14
JOURNAL OF BACTERIOLOGY, 0021-9193/97/$04.0010 Nov. 1997, p. 6537–6550 Vol. 179, No. 21 Copyright © 1997, American Society for Microbiology Positive Regulation of Shigella flexneri Virulence Genes by Integration Host Factor MEGAN E. PORTER AND CHARLES J. DORMAN* Department of Microbiology, Moyne Institute of Preventive Medicine, University of Dublin, Trinity College, Dublin 2, Republic of Ireland Received 12 June 1997/Accepted 19 August 1997 In Shigella flexneri, expression of the plasmid-encoded virulence genes is regulated via a complex cascade involving DNA topology, specific transactivators, and the nucleoid-associated protein H-NS, which represses transcription under inappropriate environmental conditions. We have investigated the involvement of a second nucleoid-associated protein, integration host factor (IHF), in virulence gene expression. We found that tran- scription of the invasion-specific genes is repressed in a strain harboring an ihfA mutation, particularly on entry into the stationary phase. Expression of the virB gene, whose product is required for the activation of these structural genes, is also enhanced by IHF in the stationary phase. In contrast, the virF gene, which encodes an activator of virB, is stimulated by IHF in both the logarithmic and early stationary phases of growth, as is another virF-regulated gene, icsA. We have identified regions of the virF, virB, and icsA promoters which form IHF-dependent protein-DNA complexes in vitro and have located sequences within these regions with similarity to the consensus IHF binding site. Moreover, results from experiments in which the virF or virB gene was expressed constitutively confirm that IHF has a direct input at the level of both virF and virB transcription. Finally, we provide evidence that at the latter promoter, the primary role of IHF may be to overcome repression by the H-NS protein. To our knowledge, this is the first report of a role for IHF in controlling gene expression in S. flexneri. The facultative intracellular pathogen Shigella flexneri is the causative agent of bacillary dysentery. Its ability to cause dis- ease is dependent upon the presence of a 230-kb plasmid, of which a 31-kb region encodes all proteins required for invasion of host epithelial cells (reviewed in reference 23). Within this region, one DNA strand encodes the Ipa invasins and IcsB, required for inter- and intracellular spreading, while upstream, and transcribed divergently, are the mxi and spa operons, which are responsible for surface presentation and secretion of the invasion proteins by a type III secretion mechanism (2–5, 47, 55) (Fig. 1). Expression of these genes is controlled by multiple environmental stimuli acting through a regulatory cas- cade involving proteins encoded by both the high-molecular- weight plasmid and the chromosome (7, 23, 27, 37, 42). The primary regulator VirF, an AraC-like protein, is a positive regulator of a second regulatory gene, virB. The product of the virB gene in turn activates the transcription of the structural genes described above (1, 50, 51). Additionally, VirF activates an unlinked gene, icsA, required for intercellular spreading (6, 30, 46). Both virF and virB are plasmid encoded, and while a MalE-VirF fusion protein has been shown to bind to the 217 to 2117 region of the virB promoter in vitro (51), no DNA binding activity of VirB has yet been demonstrated. Transcrip- tional control by environmental stimuli occurs at each level of the regulatory cascade; each gene has an individual response to factors such as temperature, pH, and osmolarity, with virF being the least regulated and the structural genes being more stringently controlled (42). Expression of the virulence gene regulon can be induced by growing the bacteria at 37°C in medium with an osmolarity similar to that of physiological saline, whereas repression occurs at 30°C or in low-osmolarity medium (26, 32, 40). Various environmental stimuli including those which mod- ulate virulence gene expression in S. flexneri—growth temper- ature, osmotic changes, and pH—can induce alterations in DNA topology that affect the transcription of many bacterial genes (24, 28), and thermally induced supercoiling changes have been shown to be important for the coordinated control of virulence gene expression in S. flexneri (17, 52). More spe- cifically, the virF, virB, and icsB promoters are known to be sensitive to the level of DNA superhelicity, and differential sensitivities to DNA supercoiling among the individual pro- moters of the virulence gene cascade may account for the nonuniform response of each gene to environmental stimuli (42, 51, 52). Repression of the S. flexneri virulence regulon by the chro- mosomally encoded protein H-NS under conditions of low temperature and low osmolarity is well established (27, 33, 40). H-NS usually mediates its effects by acting as a transcriptional repressor, often binding close to the promoter region (18). Such binding of H-NS to the 220 to 120 region of the virB promoter has been demonstrated in vitro (51). This site over- laps the binding site for RNA polymerase, and therefore H-NS probably acts by inhibiting DNA contact by the polymerase. More recent results indicate that H-NS also negatively regu- lates the virF and icsA genes (10, 12, 42). H-NS belongs to a class of proteins known as the bacterial histone-like or nucleoid-associated proteins. In this study, we have investigated the contribution of a second histone-like protein, integration host factor (IHF), to the regulation of S. flexneri virulence genes. IHF is a heterodimeric protein con- sisting of an a and a b subunit, encoded by the ihfA and ihfB genes, respectively (see references 20 and 22 for reviews). Upon interaction with its target DNA, IHF introduces a sharp bend of more than 160° in the DNA (45). In contrast to H-NS, which has no consensus binding sequence, a single IHF het- erodimer has been shown to bind to a region of approximately 35 bp, which contains a 13-bp consensus sequence (11, 21). * Corresponding author. Phone: 353 1 608 2013. Fax: 353 1 679 9294. E-mail: [email protected]. 6537 on July 1, 2020 by guest http://jb.asm.org/ Downloaded from

Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

JOURNAL OF BACTERIOLOGY,0021-9193/97/$04.0010

Nov. 1997, p. 6537–6550 Vol. 179, No. 21

Copyright © 1997, American Society for Microbiology

Positive Regulation of Shigella flexneri Virulence Genesby Integration Host Factor

MEGAN E. PORTER AND CHARLES J. DORMAN*

Department of Microbiology, Moyne Institute of Preventive Medicine,University of Dublin, Trinity College, Dublin 2, Republic of Ireland

Received 12 June 1997/Accepted 19 August 1997

In Shigella flexneri, expression of the plasmid-encoded virulence genes is regulated via a complex cascadeinvolving DNA topology, specific transactivators, and the nucleoid-associated protein H-NS, which repressestranscription under inappropriate environmental conditions. We have investigated the involvement of a secondnucleoid-associated protein, integration host factor (IHF), in virulence gene expression. We found that tran-scription of the invasion-specific genes is repressed in a strain harboring an ihfA mutation, particularly onentry into the stationary phase. Expression of the virB gene, whose product is required for the activation ofthese structural genes, is also enhanced by IHF in the stationary phase. In contrast, the virF gene, whichencodes an activator of virB, is stimulated by IHF in both the logarithmic and early stationary phases ofgrowth, as is another virF-regulated gene, icsA. We have identified regions of the virF, virB, and icsA promoterswhich form IHF-dependent protein-DNA complexes in vitro and have located sequences within these regionswith similarity to the consensus IHF binding site. Moreover, results from experiments in which the virF or virBgene was expressed constitutively confirm that IHF has a direct input at the level of both virF and virBtranscription. Finally, we provide evidence that at the latter promoter, the primary role of IHF may be toovercome repression by the H-NS protein. To our knowledge, this is the first report of a role for IHF incontrolling gene expression in S. flexneri.

The facultative intracellular pathogen Shigella flexneri is thecausative agent of bacillary dysentery. Its ability to cause dis-ease is dependent upon the presence of a 230-kb plasmid, ofwhich a 31-kb region encodes all proteins required for invasionof host epithelial cells (reviewed in reference 23). Within thisregion, one DNA strand encodes the Ipa invasins and IcsB,required for inter- and intracellular spreading, while upstream,and transcribed divergently, are the mxi and spa operons,which are responsible for surface presentation and secretion ofthe invasion proteins by a type III secretion mechanism (2–5,47, 55) (Fig. 1). Expression of these genes is controlled bymultiple environmental stimuli acting through a regulatory cas-cade involving proteins encoded by both the high-molecular-weight plasmid and the chromosome (7, 23, 27, 37, 42). Theprimary regulator VirF, an AraC-like protein, is a positiveregulator of a second regulatory gene, virB. The product of thevirB gene in turn activates the transcription of the structuralgenes described above (1, 50, 51). Additionally, VirF activatesan unlinked gene, icsA, required for intercellular spreading (6,30, 46). Both virF and virB are plasmid encoded, and while aMalE-VirF fusion protein has been shown to bind to the 217to 2117 region of the virB promoter in vitro (51), no DNAbinding activity of VirB has yet been demonstrated. Transcrip-tional control by environmental stimuli occurs at each level ofthe regulatory cascade; each gene has an individual response tofactors such as temperature, pH, and osmolarity, with virFbeing the least regulated and the structural genes being morestringently controlled (42). Expression of the virulence generegulon can be induced by growing the bacteria at 37°C inmedium with an osmolarity similar to that of physiologicalsaline, whereas repression occurs at 30°C or in low-osmolaritymedium (26, 32, 40).

Various environmental stimuli including those which mod-ulate virulence gene expression in S. flexneri—growth temper-ature, osmotic changes, and pH—can induce alterations inDNA topology that affect the transcription of many bacterialgenes (24, 28), and thermally induced supercoiling changeshave been shown to be important for the coordinated controlof virulence gene expression in S. flexneri (17, 52). More spe-cifically, the virF, virB, and icsB promoters are known to besensitive to the level of DNA superhelicity, and differentialsensitivities to DNA supercoiling among the individual pro-moters of the virulence gene cascade may account for thenonuniform response of each gene to environmental stimuli(42, 51, 52).

Repression of the S. flexneri virulence regulon by the chro-mosomally encoded protein H-NS under conditions of lowtemperature and low osmolarity is well established (27, 33, 40).H-NS usually mediates its effects by acting as a transcriptionalrepressor, often binding close to the promoter region (18).Such binding of H-NS to the 220 to 120 region of the virBpromoter has been demonstrated in vitro (51). This site over-laps the binding site for RNA polymerase, and therefore H-NSprobably acts by inhibiting DNA contact by the polymerase.More recent results indicate that H-NS also negatively regu-lates the virF and icsA genes (10, 12, 42).

