, Kreft

529

J, Kuhn M, Kunst F, Kurapkat G, Madueno E, Maitournam A, Vicente JM, Ng E,

530

Nedjari H, Nordsiek G, Novella S, de Pablos B, Perez-Diaz JC, Purcell R, Remmel B,

531

Rose M, Schlueter T, Simoes N, Tierrez A, Vazquez-Boland JA, Voss H, Wehland J,

532

Cossart P. 2001. Comparative genomics of Listeria species. Science. 294:849-852.

533

19.

Nielsen PK, Andersen AZ, Mols M, van der Veen S, Abee T, Kallipolitis BH. 2012.

534

Genome-wide transcriptional profiling of the cell envelope stress response and the role of

535

LisRK and CesRK in Listeria monocytogenes. Microbiology. 158:963-974.

536

20.

Bergholz TM, Tang S, Wiedmann M, Boor KJ. 2013. Nisin resistance of Listeria

537

monocytogenes is increased by exposure to salt stress and is mediated via LiaR. Appl.

538

Environ. Microbiol. 79:5682-5688.

539

21.

Collins B, Guinane CM, Cotter PD, Hill C, Ross RP. 2012. Assessing the

540

contributions of the LiaS histidine kinase to the innate resistance of Listeria

541

monocytogenes to nisin, cephalosporins, and disinfectants. Appl. Environ. Microbiol.

542

78:2923-2929.

543

22.

Gravesen A, Kallipolitis B, Holmstrom K, Hoiby PE, Ramnath M, Knochel S. 2004.

544

pbp2229-mediated nisin resistance mechanism in Listeria monocytogenes confers cross-

545

protection to class IIa bacteriocins and affects virulence gene expression. Appl. Environ.

546

Microbiol. 70:1669-1679.

547

23.

Cotter PD, Guinane CM, Hill C. 2002. The LisRK signal transduction system

548

determines the sensitivity of Listeria monocytogenes to nisin and cephalosporins.

549

Antimicrob. Agents Chemother. 46:2784-2790.

25

550

24.

Collins B, Curtis N, Cotter PD, Hill C, Ross RP. 2010. The ABC transporter AnrAB

551

contributes to the innate resistance of Listeria monocytogenes to nisin, bacitracin, and

552

various beta-lactam antibiotics. Antimicrob. Agents Chemother. 54:4416-4423.

553

25.

Mandin P, Fsihi H, Dussurget O, Vergassola M, Milohanic E, Toledo-Arana A, Lasa

554

I, Johansson J, Cossart P. 2005. VirR, a response regulator critical for Listeria

555

monocytogenes virulence. Mol. Microbiol. 57:1367-1380.

556

26.

Collins B, Joyce S, Hill C, Cotter PD, Ross RP. 2010. TelA contributes to the innate

557

resistance of Listeria monocytogenes to nisin and other cell wall-acting antibiotics.

558

Antimicrob. Agents Chemother. 54:4658-4663.

559

27.

Thedieck K, Hain T, Mohamed W, Tindall BJ, Nimtz M, Chakraborty T, Wehland

560

J, Jansch L. 2006. The MprF protein is required for lysinylation of phospholipids in

561

listerial membranes and confers resistance to cationic antimicrobial peptides (CAMPs) on

562

Listeria monocytogenes. Mol. Microbiol. 62:1325-1339.

563

28.

Abachin E, Poyart C, Pellegrini E, Milohanic E, Fiedler F, Berche P, Trieu-Cuot P.

564

2002. Formation of D-alanyl-lipoteichoic acid is required for adhesion and virulence of

565

Listeria monocytogenes. Mol. Microbiol. 43:1-14.

566

29.

10:179-186.

567 568

Peschel A. 2002. How do bacteria resist human antimicrobial peptides? Trends Microbiol.

30.

Wiedmann M, Arvik TJ, Hurley RJ, Boor KJ. 1998. General stress transcription factor

569

sigmaB and its role in acid tolerance and virulence of Listeria monocytogenes. J.

570

Bacteriol. 180:3650-3656.

26

571

31.

Kang J, Stasiewicz MJ, Murray D, Boor KJ, Wiedmann M, Bergholz TM. 2014.

