JB Accepted Manuscript Posted Online 2 March 2015 J. Bacteriol. doi:10.1128/JB.02409-14 Copyright © 2015, American Society for Microbiology. All Rights Reserved.

1

2

The mechanism for inhibition of Vibrio cholerae ToxT activity by the unsaturated fatty acid

3

components of bile

4

5

Sarah C. Plecha and Jeffrey H. Withey#

6

Department of Immunology and Microbiology

7

Wayne State University School of Medicine, Detroit, MI 48201

8

9

10

# To whom correspondence should be addressed: Jeffrey H. Withey, Department of

11

Immunology and Microbiology, Wayne State University, 540 E. Canfield, Detroit, MI 48201,

12

USA, Tel.: (313) 577-1316; Fax: (313) 577-1155; E-mail: [email protected]

13

14

15

16

Running Title: Mechanism for ToxT inhibition by bile

17

1

18

ABSTRACT

19

The gram-negative curved bacillus Vibrio cholerae causes the severe diarrheal illness

20

cholera. During host infection, a complex regulatory cascade results in production of ToxT, a

21

DNA-binding protein that activates the transcription of major virulence genes that encode

22

cholera toxin (CT) and toxin-coregulated pilus (TCP). Previous studies have shown that bile and

23

its unsaturated fatty acid (UFA) components reduce virulence gene expression and therefore

24

are likely important signals upon entering the host. However, the mechanism for the bile-

25

mediated reduction of TCP and CT expression has not been clearly defined. There are two likely

26

hypotheses to explain this reduction: 1) UFAs decrease DNA binding by ToxT or 2) UFAs

27

decrease dimerization of ToxT. The work presented here elucidates that bile/UFAs directly

28

affect DNA binding by ToxT. UFAs, specifically linoleic acid, can enter V. cholerae when added

29

exogenously and are present in the cytoplasm where they can then interact with ToxT.

30

Electrophoretic mobility shift assays (EMSAs) with ToxT and various virulence promoters in the

31

presence or absence of UFAs showed a direct reduction in ToxT binding to DNA, even in

32

promoters with only one ToxT binding site. Virstatin, a synthetic ToxT inhibitor, was previously

33

shown to reduce ToxT dimerization. Here we show that virstatin only affects DNA binding at

34

ToxT promoters with two binding sites, unlike linoleic acid which affects ToxT binding

35

promoters having either one or two ToxT binding site. This suggests a mechanism in which

36

UFAs, unlike virstatin, do not affect dimerization but affect monomeric ToxT binding to DNA.

37

38

IMPORTANCE Vibrio cholerae must produce the major virulence factors cholera toxin (CT) and toxin2

39

coregulated pilus (TCP) to cause cholera. CT and TCP production depends on ToxT, the major

40

virulence transcription activator. ToxT activity is negatively regulated by unsaturated fatty acids

41

(UFA) present in the lumen of the upper small intestine. This study investigates the mechanism

42

for inhibition of ToxT activity by UFAs and finds that UFAs directly reduce specific ToxT binding to

43

DNA at virulence promoters and subsequently reduce virulence gene expression. UFAs inhibit

44

ToxT monomers from binding DNA. This differs from the inhibitory mechanism of a synthetic

45

ToxT inhibitor, virstatin, which inhibits ToxT dimerization. Understanding the mechanisms for

46

inhibition of virulence could lead toward better cholera therapeutics. INTRODUCTION

47

48

V. cholerae, a gram-negative curved bacillus possessing a single polar flagellum, is the

49

causative agent of cholera. Cholera is a diarrheal disease contracted by consuming

50

contaminated food or water and is characterized by severe diarrhea that leads to dehydration,

51

and can ultimately cause death if left untreated. Each year, there are an estimated 1.4-4.3

52

million cholera cases and 20,000-142,000 deaths from the disease (1, 2). To cause disease, V.

53

cholerae must colonize the upper small intestine, where it expresses virulence genes including

54

those that encode the two most important virulence factors: toxin co-regulated pilus (TCP) and

55

cholerae toxin (CT). TCP is required for intestinal colonization, while CT is responsible for the

56

massive secretion of electrolytes and water into the lumen, causing diarrhea. Transcription of

57

these and other virulence genes are regulated by the major virulence transcription activator

58

ToxT.

59

Virulence gene expression is regulated by what is historically known as the ToxR regulon

3

60

(3, 4). toxT transcription is activated by two pairs of inner membrane proteins, ToxR/ToxS and

61

TcpP/TcpH, that bind upstream of toxT, as well as by a ToxT positive feedback loop (5-9). ToxT is

62

a 32-kDA member of the AraC/XylS transcriptional regulator family, having a conserved 100-

63

amino-acid DNA binding domain, consisting of two helix-turn-helix motifs, in its C-terminal

64

domain (CTD).