H-NS belongs to a class of proteins known as the bacterialhistone-like or nucleoid-associated proteins. In this study, wehave investigated the contribution of a second histone-likeprotein, integration host factor (IHF), to the regulation of S.flexneri virulence genes. IHF is a heterodimeric protein con-sisting of an a and a b subunit, encoded by the ihfA and ihfBgenes, respectively (see references 20 and 22 for reviews).Upon interaction with its target DNA, IHF introduces a sharpbend of more than 160° in the DNA (45). In contrast to H-NS,which has no consensus binding sequence, a single IHF het-erodimer has been shown to bind to a region of approximately35 bp, which contains a 13-bp consensus sequence (11, 21).

* Corresponding author. Phone: 353 1 608 2013. Fax: 353 1 679 9294.E-mail: [email protected].

6537

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 2: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

Furthermore, both the DNA structure and sequence surround-ing this core site contribute to its binding affinity. Like H-NS,the IHF protein is involved in a number of processes includingreplication, site-specific recombination, and transcriptionalregulation. We have identified a role for IHF in the positiveregulation of S. flexneri virulence gene expression and shownthat it is required for continued expression in the stationaryphase. For the virF and icsA genes, IHF also contributes toexpression during the logarithmic phase, suggesting that themechanism of regulation differs at the individual promoters.Gel mobility shift assays were used to define binding sites forIHF within the virF, virB, and icsA promoter regions. Finally,we provide evidence that the role of IHF at the virB promotermay be to overcome repression by H-NS.

MATERIALS AND METHODS

Strains, plasmids, and growth conditions. The bacterial strains and plasmidsused in this study are listed in Table 1. All strains are derivatives of S. flexneri 2aor Escherichia coli K-12. The ihfAD82::Tn10 allele from CH1480 and theihfA::Tn10 allele from CH1483 (16) were moved by P1cml transduction into S.flexneri BS184 to generate strains CJD1103 and CJD1104, respectively. Trans-ductions were carried out by the method of Silhavy et al. (48). Plasmid pBR322-himA1 (35) was introduced by electroporation into strains CJD1103 (generatingstrain CJD1105) and CJD1104 (generating strain CJD1106); as a control, thecloning vector pBR322 was also introduced into CJD1103 (generating strainCJD1107) and CJD1104 (generating strain CJD1108). To construct a strainharboring insertions in both ihfA and hns, the hns::Tn10 allele was cotransducedwith a linked hnrG::Apr lesion from CJD899 into BS184 (to generate strainCJD1122) and CJD1104 (to generate CJD1123) by P1cml transduction. Apr

transductants were selected, and the presence of the hns::Tn10 insertion wasconfirmed by Southern blotting (data not shown). The hnrG::Apr allele (hns1)was transduced from CJD896 into BS184, resulting in strain CJD1121. Theconstruction of an S. flexneri virF mutant strain, CJD1006, has been describedpreviously (42). To construct a virB mutant derivative of S. flexneri, a 600-bpinternal BamHI-DraI fragment of the virB gene was cloned into the ampicillin-resistant suicide vector pGCS82. The resulting plasmid, pMEP151, was trans-ferred to the tetracycline-resistant S. flexneri CJD1005 by conjugation, as de-scribed previously (42). In colonies where a single homologous recombinationalevent had occurred, integration of the suicide vector pMEP151 resulted in asingle copy of the vector sequences flanked by two partial copies of virB, thusinactivating the virB gene. Bacteriophage P1cml was used to transduce the virBlesion to BS184, resulting in strain CJD1018. The ihfA::Tn10 insertion fromCH1483 was transduced into CJD1006 (creating strain CJD1117) and CJD1018(creating strain CJD1118). To construct plasmids expressing the virF or virBgenes constitutively, plasmid pMEP539 was first constructed by cloning the Cmr

gene from pACYC184 into pBCP378. Plasmid pMEP537 carries the virF geneamplified by PCR with the oligonucleotide primers 59-GCCATACATATGATGGATATGGGACATAAA-39 and 59-CATAGTCGACTTAAAATTTTTTATGATATAAG-39, and cloned into the NdeI-SalI sites of pMEP539. To constructplasmid pMEP538, the virB gene was amplified by PCR with the oligonucleotideprimers 59-GCCATACATATGGTGGATTTGTGCAACGAC-39 and 59-GCCATACATATGTTATGAAGACGATAGATGGCG-39 and then cloned into theNdeI site of pMEP539. Constructs were checked by restriction digestion. Bacte-rial cultures were grown routinely in Luria (L) broth (36) or on L plates (L brothcontaining 1.5% [wt/vol] agar). Where required, antibiotic selection was donewith ampicillin (50 mg/ml), chloramphenicol (15 mg/ml), tetracycline (20 mg/ml),or kanamycin (20 mg/ml).

Enzyme assays. Transcription of the mxiC-lacZ fusion was monitored by b-galactosidase assays of overnight cultures grown at either 30 or 37°C by themethod of Miller (36). Assays were performed in duplicate, and the data wereexpressed as the mean of the two measurements. Standard deviations were lessthan 10%. Experiments were performed at least three times, and typical data areshown.