572

Optimization of combinations of bactericidal and bacteriostatic treatments to control

573

Listeria monocytogenes on cold-smoked salmon. Int. J. Food Microbiol. 179:1-9.

574

32.

Branen JK, Davidson PM. 2004. Enhancement of nisin, lysozyme, and monolaurin

575

antimicrobial activities by ethylenediaminetetraacetic acid and lactoferrin. Int. J. Food

576

Microbiol. 90:63-74.

577

33.

Brandt AL, Castillo A, Harris KB, Keeton JT, Hardin MD, Taylor TM. 2010.

578

Inhibition of Listeria monocytogenes by food antimicrobials applied singly and in

579

combination. J. Food Sci. 75:M557-563.

580

34.

Matsuo M, Oogai Y, Kato F, Sugai M, Komatsuzawa H. 2011. Growth-phase

581

dependence of susceptibility to antimicrobial peptides in Staphylococcus aureus.

582

Microbiology. 157:1786-1797.

583

35.

Peschel A, Otto M, Jack RW, Kalbacher H, Jung G, Gotz F. 1999. Inactivation of the

584

dlt operon in Staphylococcus aureus confers sensitivity to defensins, protegrins, and other

585

antimicrobial peptides. J. Biol. Chem. 274:8405-8410.

586

36.

Vadyvaloo V, Arous S, Gravesen A, Hechard Y, Chauhan-Haubrock R, Hastings

587

JW, Rautenbach M. 2004. Cell-surface alterations in class IIa bacteriocin-resistant

588

Listeria monocytogenes strains. Microbiology. 150:3025-3033.

589

37.

Ollinger J, Bowen B, Wiedmann M, Boor KJ, Bergholz TM. 2009. Listeria

590

monocytogenes sigmaB modulates PrfA-mediated virulence factor expression. Infect.

591

Immun. 77:2113-2124.

592 593

38.

Fay MP, Shaw PA. 2010. Exact and asymptotic weighted logrank tests for interval censored data: The interval R package. J. Stat. Softw. 36(2).

27

594

39.

microbial resistance. Nat. Rev. Microbiol. 4:529-536.

595 596

Peschel A, Sahl HG. 2006. The co-evolution of host cationic antimicrobial peptides and

40.

Tang S, Stasiewicz MJ, Wiedmann M, Boor KJ, Bergholz TM. 2013. Efficacy of

597

different antimicrobials on inhibition of Listeria monocytogenes growth in laboratory

598

medium and on cold-smoked salmon. Int. J. Food Microbiol. 165:265-275.

599

41.

Neetoo H, Ye M, Chen H. 2008. Potential antimicrobials to control Listeria

600

monocytogenes in vacuum-packaged cold-smoked salmon pate and fillets. Int. J. Food

601

Microbiol. 123:220-227.

602

42.

van Schaik W, Gahan CG, Hill C. 1999. Acid-adapted Listeria monocytogenes displays

603

enhanced tolerance against the lantibiotics nisin and lacticin 3147. J. Food Prot. 62:536-

604

539.

605

43.

Tessema GT, Moretro T, Snipen L, Heir E, Holck A, Naterstad K, Axelsson L. 2012.

606

Microarray-based transcriptome of Listeria monocytogenes adapted to sublethal

607

concentrations of acetic acid, lactic acid, and hydrochloric acid. Can. J. Microbiol.

608

58:1112-1123.

609

44.

Bowman JP, Lee Chang KJ, Pinfold T, Ross T. 2010. Transcriptomic and phenotypic

610

responses of Listeria monocytogenes strains possessing different growth efficiencies

611

under acidic conditions. Appl. Environ. Microbiol. 76:4836-4850.

612

45.

Stasiewicz MJ, Wiedmann M, Bergholz TM. 2011. The transcriptional response of

613

Listeria monocytogenes during adaptation to growth on lactate and diacetate includes

614

synergistic changes that increase fermentative acetoin production. Appl. Environ.

615

Microbiol. 77:5294-5306.

28

616

46.

Kazmierczak MJ, Wiedmann M, Boor KJ. 2006. Contributions of Listeria

617

monocytogenes sigmaB and PrfA to expression of virulence and stress response genes

618

during extra- and intracellular growth. Microbiology. 152:1827-1838.