65

sequences called toxboxes that vary in configuration at individual genes (10-14). ToxT directly

66

controls transcription of genes encoding not only TCP and CT, but also other accessory virulence

67

genes including acfA, acfD, aldA, tagA, tarA, tarB, and tcpI (11-13, 15, 16). Except for the aldA

68

promoter, in which there is a single toxbox, there are two toxboxes present at ToxT-activated

69

promoters (11-14, 16). ToxT can bind to single toxboxes as a monomer but it is thought that full

70

activation only occurs upon ToxT dimerization on the DNA, at least at some genes (12, 13). The

71

ToxT N-terminal domain (NTD), does not share significant sequence homology with other

72

proteins, but has some structural similarity to the AraC NTD, which is the domain necessary for

73

AraC dimerization and arabinose binding (17). NTDs of ToxT have been shown to interact when

74

separated from the CTD, although the true role of dimerization in ToxT function is not fully

75

understood (16-22).

Located upstream of all ToxT-activated genes are 13-bp degenerate DNA

76

There are a number of host and environmental factors that affect ToxT activity, including

77

temperature, pH, bile and bicarbonate (23-27). Bile is produced by the liver and then

78

subsequently stored in the gallbladder. Upon eating, bile is released into the duodenum where

79

it acts to solubilize lipids. Bile itself is a complex, heterogeneous mixture that includes bile salts,

80

cholesterol, bilirubin, and saturated and unsaturated fatty acids (UFAs). V. cholerae encounters

81

bile early during infection and it is proposed to be a natural effector of ToxT as it is found in the 4

82

same places that V. cholerae colonizes (24, 25, 27). V. cholerae has reduced virulence gene

83

expression in the presence of bile and/or UFAs, with increased motility gene expression, biofilm

84

formation, the induction of efflux pumps, and increased amounts of outer membrane proteins

85

OmpU, OmpT and TolC (24, 25, 28-31). Bile and its UFA components have been previously

86

shown to decrease CT and TCP production, but the mechanism of this effect on virulence gene

87

expression levels is not fully understood (24, 25). Bile also has another role in that it is able to

88

act as a bactericide for the benefit of the host, but enteric bacteria, such as V. cholerae, are not

89

only adapted to live in the presence of bile, but potentially can recognize bile as a signal to

90

ensure survival in the host (24, 27). Another ToxT inhibitor is a synthetic molecule called

91

virstatin (21, 22). The mechanism of action for virstatin has been proposed to be through

92

inhibition of ToxT dimerization (21). Whether UFAs and virstatin inhibit ToxT activity via the

93

same mechanism is unknown.

94

In this study we show that UFAs, specifically the unsaturated fatty acid linoleic acid, are

95

able to enter the bacterial cytoplasm where ToxT is also present. We report that linoleic acid

96

causes decreased transcription of ToxT-controlled virulence genes assessed. The mechanism by

97

which this occurs is decreased binding affinity at ToxT-activated promoters in the presence of

98

linoleic acid, regardless of the number or configuration of the toxboxes. Our results indicating

99

that UFAs decrease ToxT binding to DNA even at promoters having one ToxT binding site

100

suggest a mechanism in which UFAs affect the ability of monomeric ToxT to bind to DNA, in

101

contrast to the results we observed using virstatin. Thus these two ToxT inhibitors exhibit

102

somewhat different modes of action. These results give us a clearer understanding of the

103

regulatory networks controlling virulence gene expression during human infection. 5

MATERIALS AND METHODS

104 105

V. cholerae strains and growth conditions. All V. cholerae strains used in the study are

106

derived from classical biotype O395. Strains were maintained in Luria Broth (LB) containing

107

20% glycerol and stored at -70°C. All promoter::lacZ fusions used for β-galactosidase assays

108

were made in plasmid pTL61T as described in previous studies (26, 32). Overnight cultures were

109

grown for ~16 hours at 37°C in LB and then diluted 1:40 into LB pH 6.5 at 30°C for virulence

110

inducing conditions in the presence or absence of freshly prepared .05% bile (sodium choleate),

111

160µM linoleic acid, or 100µM virstatin. V. cholerae strains were grown with antibiotic

112

concentration of 100µg/mL streptomycin and strains with the plasmid pTL61T were grown with

113

100µg/mL ampicillin.

114

14

C linoleic acid uptake. V. cholerae classical biotype strain O395 was grown overnight in

115

LB at 37°C, subcultured 1:40 in LB pH 6.5, and grown for two hours in the absence of linoleic

116

acid. At two hours, 0.1μCi

117

added for each milliliter of the subculture. Upon addition of the radiolabeled linoleic acid, 1 mL

118

of the culture was immediately centrifuged, and the supernatant saved to compare to cell

119

pellet cpms. The cell pellet was washed 3 times with 1 mL of PBS and centrifuged each time.