RNA extraction. Cells (20 ml) grown to the optical densities at 600 nm (OD600)stated in the text, were harvested by centrifugation, added to 1.5 ml of boilingREB (20 mM sodium acetate, 2% sodium dodecyl sulfate, 0.3 M sucrose), andincubated at 100°C for 1.5 min. RNA was isolated by phenol extraction followedby two or three phenol-chloroform (1:1) extractions and one chloroform extrac-tion. Nucleic acids were precipitated with ethanol at 270°C for at least 4 h.Following resuspension in 400 ml of distilled H2O (dH2O), DNA was removedwith RNase-free DNase I (Boehringer Mannheim) as specified by the manufac-turer. The RNA was phenol, phenol-chloroform, and chloroform extracted andprecipitated at 270°C. It was resuspended in dH2O at 5 to 10 mg/ml. The RNAconcentration was determined spectrophotometrically by measuring the OD260,and the integrity of the isolated RNA was confirmed by ethidium bromide gelelectrophoresis.

RNA analysis by Northern blotting. Samples (5 mg [RNA probes] or 20 mg[oligonucleotide probes]) of total cellular RNA were denatured at 65°C, elec-trophoresed on 1% formaldehyde–agarose gels in 3-(N-morpholino)propanesul-fonic acid (MOPS) buffer, and transferred to Hybond-N nylon filters (Amer-sham) as described by the manufacturer. Digoxigenin-labeled antisense RNAprobes specific for virF, virB, icsB, ipgD, and hupA were synthesized by in vitrotranscription of an internal fragment of the respective gene cloned into theSP6/T7 expression vector pSPT18 by using a kit supplied by Boehringer Mann-heim. A synthetic oligonucleotide homologous to the 59 open reading frameof icsA (59-TATTGGCCCCCCGAGCAACAGGGATGCACC-39) was labeledwith digoxigenin-ddUTP by using an end-labeling kit supplied by BoehringerMannheim. Following Northern transfer, probes were hybridized to the immo-bilized RNA and bound probe was detected with the CSPD chemiluminescencedetection system (Boehringer Mannheim) and Hyperfilm-MP X-ray film (Am-ersham). Densitometric analysis was performed with a Bio-Rad model GS-670imaging densitometer and Molecular Analyst software. Each experiment wasperformed at least three times, and typical data are shown.

RNA 5* analysis by primer extension. The icsB and ipgD transcriptional startsites were determined by primer extension analysis with 50 mg of total RNA.Synthetic oligonucleotides homologous to nucleotides 46 to 70 of the icsB openreading frame (59-CTGTATCTTCAACGCGTATAGGCCC-39) and nucleo-tides 44 to 67 of the ipgD open reading frame (59-CTGCGCCTTTATAGGAATCTCCGC-39) were end labeled with 25 mCi of [g-32P]ATP (10 mCi/ml; Amer-sham). Labeled primer (0.3 pmol) was precipitated with the target RNA andresuspended in 30 ml of dH2O. Following incubation at 90°C for 3 min, themixture was allowed to cool slowly to room temperature. Primer extension wasperformed at 42°C for 90 min with 100 U of Superscript RT (Gibco BRL) in thepresence of 40 U of human placental RNase inhibitor (Boehringer Mannheim).RNase A was added, and incubation was continued for a further 10 min at 42°C.cDNA transcripts were extracted, precipitated, resuspended, and then electro-phoresed on 7 M urea–5% polyacrylamide gels next to sequencing markersgenerated by dideoxy sequencing reactions with the same primer and plasmidpMEP73 as a template. The gels were dried and exposed to Hyperfilm-MP X-rayfilm (Amersham). The experiment was repeated twice, and typical data areshown.

Preparation of crude protein extracts. Crude protein extracts were preparedfrom 20-ml cultures of BS184, CJD1104, and CJD1106. Cultures of the respec-tive strains were grown overnight, subinoculated 1:500 into fresh broth, andgrown to an OD600 of 1.5. The cells were harvested by centrifugation andresuspended in 250 ml of cold 10% sucrose–50 mM Tris-HCl (pH 7.5)–100 mMNaCl–1 mM EDTA–5 mM dithiothreitol. Then 7.5 ml of lysozyme (10 mg/ml)was added to each sample. After a 10-min incubation on ice, the mixture was

FIG. 1. Schematic map of the 31-kb virulence region on the S. flexneri 230-kb plasmid and unlinked virulence genes virF and icsA. The orientation of transcriptionis indicated by horizontal arrows, and the locations of promoters (P) are marked. The vertical arrows represent the points of input for the positive regulatory proteinsVirF and VirB. The location of the MudI1734 transcriptional fusion to the mxiC gene is shown by dashed lines. Not drawn to scale.

6538 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 3: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

sonicated and cell debris was pelleted by centrifugation. Finally, each lysate wasmixed with an equal volume of glycerol and stored at 220°C.