619

47.

Kovacs M, Halfmann A, Fedtke I, Heintz M, Peschel A, Vollmer W, Hakenbeck R,

620

Bruckner R. 2006. A functional dlt operon, encoding proteins required for incorporation

621

of d-alanine in teichoic acids in gram-positive bacteria, confers resistance to cationic

622

antimicrobial peptides in Streptococcus pneumoniae. J. Bacteriol. 188:5797-5805.

623

48.

Samant S, Hsu FF, Neyfakh AA, Lee H. 2009. The Bacillus anthracis protein MprF is

624

required for synthesis of lysylphosphatidylglycerols and for resistance to cationic

625

antimicrobial peptides. J. Bacteriol. 191:1311-1319.

626

49.

Peschel A, Jack RW, Otto M, Collins LV, Staubitz P, Nicholson G, Kalbacher H,

627

Nieuwenhuizen WF, Jung G, Tarkowski A, van Kessel KP, van Strijp JA. 2001.

628

Staphylococcus aureus resistance to human defensins and evasion of neutrophil killing

629

via the novel virulence factor MprF is based on modification of membrane lipids with l-

630

lysine. J. Exp. Med. 193:1067-1076.

631

50.

Kramer NE, van Hijum SA, Knol J, Kok J, Kuipers OP. 2006. Transcriptome

632

analysis reveals mechanisms by which Lactococcus lactis acquires nisin resistance.

633

Antimicrob. Agents Chemother. 50:1753-1761.

634

51.

Saar-Dover R, Bitler A, Nezer R, Shmuel-Galia L, Firon A, Shimoni E, Trieu-Cuot

635

P, Shai Y. 2012. D-alanylation of lipoteichoic acids confers resistance to cationic

636

peptides in group B streptococcus by increasing the cell wall density. PLoS Pathog.

637

8:e1002891.

29

638

52.

view. Infection 16 Suppl 2:S92-97.

639 640

53.

Weidenmaier C, Peschel A. 2008. Teichoic acids and related cell-wall glycopolymers in Gram-positive physiology and host interactions. Nat. Rev. Microbiol. 6:276-287.

641 642

Fiedler F. 1988. Biochemistry of the cell surface of Listeria strains: a locating general

54.

Kingston AW, Liao X, Helmann JD. 2013. Contributions of the sigmaW , sigmaM and

643

sigmaX regulons to the lantibiotic resistome of Bacillus subtilis. Mol. Microbiol. 90:502-

644

518.

645

55.

antimicrobial peptides in Clostridium difficile. Microbiology. 157:1457-1465.

646 647

56.

57.

Crandall AD, Montville TJ. 1998. Nisin resistance in Listeria monocytogenes ATCC 700302 is a complex phenotype. Appl. Environ. Microbiol. 64:231-237.

650 651

Ming XT, Daeschel MA. 1993. Nisin resistance of foodborne bacteria and the specific resistance responses of Listeria monocytogenes Scott A. J. Food Prot. 56:944-948.

648 649

McBride SM, Sonenshein AL. 2011. The dlt operon confers resistance to cationic

58.

Li J, Chikindas ML, Ludescher RD, Montville TJ. 2002. Temperature- and surfactant-

652

induced membrane modifications that alter Listeria monocytogenes nisin sensitivity by

653

different mechanisms. Appl. Environ. Microbiol. 68:5904-5910.

654

59.

Mastronicolis SK, Berberi A, Diakogiannis I, Petrova E, Kiaki I, Baltzi T, Xenikakis

655

P. 2010. Alteration of the phospho- or neutral lipid content and fatty acid composition in

656

Listeria monocytogenes due to acid adaptation mechanisms for hydrochloric, acetic and

657

lactic acids at pH 5.5 or benzoic acid at neutral pH. Antonie van Leeuwenhoek. 98:307-

658

316.

659 660

60.

Beales N. 2004. Adaptation of microorganisms to cold temperatures, weak acid preservatives, low pH, and osmotic stress: A review. Compr Rev Food Sci F 3:1-20.

30

661

61.