120

Cell pellets were resuspended in 100μL PBS and added to 5mL Scintillation cocktail (Fisher

121

Scientific). The same procedure was followed for other aliquots of the subculture at times 5, 15

122

and 30 minutes. After uptake, cpm was measured for each time point using an LS6000IC liquid

123

scintillation counting system (Beckman).

124 125

14

C-radiolabeled linoleic acid (58.2mCi/mmol) (Perkin-Elmer) was

14

C linoleic acid fractionation assay. After overnight growth, V. cholerae O395 classical

biotype was subcultured 1:40 in the absence of linoleic acid. After 2 hours, 0.1μCi

14

C-

6

126

radiolabeled linoleic acid (58.2mCi/mmol) (Perkin-Elmer) was added to one milliliter of the

127

subculture and incubated at room temperature. After one hour, the bacteria were harvested by

128

centrifugation and washed three times in PBS. Bacteria were then resuspended in 20mM Tris-

129

HCl, pH 8.5 and 500 mM NaCl) and freeze-thawed in an ethanol and dry ice bath, then thawed

130

at 37°C. This process was repeated for a total of three freeze-thaw cycles. The bacteria were

131

then fractionated by centrifugation for 10 min at 15,000 x g to separate the membrane and

132

cytoplasm. The cytoplasm was taken as the supernatant and the pellet was resuspended in

133

PBS. Each fraction was then measured for cpm of each fraction using a LS6000IC liquid

134

scintillation counting system (Beckman).

135

Western blot analysis. Immunodetection of ToxT was performed using the same

136

procedure as previously described (26). Briefly, bacteria were harvested after inducing

137

conditions grown with DMSO only, 160µM linoleic acid, or 100µM virstatin.

138

normalized at an optical density of 600nM and resuspended in 2x protein buffer. Samples were

139

then boiled, separated by 14% SDS-PAGE, blotted, and blocked for 30 minutes in TBS-Tween

140

buffer. After blocking, the blots were incubated overnight in a 1:2,500 dilution of rabbit

141

polyclonal anti-ToxT serum, washed three times in TBS buffer, and secondary goat anti-rabbit

142

IgG conjugated to alkaline phosphatase (IgG-AP) was used at a dilution of 1:5000 (Southern

143

Biotech). Blots were washed again and then developed using immune-BCIP (MP Biomedicals).

Cells were

144

To ensure complete fractionation, immunodetection of cytoplasmic ToxT and the outer

145

membrane protein OmpU was performed. After bacteria was separated into cytoplasmic and

146

membrane fractions, with whole cell lysate used as a control, a 14% SDS-PAGE gel was run,

147

blotted, and blocked as described above. The membrane was then incubated overnight in anti7

148

ToxT serum, to detect the cytoplasmic fraction, and 1:500 dilution of goat polyclonal anti-

149

OmpU serum to detect the membrane fraction. Blots were washed three times in TBS-Tween

150

buffer, and secondary rabbit anti-goat IgG-AP (to detect OmpU) followed by secondary goat

151

anti-rabbit IgG-AP (to detect ToxT) were used. Blots were washed again and then developed

152

using immune-BCIP.

153

β-galactosidase assays. Strains to be analyzed were grown overnight in LB at 37°C and

154

then subsequently subcultured 1:40 into LB pH 6.5 in the presence or absence of 160µM

155

linoleic acid dissolved in DMSO. Cultures were grown under virulence inducing conditions

156

(aeration at 30°C) for 3 hours and then analyzed.

157

performed following the same established protocol (33).

The β-galactosidase assay was then

158

qPCR. As in previous assays, after overnight growth, V. cholerae classical biotype O395

159

was subcultured 1:40, either with or without 160µM linoleic acid. RNA from three biological

160

samples in each condition was extracted using the RNeasy Bacteria Protect Mini Kit (Qiagen)

161

and manufacturer’s protocols were followed. DNA contamination was removed using an on-

162

column RNase-free DNase kit (Qiagen) and confirmed free of DNA by the absence of bands

163

using logarithmic PCR. To measure the relative mRNA levels of tcpA and aldA the following

164

primers were used: forward tcpA (5’-ACGCAAATGCTGCTACACAG-3’) reverse tcpA (5’-

165

CCCCTACGCTTGTAACCAAA-3’), and forward aldA (5’-TTGGTGGGCATCCTAACAAT-3’) reverse

166

aldA (5’-ACACCGGCACCTAAACCATA-3’). qRT-PCR was performed using one-step SYBR green

167

Mastermix (Invitrogen) and the following program: cDNA synthesis for 10min. at 55° C,

168

denaturing step for 5min. at 95° C, followed by 35 cycles of 95° C for 10 s, then 55° C for 30 s.