Electrophoretic mobility shift assay with crude protein extracts. PCR frag-ments of the various promoters under study were amplified with primers asstated in the text, and DNA was extracted from whole cells of BS184 as atemplate. DNA fragments were labeled by including 30 mCi of [a-32P]dATP inthe PCR mixture, and radioactive probes were purified on a NAP5 column(Pharmacia Biotech) as specified by the manufacturer. Approximately 1 ng of

probe was used for each 20 ml of reaction mixture in 50 mM Tris-HCl (pH8.0)–20 mM KCl–10% (vol/vol) glycerol–100 mg of bovine serum albumin perml–25 mg poly (dI-dC) per ml. Each reaction mixture therefore contained ap-proximately a 500-fold excess of nonspecific synthetic competitor. Followingincubation of the reaction mixture at 28°C for 5 min, 5 ml of crude protein lysate(corresponding to approximately 200 ng of total cell protein) was added and theincubation was allowed to continue for a further 15 min at 28°C. The sampleswere then electrophoresed on a native 4% (wt/vol) polyacylamide gel in 0.53Tris-borate-EDTA (TBE) at 4°C. Radioactive fragments were visualized by au-toradiography with Hyperfilm-MP film (Amersham).

RESULTS

IHF positively regulates the expression of the invasion genemxiC. Previous studies have shown that expression of theS. flexneri invasion genes is affected by the nucleoid-associatedprotein H-NS and DNA topology in addition to the specificregulators VirF and VirB (1, 17, 27, 42). Since the regulationof virulence genes in this bacterium appears to be highly com-plex, it remained possible that other proteins were involved. Asa first step to investigating the potential input of a secondnucleoid-associated protein, IHF, in virulence gene expression,we constructed a strain containing a null mutation in ihfA, thegene encoding the a subunit of IHF. The ihfA::Tn10 allelefrom E. coli CH1483 (16) was introduced into S. flexneri BS184to generate CJD1104 (Table 1). BS184 and its derivatives har-bor a lacZ transcriptional fusion to mxiC, enabling the expres-sion of this invasion gene to be monitored by measuring itsb-galactosidase activity (Fig. 1). Table 2 compares the expres-sion of mxiC in cultures of the ihfA mutant strain CJD1104 andits isogenic wild-type strain, BS184, grown for 12 h. The genesin the regulatory cascade under study are thermally regulatedat the level of transcription, and thermal regulation of mxiCwas maintained in CJD1104; however, there was a reduction inexpression at both 37°C (twofold) and at 30°C (threefold)compared to the wild type. To confirm that the ihfA lesion wasresponsible for the phenotypic effects observed, CJD1104 wastransformed with either a plasmid carrying a cloned copy of theE. coli ihfA gene, pBR322-ihfA1 (35), or the cloning vectorpBR322 to give strains CJD1106 and CJD1108, respectively.When these strains were assayed, it was found that the level ofb-galactosidase activity was restored to that of the wild type inthe presence of pBR322-ihfA1 but not of the cloning vector,confirming that the lack of the IHFa subunit alone was respon-sible for the two- to threefold decrease in mxiC expression.Furthermore, transforming BS184 with either pBR322-ihfA1

or pBR322 had no effect on mxiC expression (data not shown).When the independently isolated ihfAD82::Tn10 allele fromE. coli CH1480 (16) was transduced into BS184 (generatingstrain CJD1103), identical results were obtained (Table 2).These results suggest that IHF plays a small but significant rolein positively regulating mxiC expression.

TABLE 1. Bacterial strains, plasmids, and bacteriophageused in this study

Strain, plasmid,or bacteriophage Relevant detailsa Source or

reference

StrainsS. flexneri 2a

BS184 mxiC::MudI1734 32CJD1005 BS184 pTet 41CJD1006 BS184 virF::pMEP312 41CJD1018 BS184 virB::pMEP151 41CJD1103 BS184 ihfAD82::Tn10 (CH1480 3

BS184)This work

CJD1104 BS184 ihfA::Tn10 (CH1483 3BS184)

This work

CJD1105 CJD1103 pBR322-ihfA1 This workCJD1106 CJD1104 pBR322-ihfA1 This workCJD1107 CJD1103 pBR322 This workCJD1108 CJD1104 pBR322 This workCJD1111 CJD1006 pMEP537 This workCJD1114 CJD1018 pMEP538 This workCJD1115 CJD1006 pMEP539 This workCJD1116 CJD1018 pMEP539 This workCJD1117 CJD1006 ifhA::Tn10 (CH1483 3

CJD1006)This work

CJD1118 CJD1018 ihfA::Tn10 (CH1483 3CJD1018)

This work

CJD1119 CJD1117 pMEP537 This workCJD1120 CJD1118 pMEP538 This workCJD1121 BS184 hnrG::Apr This workCJD1122 BS184 hns-205::Tn10 hnrG::Apr

(CJD899 3 BS184)This work

CJD1123 CJD1104 hns-205::Tn10 hnrG::Apr

(CJD899 3 CJD1104)This work

E. coli K-12CH1480 ihfAD82::Tn10 16CH1483 ihfA::Tn10 16CJD896 hnrG::Apr 19CJD899 hns-205::Tn10 hnrG::Apr 19

PlasmidspBCP378 ColE1 replicon; Apr 54pBR322 ColE1 replicon; Apr Tetr 9pBR322-ihfA1 Functional ihfA gene cloned into

pBR322; Apr35

pGCS82 oriR6K oriTRK2; Apr 29pMEP73 0.6-kb icsB-ipgD promoter region

in pBS PstI-HindIII site; AprThis laboratory

pMEP151 Internal fragment of virB clonedinto pGCS82; Apr

41

pMEP537 virF gene in pMEP539; Cmr Apr This workpMEP538 virB gene in pMEP539; Cmr Apr This workpMEP539 Cmr gene cloned into pBCP378

polylinker; Cmr AprThis work

pSPT18 Low-copy-number plasmid withSP6 and T7 promoters flankingthe pUC18 multicloning site; Apr

Boehringer Mann-heim

pTet pACYC184 with a 339-bp deletionin the Cmr gene; Tetr

41

BacteriophageP1cml P1::Tn9 clr100(Ts) 49

a Transductional crosses are represented by A 3 B, in which A is the donorand B is the recipient. Apr, ampicillin resistance; Cmr, chloramphenicol resis-tance; Tetr, tetracycline resistance.