Zivanovic S, Basurto CC, Chi S, Davidson PM, Weiss J. 2004. Molecular weight of

662

chitosan influences antimicrobial activity in oil-in-water emulsions. J. Food Prot. 67:952-

663

959.

664

62.

Loeffler M, McClements DJ, McLandsborough L, Terjung N, Chang Y, Weiss J.

665

2014. Electrostatic interactions of cationic lauric arginate with anionic polysaccharides

666

affect antimicrobial activity against spoilage yeasts. J. Appl. Microbiol. 117:28-39.

667

63.

Stephenson K, Hoch JA. 2004. Developing inhibitors to selectively target two-

668

component and phosphorelay signal transduction systems of pathogenic microorganisms.

669

Curr. Med. Chem. 11:765-773.

670

64.

systems as therapeutic targets. Curr. Opin. Pharmacol. 2:507-512.

671 672

65.

Bergholz TM, Moreno Switt AI, Wiedmann M. 2014. Omics approaches in food safety: fulfilling the promise? Trends Microbiol. 22:275-281.

673 674

Stephenson K, Hoch JA. 2002. Two-component and phosphorelay signal-transduction

66.

Nelson KE, Fouts DE, Mongodin EF, Ravel J, DeBoy RT, Kolonay JF, Rasko DA,

675

Angiuoli SV, Gill SR, Paulsen IT, Peterson J, White O, Nelson WC, Nierman W,

676

Beanan MJ, Brinkac LM, Daugherty SC, Dodson RJ, Durkin AS, Madupu R, Haft

677

DH, Selengut J, Van Aken S, Khouri H, Fedorova N, Forberger H, Tran B,

678

Kathariou S, Wonderling LD, Uhlich GA, Bayles DO, Luchansky JB, Fraser CM.

679

2004. Whole genome comparisons of serotype 4b and 1/2a strains of the foodborne

680

pathogen Listeria monocytogenes reveal new insights into the core genome components

681

of this species. Nucleic Acids Res. 32:2386-2395.

31

682

67.

Bergholz TM, Bowen B, Wiedmann M, Boor KJ. 2012. Listeria monocytogenes shows

683

temperature-dependent and -independent responses to salt stress, including responses that

684

induce cross-protection against other stresses. Appl. Environ. Microbiol. 78:2602-2612.

685

68.

Lauer P, Chow MY, Loessner MJ, Portnoy DA, Calendar R. 2002. Construction,

686

characterization, and use of two Listeria monocytogenes site-specific phage integration

687

vectors. J. Bacteriol. 184:4177-4186.

688 689 690

32

691

Table 1. Strains and plasmids used in this study. Strain / Plasmid Strains FSL F6-0366 FSL B2-0315 FSL B2-0377 FSL B2-0379 FSL B2-0380 FSL K5-0022 FSL K5-0024 FSL K5-0025 FSL K5-0026 FSL K5-0027 FSL K5-0028 FSL K5-0029 Plasmids pJK1 pPL2

Strain Alias / Relevant Genotype

Reference

H7858; parent strain; serotype 4b H7858 ΔliaR H7858 ΔlisR H7858 ΔvirR H7858 ΔcesR H7858 ΔvirS Pdlt-GUS FSL F6-0366 Pdlt-GUS FSL B2-0379 (ΔvirR) Pdlt-GUS FSL K5-0022 (ΔvirS) H7858 ΔdltA H7858 ΔmprF H7858 ΔanrAB

(66) (67) This work This work This work This work This work This work This work This work This work This work

Pdlt-uidA(GUS) on pPL2 Integrative shuttle vector (Cmr)

This work (68)

692 693

33

694 695

Table 2. MIC determinations for the parent strain (H7858) and its isogenic TCS response

696

regulator mutantsa,b

Antimicrobial (µg/ml) NIS LAE EPL CHI

H7858 12.5 12.5 100 >200

ΔliaR 6.25 12.5 50 >200

MIC for ΔlisR 12.5 6.25 100 >200

ΔvirR 1.56 6.25 25 100

ΔcesR 12.5 12.5 50 >200

697 698 699 700

a

MICs are determined from three biological replicates and represented as the mode.

b

MICs in bold are statistically different compared to the parent strain with p value = 0.1 as

determined by the logrank test.