169

The level of each mRNA was normalized to the level of rpoB using primers forward rpoB (5’8

170

ACCTGAAGGTCCAAACATCG-3’) and reverse rpoB (5’-CAAAACCGCCTTCTTCTGTC-3’). Relative

171

levels of transcript with the addition of linoleic were calculated using 2-ΔΔCT and analyzed

172

comparing the ΔΔCT values as previously described (34).

173

Protein Purification. Maltose binding protein-ToxT fusion (MBP-ToxT) and MBP-CRP

174

purification was performed as previously described using E. coli strain BL21(DE3) with the

175

plasmid pMAL-c2E carrying either MBP-ToxT or MBP-CRP (32, 35). Briefly, after MBP-ToxT or

176

MBP-CRP induction by IPTG, cells were lysed by French press, run over an amylose column, and

177

fractions containing MBP-ToxT or MBP-CRP were saved and dialyzed into 50mM Na2HPO4 (pH

178

8.0), and 100mM NaCl and then dialyzed into the same solution containing 20% glycerol to save

179

as freezer stock at -80° C.

180

Electrophoretic mobility shift assays (EMSA). EMSAs were performed as previously

181

described (35). Purified MBP-ToxT, or MBP-CRP with cAMP (New England Biolabs), was

182

incubated with DNA probes made by PCR from the promoter sequence of interest using one

183

primer radiolabeled with γ-32P (Perkin-Elmer) by T4 polynucleotide kinase (New England

184

Biolabs). Binding reactions contained various amounts of MBP-ToxT with constant 10 µg/mL

185

salmon sperm DNA as nonspecific competitor, 10mM Tris-acetate (pH 7.4), 1mM Potassium

186

EDTA (pH 7.0), 100mM KCl, 1mM dithiothreitol (DTT), 0.3mg/mL bovine serum albumin (BSA)

187

and 10% glycerol in a volume of 30µL. To each reaction, a constant concentration of the labeled

188

DNA probe was added. In reactions containing linoleic acid, the final concentration was 32µM

189

for each reaction, and for reactions containing virstatin, the final concentration was 50µM,

190

except at the aldA promoter where a higher concentration of 100µM was used. All other

191

reactions contained 3.33% (1μL in 30μL) DMSO as a solvent control. Binding reactions were 9

192

incubated for 30 minutes at 37°C and then loaded into a 6% polyacrylamide gel at 4°C. Gels

193

were dried for 1 hour and then analyzed by autoradiography.

194

Binding curve analysis. Autoradiographs were analyzed using ImageJ software (NIH) as

195

previously described with nonspecific binding omitted from further analysis (32). Briefly, to

196

determine the Kd for samples containing either linoleic acid or virstatin and compared to

197

protein bound to DNA without inhibitor, the percent of protein bound with labeled DNA was

198

determined

199

%Bound=Bmax*[Protein]h/(Kdh + [Protein]h) with Bmax constraint set to 100 using Graphpad Prism

200

5 software. The Kd values for each condition were compared to each other using the extra sum

201

of squares F test to determine if the two values were statistically different.

202

for

each

lane.

This

was

then

fit

to

the

equation

RESULTS

203

ToxT protein is produced in the presence of linoleic acid. Previous work has

204

demonstrated that UFAs inhibit virulence gene expression by acting through ToxT (25). Oleic

205

acid has been previously shown to negatively regulate TCP and CT expression levels (17). In

206

preliminary experiments, we determined that linoleic acid had the strongest negative effect on

207

ToxT activity of several different UFAs tested (data not shown), and previous studies have

208

shown that linoleic acid makes up the largest proportion of unsaturated fatty acid content in

209

human bile (36). To date, direct interaction of linoleic acid and ToxT has not been observed,

210

although a palmitoleic acid molecule was visualized within the ToxT NTD in the solved crystal

211

structure (17). To begin our study, we investigated whether ToxT protein production was

212

affected by the addition of linoleic acid in the media. It had been previously determined that

213

there are comparable levels of toxT expression in cultures grown with and without linoleic acid 10

214

(25). To confirm this, we analyzed the β-galactosidase production from a toxT::lacZ reporter

215

plasmid in V. cholerae grown in the presence or absence of linoleic acid (Figure 1A). When V.

216

cholerae was grown under virulence-inducing conditions (LB pH=6.5, aeration) in the presence

217

of linoleic acid, the amount of β-galactosidase activity indicated that there was no statistically

218

significant difference (P=.799) in toxT expression whether the culture had linoleic acid or lacked

219

it. This is consistent with previous data confirming that linoleic acid does not

220

affect toxT transcription.

221

To further examine the effect of linoleic acid on ToxT, and confirm that there were no

222

effects of linoleic acid on translation, we carried out Western blot analysis to assess the ToxT

223

protein levels (Figure 1B). V. cholerae was grown under virulence-inducing conditions in the

224

presence or absence of linoleic acid and cell extracts were harvested. ToxT-specific polyclonal

225

antibodies were used to detect protein levels. No differences in ToxT levels were observed

226

regardless of the presence or absence of linoleic acid, as a ToxT-specific band was visible in

227

Western blots at the same intensity under both conditions (Figure 1B, lanes 4 and 5). This band

228

was not detected in the ΔtoxT control strain (Figure 1B, lane 3). Thus, ToxT was stably

229

produced regardless of the presence of linoleic acid while bacteria were grown under virulence-

230

inducing conditions.