TABLE 2. Regulation of the mxiC-lacZ transcriptional fusionby IHF in cultures grown overnight at 30 or 37°C

Straina Relevant genotypeb-Galactosidase activityb at:

30°C 37°C

BS184 Wild type 72.7 (1.5) 1,062.4 (5.0)CJD1104 ihfA::Tn10 23.5 (0.27) 509.6 (10.7)CJD1108 ihfA::Tn10 pBR322 18.1 (1.7) 539.5 (27.3)CJD1106 ihfA::Tn10 pBR322-ihfA1 50.4 (1.5) 1,040.3 (7.4)CJD1103 ihfAD82 24.6 (0.56) 522.1 (18.3)CJD1107 ihfAD82 pBR322 23.2 (1.2) 471.1 (15.8)CJD1105 ihfAD82 pBR322-ihfA1 51.0 (2.1) 1,016.0 (25.0)

a See Table 1 for details of the strains.b Values are given in Miller units. Standard deviations are shown in parentheses.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6539

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 4: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

IHF is required for the expression of mxiC in the stationaryphase. The results described above were obtained with bacte-rial cultures grown to the stationary phase, when the expres-sion of mxiC reaches a maximum. Furthermore, it has beenshown previously that the level of IHF increases 5- to 10-foldupon entry into the stationary phase (14), suggesting that reg-ulation by IHF may vary depending upon the stage of growth.By comparing mxiC expression throughout the growth phase,we were able to show that the level of b-galactosidase activityin CJD1104 was only slightly reduced compared to that in thewild-type strain BS184 until an OD600 of ca. 1.0 was reached(Fig. 2). At this point, the expression of mxiC in the ihfAmutant strain CJD1104 reached a plateau, whereas the b-galactosidase activity in the wild type continued to increaseuntil it reached a final level two- to threefold higher than thatin CJD1104. A similar pattern of expression was observed incultures grown at 30 or 37°C, although at 30°C there was aslightly larger difference between the two strains earlier in the

growth phase (compare Fig. 2A and B). Thus, it appears thatIHF is required for the continued expression of mxiC in thestationary phase and that activation by IHF is not temperaturedependent.

Role of IHF in the kinetics of thermal induction of mxiC. Anincrease in temperature, indicative of entry into the humanhost, is the key environmental signal resulting in an invasivephenotype in S. flexneri. Although the above data suggest thatIHF can exert an effect at both 30 and 37°C it remainedpossible that IHF was involved in the induction of virulencegene expression following a thermal upshift from 30 to 37°C.Wild-type strain BS184 and its ihfA::Tn10 mutant derivativeCJD1104 were grown overnight, diluted 1:100 into fresh me-dium, and grown at 30°C until either the early exponential(OD600 5 0.3) or late exponential (OD600 5 0.6) phase beforebeing shifted to 37°C. Culture samples were taken at varioustime points and assayed for b-galactosidase activity. In bothcases, it was found that upon induction, the b-galactosidaseactivity in CJD1104 was reduced compared to that in BS184(Fig. 3). When the thermal upshift occurred early in the growthphase (Fig. 3A) the fold increase in expression over that of theuninduced culture was similar in both strains. In contrast, whenthe shift occurred immediately prior to the onset of stationaryphase, the fold induction was lower in CJD1104 (threefold; 270min after the upshift) than in BS184 (sixfold; 270 min after theupshift) (Fig. 3B). Therefore, IHF contributes to the inductionof virulence gene expression in response to the thermal signal,particularly in cells entering the stationary phase of growth.

Interestingly, while the mxiC fusion could be induced to over1,200 Miller units in the wild-type strain BS184 and to over 600Miller units in the ihfA::Tn10 mutant in the early exponentialphase, it could be induced to only just over 200 Miller units inBS184 or to approximately 100 Miller units in the ihf mutant inlate-exponential-phase cultures (Fig. 3). These data suggestthat the thermal upshift is less effective at inducing the expres-sion of the invasion genes in late-exponential-phase cultures.The reason for this is unknown.