701 702

34

703

FIGURE LEGENDS

704

Figure 1. Membrane integrity of the parent strain H7858 and ΔvirR strain under NIS, LAE, EPL,

705

CHI exposure for 24 h at 7oC. Black and gray bars represent the relative dye ratio of the parent

706

strain and ΔvirR, respectively. The relative dye ratios are calculated as the fluorescence ratio

707

(SYTO9 to propidium iodide) for antimicrobial-treated parent strain and ΔvirR divided by the

708

fluorescence ratio of untreated parent strain and ΔvirR. Asterisks indicate the relative dye ratios

709

for ΔvirR under each antimicrobial stress that are significantly (p < 0.05) different compared to

710

the parent strain. The data represent the mean and the standard deviation of three biological

711

replicates.

712

Figure 2. Log decrease in cell density for the parent strain H7858, ΔvirR, ΔdltA, ΔmprF, and

713

ΔanrAB exposed to NIS (0.5 µg/ml), LAE (20 µg/ml), EPL (500 µg/ml), and CHI (50 µg/ml) for

714

24 h at 7oC. Within each antimicrobial, strains with the same letter are not significantly different

715

from each other (p < 0.05). The data represent the mean and the standard deviation of three

716

biological replicates.

717

Figure 3. Relative binding of cytochrome c for the parent strain H7858, ΔvirR, ΔdltA, and

718

ΔmprF grown to log phase at 7oC. Strains with the same letter are not significantly different from

719

each other (p < 0.05). The data represent the mean and the standard deviation of three biological

720

replicates.

721

Figure 4. GUS activities for dlt promoter-GUS reporter fusions expressed in the parent strain,

722

ΔvirR, ΔvirS backgrounds in the presence of 100ng/ml NIS at 7oC. Asterisks indicate

723

significantly (p < 0.05) different GUS activity levels in each genetic background. The data

724

represent the mean and the standard deviation of three biological replicates. 35

725

Figure 5. Log decrease in cell density for the parent strain H7858 and ΔvirR strains pre-exposed

726

to no acid (CTRL), 2% potassium lactate (PL), and 0.14% sodium diacetate (SD) for 8 h at 7oC

727

prior to 24 h challenge with NIS, LAE, EPL, and CHI at 7oC. Closed circles and open circles

728

represent the log decrease in cell density for the parent strain and ΔvirR strains, respectively.

729

Asterisks above and below the dots indicate the log decrease in cell numbers for acid-exposed

730

parent strain and ΔvirR, respectively, that are significantly (p < 0.05) different compared to no

731

prior acid exposure (CTRL). The data represent the mean and the standard deviation of three

732

biological replicates.

733

Figure 6. Model of VirR-mediated resistance to food antimicrobials based on our data and

734

previously proposed modes of action for these antimicrobials. The figure was adapted from a

735

previous publication (53). The L. monocytogenes cell envelope consisting of the cell wall (gray)

736

and the cytoplasmic membrane (below gray area) is shown. Teichoic acids that are anchored to

737

the cell wall and the cytoplasmic membrane represent wall teichoic acids and lipoteichoic acids,

738

respectively. D-alanylation of teichoic acids (mediated by DltABCD) and L-lysinylation of

739

membrane phospholipids (mediated by MprF) are indicated by the letter A or L in a circle,

740

respectively. Letters below each strain designate the phenotype when exposed to each

741

antimicrobial as inferred from the survival assays: R, resistant; S, sensitive; MS, most sensitive.

742

Potential electrostatic repulsion or exclusion of antimicrobials facilitated by teichoic acid as well

743

as phospholipid modifications are indicated by deflected arrows with a horizontal line whereas a

744

plain arrow indicates antimicrobial penetration to the target site. The combination of two arrows

745

indicates partial resistance provided by the specific modification.

746

36

VirR-Mediated Resistance of Listeria monocytogenes against Food Antimicrobials and Cross-Protection Induced by Exposure to Organic Acid Salts.

Formulations of ready-to-eat (RTE) foods with antimicrobial compounds constitute an important safety measure against foodborne pathogens such as Liste...
909KB Sizes 32 Downloads 5 Views