231

Linoleic acid enters the V. cholerae cytoplasm. As ToxT protein levels were roughly

232

equivalent with or without linoleic acid, we next assessed whether a direct interaction between

233

linoleic acid and ToxT would be possible. Such an interaction would require that linoleic acid be

234

imported into the bacterial cytosol. To determine whether this occurs, we performed an uptake

235

assay using radiolabeled linoleic acid (Fig. 2). V. cholerae was subcultured under virulence11

236

inducing conditions for two hours and then 14C-labeled linoleic acid was added to the culture.

237

Aliquots of the culture were taken at times 0, 5, 15, and 30 minutes after addition of

238

radiolabeled linoleic acid and radioactivity was quantified by a scintillation counter. The amount

239

of radioactivity found in the cell pellet was compared to the supernatant. Results from this

240

assay show that linoleic acid enters the bacteria with a best fit equation for one-phase

241

association of y = 24622 + 156786*(1 - e-.036x) and an R2-value of .989.

242

These data, however, do not indicate whether linoleic acid is present inside the

243

bacterium or simply associated with the cell surface or outer membrane. To determine whether

244

linoleic acid enters the cytosol,

245

grown under virulence-inducing conditions and then the bacteria were fractionated into

246

membrane and cytoplasmic portions. This allowed us to see where in the bacterium linoleic

247

acid was localized: the membrane, the cytoplasm, or both. Cultures were incubated for one

248

hour with the radiolabeled linoleic acid. These cultures were then pelleted and washed and the

249

bacteria were separated into a cytoplasm and envelope fraction. The 14C-linoleic acid in each

250

fraction was then quantified using a scintillation counter. Results of this experiment show that

251

most of the linoleic acid is able to enter the cell (Figure 3A). A control Western blot was

252

performed to ensure complete fractionation of the bacteria and showed no ToxT in the

253

membrane fraction and no OmpU in the cytoplasmic fraction (Figure 3B). Directly comparing

254

the amounts of linoleic acid in the two fractions showed that about 65% of the total 14C-linoleic

255

acid was present in the cytoplasmic fraction, while 35% was present in the membrane fraction.

256

These data suggest that a significant fraction of extracellular linoleic acid can enter the cytosol,

257

where it could then interact with ToxT.

14

C-labeled linoleic acid was added to V. cholerae cultures

12

258

Linoleic acid decreases ToxT activation of virulence gene transcription. tcpA and ctxAB

259

expression have been previously shown to be inhibited by the addition of oleic acid to the

260

media, but the effects of linoleic acid were not assessed in those studies (17). We assessed

261

whether the inhibitory effect of linoleic acid that we observed extended to a variety of ToxT-

262

activated promoters with various ToxT binding site (toxbox) configurations, including tcpA,

263

ctxAB, acfA, aldA, acfD, tagA, and tcpI. All of these promoters contain two toxboxes in either

264

direct (tcpA,ctxAB) (13) or inverted repeat configurations (acfA, acfD, tagA, tcpI) (12-14) except

265

for aldA, which contains only one toxbox (16). Plasmid-borne promoter-lacZ fusions were

266

added to the wild-type V. cholerae classical strain as well as an otherwise isogenic ΔtoxT strain

267

to measure transcriptional activity at these promoters with and without 160µM linoleic acid

268

(Figure 4). At each of these promoters, linoleic acid had a statistically significant negative effect

269

that prevented ToxT from fully activating transcription at the promoters in wild-type O395. No

270

effect of linoleic was observed in the ΔtoxT strain, confirming that linoleic acid acts through

271

ToxT (Figure 4).

272

As the aldA promoter activity even in the absence of linoleic acid was much lower than

273

at the other promoters, we performed quantitative real-time PCR (qPCR) to compare

274

expression levels of this promoter to expression levels of the tcpA promoter upon the addition

275

of linoleic acid. Bacteria were cultured using the same methods as the β-galactosidase assay

276

and then the RNA was extracted. qPCR was then performed using primers for tcpA, and aldA,

277

while using rpoB as the housekeeping gene control. The qPCR results showed that both tcpA

278

and aldA were downregulated upon the addition of linoleic acid to the media (Figure 5).