Identification of targets for IHF regulation. The mxiC genestudied in the above experiments is part of a large operon, withthe promoter being situated approximately 11 kb upstream atthe ipgD gene (Fig. 1). Studying the expression of the fusion oflacZ to mxiC in BS184 and its derivatives allows the regulationof these structural genes to be monitored. Their activationrequires the product of the virB gene, which in turn is abso-lutely dependent upon VirF for expression. Since alterations inthe levels of either of these proteins as a result of the ihfA::Tn10 mutation could affect mxiC expression indirectly, we de-cided to use Northern analysis to establish which gene(s) isregulated by IHF. Cultures of BS184 and CJD1104 (ihfA::Tn10), grown overnight, were subinoculated 1:100 into freshmedium and incubated at either 30 or 37°C. Samples weretaken at OD600 of 0.6 and 1.2, and the RNA was extracted andsubjected to Northern blotting to detect transcripts of virF,virB, and the two structural genes ipgD and icsB (Fig. 4). Foreach gene studied, the transcript pattern and thermoregulationin the wild-type strain were the same as observed previously(42).

As predicted from the data on mxiC expression describedabove, transcription of the first gene in the operon, ipgD, wasreduced twofold in the ihfA::Tn10 mutant strain compared tothe wild type in the stationary phase whereas only a slight dropin transcript levels was observed earlier in the growth phase.Furthermore, icsB, which is transcribed divergently from ipgD,showed a similar pattern of IHF dependency. Both the icsBand ipgD promoters require VirB for activation. Interestingly,the virB gene was also regulated twofold by IHF in the station-

FIG. 2. Expression of the mxiC::lacZ transcriptional fusion in response togrowth phase and IHF at 30°C (A) or 37°C (B). Cultures of BS184 (wild type)(solid symbols) and CJD1104 (ihfA::Tn10) (open symbols) grown overnight werediluted 1:100 into fresh medium and grown for 12 h. Samples were taken atregular intervals and assayed for b-galactosidase activity (triangles). OD600 datafor each culture are presented on the same graph (squares).

6540 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 5: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

ary phase, with again no apparent difference in transcript levelsbetween the two strains during the exponential phase, suggest-ing that regulation of the icsB and ipgD promoters by IHF maybe indirect.

Surprisingly, when the same RNA samples were probedfor the virF transcript, a twofold decrease was observed inCJD1104 in both the stationary- and exponential-phase sam-ples, implying that transcription of virF, as well as that of virB,is regulated by IHF. Furthermore, the pattern of regulation isdifferent at the two promoters: virF, but not its subservientgenes, requires IHF for maximum expression throughout thegrowth phase. This also suggests that the effect of IHF at virBmay be direct and not solely a result of its input at the up-

stream virF promoter. An unassociated virulence gene, icsA (orvirG), which is required for intracellular spreading, is alsodependent upon VirF for expression (46). Here there is nointermediary protein; VirF directly activates icsA (1). In anattempt to determine whether icsA is also regulated by IHF, anoligonucleotide specific for the coding region of icsA was usedto probe the same RNA samples used previously. Figure 4shows that icsA is also modulated by IHF in a manner similarto virF, with the transcript levels being reduced in CJD1104during both exponential and stationary phases.

While the transcription of each gene studied was lower at30°C, the fold reduction in expression due to the ihfA::Tn10insertion was the same irrespective of the growth temperature,providing further evidence that regulation by IHF is tempera-ture independent. As a control for concentration and integrity,the RNA was probed for the transcript of the hupA gene,which encodes the a subunit of the histone-like protein HU.No significant variation in hupA transcription was detected(Fig. 4).

IHF dependency can be overcome by expressing VirB con-stitutively. To examine further the possibility that IHF plays anactivating role at both the virF and virB promoters, the effect ofconstitutively expressing virF or virB was examined. StrainCJD1111 contains an inactivated virF gene on the virulenceplasmid and harbors a cloned copy of virF in the expressionplasmid pMEP539. Expression of virF from this plasmid com-plements the virF mutation (Table 3). Introduction of theihfA::Tn10 allele to this strain caused a 1.6-fold reduction inmxiC expression. In strain CJD1114, containing a virB muta-tion and expressing virB from the same expression vector,introduction of the same ihfA mutation caused no reduction inb-galactosidase activity. Constitutive expression of the virBgene eliminated the effect of the ihfA mutation, suggesting thatIHF contributes directly to the expression of virB but onlyindirectly to that of VirB-regulated genes. Constitutive expres-sion of virF merely reduced the effect of an ihfA lesion onmxiC-lacZ expression, probably because of the intact IHF de-pendency of virB in strain CJD1119.

Mapping of the icsB and ipgD transcriptional start sites.The foregoing results suggested that IHF is likely to bind to theregulatory region of virF and virB but not to the icsB-ipgDpromoter region. Before any putative binding sites for IHFwere determined, mapping of the transcriptional start sites forboth icsB and ipgD was required: those for virF, virB, and icsAhave been published previously (30, 37, 51). Total RNA ex-tracted from BS184 grown to the exponential phase (OD600 50.6) was subjected to primer extension analysis with primersspecific for either icsB or ipgD transcripts (see Materials andMethods for details). Polyacrylamide gel electrophoresis of theresulting extension products identified a transcript initiating atthe C residue 52 bp upstream of the icsB coding region (Fig.5A). For ipgD, we detected two transcripts, the first initiating atadjacent A and C residues 44 and 45 bp upstream of thetranslation initiation codon, and the second (minor) initiatingat the C residue 178 bp upstream of the coding region (Fig.5B). For both promoters, potential 210 and 235 sequenceswere found at an appropriate distance upstream of each ofthese transcriptional start sites (Fig. 5C).