13

279

Linoleic acid decreases ToxT-DNA binding affinity. EMSA was used to determine if

280

linoleic acid is able to directly affect ToxT binding to DNA at the ToxT-activated promoters. This

281

method allows the binding pattern of ToxT to be characterized and gives a rough estimation of

282

the equilibrium dissociation constant, Kd, which represents the concentration of ToxT required

283

for binding 50% of the DNA at equilibrium. As the concentration of active ToxT varies from

284

preparation to preparation, the Kd values can only be compared per experimental condition,

285

and are only an estimate. The Kd values with and without linoleic acid give a direct comparison

286

of the effect of the fatty acid on the given promoter, as a higher Kd value indicates a lower

287

binding affinity of ToxT for the DNA and a lower Kd value indicates higher binding affinity.

288

To begin, we used the primary promoter in the tcp operon, PtcpA, as our 32P-labeled DNA

289

probe. Using purified MBP-ToxT, we added increasing amounts of protein to binding reactions

290

containing a constant concentration of labeled DNA probe. Additionally, either DMSO alone

291

(Figure 6A, lanes 1-7) or 32 μM linoleic acid dissolved in DMSO were added to the reactions. A

292

representative EMSA is shown in Fig. 6A. Densitometry analysis was performed to calculate the

293

percentage of bound DNA at each concentration of MBP-ToxT and a binding curve was drawn

294

for each set of reactions to determine the percentage of bound DNA at each concentration of

295

MBP-ToxT (Figure 6B). In the case of MBP-ToxT binding to PtcpA DNA, the Kd without linoleic acid

296

was 9.99 nM and the Kd with linoleic acid was 33.11 nM. These values are significantly different,

297

and indicate a lower ToxT binding affinity for PtcpA in the presence of linoleic acid. These data

298

strongly suggest there is likely a direct interaction between linoleic acid and ToxT.

299

To confirm that linoleic acid does not generally inhibit protein-DNA interactions, we

300

used cyclic AMP receptor protein (CRP) as a control, together with the region upstream of tcpP 14

301

to which CRP binds, as described above. We performed an EMSA comparing CRP binding with

302

and without the addition of linoleic acid (Figure 6C) and performed densitometric analysis to

303

determine the binding curves and Kd values. The binding curve and Kd values were statistically

304

the same for CRP binding to PtcpP with or without linoleic acid (Figure 6D).

305

We next determined the binding curves and Kd values at other ToxT-activated

306

promoters to determine whether different toxbox configurations may impact the effect of

307

linoleic acid on ToxT binding to DNA (Figure 7). Equilibrium binding experiments on PtcpA, which

308

has two direct repeat toxboxes, showed a decrease in DNA binding upon addition of linoleic

309

acid. Other promoters have direct and inverted repeat toxbox configurations, and PaldA has only

310

one toxbox (Figure 7.) If linoleic acid could decrease the binding affinity of ToxT for PaldA, it

311

could indicate the mechanism for ToxT inhibition did not involve effects on dimerization, which

312

had been previously proposed as the mechanism for decreasing virulence gene expression (18).

313

PctxAB contains two directly repeated toxboxes like PtcpA, albeit in the opposite orientation and

314

having different spacing (35). The rest of the ToxT-activated promoters previously mentioned,

315

PacfA, PacfD, PtagA, and PtcpI each contain toxboxes oriented in inverted repeat configuration as

316

well as variable spacing between the toxboxes (12-14).

317

As linoleic acid was shown above to decrease promoter activity for all of these ToxT-

318

activated promoters, we determined whether linoleic acid also affected the DNA binding

319

affinity using EMSA with each promoter region. For PctxAB, containing direct repeats, the Kd

320

without linoleic acid is 4.436 nM, and with linoleic acid is 10.05 nM, indicating decreased

321

binding affinity of ToxT (Figure 7A and 7B). As a representative of the inverted repeat toxbox

322

promoters, we chose to look at PtagA. Here, densitometry and binding curve analysis again 15

323

revealed decreased binding affinity of ToxT for the promoter upon the addition of linoleic acid,

324

with a Kd of 4.078 nM without linoleic acid, and 27.78 nM with linoleic acid (Figure 7C and 7D).

325

Lastly, we examined ToxT binding to the single toxbox promoter, PaldA. Figure 7E shows the

326

autoradiograph of ToxT binding to PaldA with and without linoleic acid. The Kd once again was

327

increased, going from 11.25 nM without linoleic acid to 48.67 nM with linoleic acid (Figure 7F).

328

These data, along with the previously described decrease in promoter activity with the addition

329

of linoleic acid as indicated by the β-galactosidase assays, suggest that linoleic acid decreases

330

binding at each promoter regardless of toxbox configuration and spacing.

331

Another negative ToxT effector, virstatin, acts differently than linoleic acid. The results

332

above suggest a direct negative effect of linoleic acid on ToxT binding to DNA. We next

333

examined whether this was also true for another negative ToxT regulator, virstatin. Previous

334

studies suggested that virstatin negatively affects the ability of ToxT to dimerize (21). To begin,

335

we looked at ToxT levels using Western blot analysis to compare V. cholerae grown with or

336

without virstatin (Figure 1B, lanes 6 and 7). The ToxT levels were unchanged, confirming that

337

ToxT production is not affected by virstatin. We then performed β-galactosidase assays on the

338

previously described ToxT-regulated genes to assess the effects of virstatin (Figure 8). These

339

data show decreased ToxT activity at all promoters except for aldA, where virstatin had no

340

significant effect. A previous study suggested that the aldA promoter was affected by virstatin

341

(21). We then performed EMSAs to determine whether virstatin affected the DNA binding

342

ability of ToxT at tcpA, ctxAB, and aldA, as representatives of the different toxbox

343

configurations (Figure 9). Virstatin decreased binding at tcpA and ctxAB promoters, as indicated

344

by analysis of the binding curve and Kd, (Figure 9A-D) but did not decrease ToxT binding at the 16

345

aldA promoter even at a higher concentration of virstatin (Figure 9E). In fact, the Kd significantly

346

decreased from 21.83nM to 11.79nM with the addition of virstatin, indicating a higher binding

347

affinity of ToxT for the aldA promoter when virstatin is present (Figure 9F).

348

349

350

DISCUSSION

351

The activation of virulence gene expression relies on the major virulence transcription

352

activator, ToxT. Previous studies have looked at the involvement of bile, UFAs, and virstatin on

353

V. cholerae virulence gene expression and ToxT dimerization (17, 21, 22, 24, 25, 27). Most of

354

this work with negative effectors has been done only on tcpA and ctxAB; here, in addition to

355

these genes, we also examined the effects of negative effectors on ToxT activity at aldA, acfA,

356

acfD, tagA, and tcpI. All ToxT-controlled promoters showed a decrease in expression in the

357

presence of linoleic acid. Prior to entry into the mucus layer of epithelial cells and colonization,

358

V. cholerae must pass through the lumen of the small intestine where high concentrations of

359

bile and UFAs are present. Linoleic acid needs to be taken up by V. cholerae and enter the

360

cytoplasm in order to directly interact with ToxT. Here, we show that linoleic acid is able to

361

enter the cell and that the majority of it enters the cytoplasm. One caveat is that there may be

362

periplasmic components present in the cytoplasmic portion in our fractionation studies.

363

However, the observation that a majority of linoleic acid was in this fraction, combined with our

364

other observations that linoleic acid directly impacts DNA binding by ToxT in vitro and ToxT

17

365

activity in vivo, strongly suggest that at least some linoleic acid is present in the bacterial

366

cytosol where it directly interacts with ToxT.

367

Previous work has described the effect of negative ToxT effectors on virulence gene

368

expression, but their direct effect on ToxT DNA binding affinity had not been previously

369

assessed. It has been proposed that the mechanism of action for virstatin and UFAs is inhibition

370

of ToxT dimerization (18, 21). The binding sites for ToxT, toxboxes, and their configurations

371

have been extensively studied (12,13,16,35), but whether toxbox configuration factors into the

372

inhibition of ToxT activity by linoleic acid had not been examined. Here we show that the

373

decrease in binding affinity in response to linoleic acid can be seen at each ToxT-activated

374

promoter that we examined, regardless of toxbox configuration. This suggests that linoleic acid

375

has a general effect on DNA binding by ToxT monomers. Previous work strongly suggested that

376

ToxT binds to DNA as a monomer (12, 13, 16) and that ToxT does not behave as a typical

377

obligatorily dimeric activator, such as AraC.

378

We also looked at virstatin and its method of action on ToxT. A previous study examined

379

virulence gene promoter activity in the presence of virstatin; however, unlike this earlier study,

380

the effect of virstatin in our hands was not as pronounced. We did observe that virstatin

381

significantly decreased promoter activity, but to a lesser extent than previously described

382

(Figure 8 (21)). While Shakhnovich et al. saw, at greatest, 20% promoter activity in the presence

383

of virstatin at the aldA, acfA, and tcpI promoters, we observed a decrease, but not to that

384

degree. Interestingly, we did not see any effect of virstatin on ToxT activation at the single

385

toxbox, aldA promoter with the β-galactosidase assay (Figure 8). In the same vein, virstatin also

18

386

did not affect the binding affinity of ToxT for the aldA promoter (Figure 9E,F). In contrast, we

387

did see a negative effect of linoleic acid on both ToxT-dependent virulence gene promoter

388

activity (Figure 4) and ToxT binding affinity for PaldA, where the Kd increased from 11.25 nM to

389

48.67 nM (Figure 7E,F).

390

These data suggest that virstatin and linoleic acid, although both negative effectors of

391

ToxT activity, do not have the same mechanism for downregulating gene expression. Another

392

study performed bacterial two-hybrid assays to look at the effect of various UFAs on

393

dimerization and the data suggest another UFA, oleic acid, does affect ToxT dimerization (25).

394

However, linoleic acid was not tested and this study was done using only the NTD of ToxT

395

instead of intact, full-length ToxT, whose dimerization has never been experimentally observed,

396

at least in published reports. More testing using full-length ToxT in V. cholerae with this two-

397

hybrid system in the presence of linoleic acid would confirm whether or not linoleic acid is

398

involved in dimerization. As there was reduced binding of ToxT at the aldA promoter, it is likely

399

that linoleic acid changes the structural conformation of ToxT, leading to inhibited DNA binding.

400

Understanding the effect of host signals on bacterial pathogenesis is useful for complete

401

understanding of the V. cholerae virulence cascade, as many host signals are used by the

402

bacteria to sense the appropriate location in which virulence gene transcription should be

403

initiated. When passing through the small intestine, V. cholerae encounters many different

404

chemical signals, such as bile and bicarbonate, which either inhibit or activate ToxT-dependent

405

gene transcription. Bile and its UFA components are at high concentrations in the duodenal

406

lumen, in which ToxT is inactive (20, 25, 27, 37). Once the mucus layer is penetrated, UFA

19

407

concentration decreases, while the bicarbonate concentration increases. At this point, along

408

the epithelial surface, ToxT becomes active, virulence genes are expressed, and V. cholerae can

409

colonize. UFAs such as linoleic acid likely keep ToxT in a form unable to bind tightly to DNA in

410

the lumen, because production of TCP would lead to bacterial aggregation prior to penetration

411

of the mucus layer and colonization of the epithelial surface would become impossible.

412

According to this model, V. cholerae is able to use both negative and positive effectors of

413

virulence in order to identify and colonize its ideal niche. Our findings here show a direct effect

414

of linoleic acid on ToxT, giving useful insight into the mechanisms of ToxT gene regulation.

415

416 417

ACKNOWLEDGEMENTS The authors thank the members of the Withey laboratory for helpful discussions. This work was

418

supported by P.H.S. grants K22AI071011 and R56AI093622 (to J.H.W.), Bill and Melinda Gates

419

Foundation grant OPP1068124 and by Wayne State University funds.

420

20

421

422

423

a)

427 428 429 430 431 432

15000 10000 5000 0

WT 160μM LA

426

n.s.

WT

425

PtoxT::lacZ Expression (Miller Units)

424

b)

20000

433 434 435 436 437 438 439 440 441

FIGURE 1. Effect of linoleic acid on PtoxT::lacZ and ToxT expression. (A). Light gray bars, wildtype V. cholerae grown without linoleic acid; dark gray bars, wild-type V. cholerae grown with 64 µM linoleic acid. Statistical significance was determined by Student's t test. (B). Effect of linoleic acid and virstatin on ToxT protein levels. V. cholerae was grown under virulence inducing conditions (LB pH 6.5, 30°C, aeration) with DMSO, linoleic acid or virstatin (lanes 4, 5, 6, 7). Purified full length ToxT-6His migrates to ~32 kDa (lane 2) and a ΔtoxT strain shows no band (lane 3) were used as controls. Samples were normalized by OD600. The Western blot was probed with anti-ToxT antibody. This is representative of three separate experiments.

442

21

443 444

14

C CPM/OD600

150000

100000

50000

0

0

10

20

30

Time (minutes) 445 446 447 448 449 450

FIGURE 2. C14-radiolabeled linoleic acid is taken up by V. cholerae. Classical strain O395 cells were grown under virulence-inducing conditions for two hours and then the radiolabeled linoleic acid was added at time 0 and analyzed at times 5, 15, and 30 minutes post-addition, measured as CPM per optical density unit at 600nm. The equation of best fit is y = 24622 + 156786*(1 - e-.036x) and an R2-value of .989.

451

22

b) 80

456

60

457 458 459 460 461 462

Percent

455

*

Membrane

a)

Cytoplasm

454

Ladder

453

Whole cell lysate

452

40 20 0

OmpU ToxT

25kDa – % Envelope

% Cytoplasm

1

2

3

4

463 464 465 466 467 468 469 470 471 472 473

FIGURE 3. C14-radiolabeled linoleic acid is able to enter the cell and enter the cytoplasm. (A). After a 2-hour subculture in virulence-inducing conditions, linoleic acid was added to the culture and allowed to sit for one hour. After one hour, V. cholerae was fractionated into an envelope and cytoplasmic portion by multiple freeze-thaw cycles. There is significantly more linoleic acid in the cytoplasmic portion as compared to the envelope portion. Error bars represent +/- standard error mean and statistical significance was determined by Student's t test, *, p

Mechanism for inhibition of Vibrio cholerae ToxT activity by the unsaturated fatty acid components of bile.

The Gram-negative curved bacillus Vibrio cholerae causes the severe diarrheal illness cholera. During host infection, a complex regulatory cascade res...
2MB Sizes 0 Downloads 10 Views