IHF binds to sites in the virF, virB, and icsA promoters. Asa first step to determining binding sites for IHF within thepromoter regions of virF, virB, icsA, icsB, and ipgD, the pro-moter-proximal sequences of these genes were searched forsequences with similarity to the IHF consensus binding se-quence 59-WATCAANNNNTTR-39, in which the underlinedCA motif is conserved in all known IHF binding sites (11, 21).For each promoter, several putative binding sites were identi-

FIG. 3. Thermal induction of the mxiC::lacZ transcriptional fusion. Culturesof BS184 (wild type) (solid symbols) and CJD1104 (ihfA::Tn10) (open symbols)grown overnight were diluted 1:100 into fresh medium and grown at 30°C untilan OD600 of 0.3 (A) or 0.6 (B) was reached (Time 5 0). Each culture was thendivided, with half being shifted to 37°C while the other half remained at 30°C.Culture samples were removed at intervals and assayed for b-galactosidaseactivity. Triangles, 37°C; circles, 30°C. Growth curves for each culture at 37°Cfollowing thermal upshift are represented by squares.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6541

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 6: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

fied. To test these for IHF binding, mobility shift assays wereperformed with radiolabeled DNA fragments of the respectivepromoters generated by PCR and crude protein lysates pre-pared from strains BS184, CJD1104, (ihfA::Tn10) andCJD1106 (CJD1104 containing pBR322-ihfA).

In virB, a DNA fragment extending from 272 to 2205 (am-plified with primers BF1 and BBS3 [Fig. 6A]) produced a clearshift when electrophoresed on a polyacrylamide gel followingincubation with equal volumes of protein lysate from eitherBS184 or CJD1106 but not with CJD1104 (Fig. 7). This shift istherefore dependent upon the presence of IHF and suggeststhat IHF binds to a site upstream of the virB promoter. From

analysis of the DNA sequence in this region, the best match tothe IHF consensus sequence (one mismatch) was found at bp2171 to 2183 with respect to the transcriptional start site, asindicated in Fig. 6A. When an adjacent DNA fragment (112 to2167), amplified with primers BBS2 and BF2, was incubatedwith the same crude protein lysates, no DNA-protein interac-tion was observed (Fig. 7) confirming the specificity of theinteraction between IHF and the DNA around nucleotides2171 to 2183.

DNA fragments encompassing the icsB and ipgD promotersidentified above (Fig. 5) were similarly tested for interactionwith IHF in a mobility shift assay. A 228-bp stretch of DNA

FIG. 4. Input of IHF to the S. flexneri virulence regulon. Cultures of BS184 (wild-type) and CJD1104 (ihfA::Tn10) were grown at either 30 or 37°C, and RNA wasextracted at OD600 values of 0.6 and 1.2. (A) Northern blots of the virF, virB, icsB, ipgD, icsA, and hupA (control) transcripts as indicated. The relative levels of eachtranscript are not comparable. (B) The bands in panel A were scanned densitometrically, and the data are expressed as a percentage of the most intense band for eachtranscript. The experiment was repeated three times, and typical data are shown.

6542 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 7: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

closest to the icsB coding region was amplified with primersicsPE and icsK, and the adjacent 244-bp fragment proximal toipgD was amplified with primers ipgPE and ipgK (details ofprimer locations are given in Fig. 5C). Following incubationof either of these fragments with protein lysates produced

from BS184, CJD1104 (ihf::Tn10), and CJD1106 (ihf::Tn10and pBR322-ihfA), no significant IHF-specific interaction wasobserved, suggesting that IHF either does not bind to or hasonly a weak interaction with the structural gene promoters(Fig. 7). This agrees with the above hypothesis that the ob-

FIG. 4—Continued.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6543

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 8: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

served effect of IHF on structural gene expression is due to itsinput at the virB promoter.

It will be recalled that IHF is also required for maximumexpression of the virF and icsA genes. Two fragments in thevicinity of the virF promoter were amplified, and an upstreamregion extending from bp 131 to 2114 (primers NFnt andFBS2 [Fig. 6B]) surprisingly showed no IHF-dependent mo-bility shift (Fig. 7). Similarly, no binding of IHF could be de-tected when sequences up to approximately bp 2200 wereincluded in the probe (data not shown). Instead IHF bound toa 127-bp fragment spanning the promoter-proximal sequenceand start of the open reading frame (primers FPE and Fseq1[Fig. 6B]). When this region was analyzed, a putative IHFbinding site was found (two mismatches with the consensus) atbp 145 to 157. Furthermore, we observed strong binding ofIHF to a 260-bp fragment (primers GBS2 and GPE2 [Fig. 6C])of icsA, which also spans the promoter-proximal sequences andinitial coding region (Fig. 7). As with virF, this region containsa putative IHF binding site downstream of the transcriptionalstart (bp 1109 to 1122) which matches the consensus exactly.Additionally, a low-affinity interaction of IHF with a moreupstream fragment of icsA (primers GBS1 and GBS3 [Fig.6C]) was also detected. Within this sequence, a putative IHFbinding site, harboring two mismatches with the consensus, waslocated at bp 2205 to 2217. It should be pointed out, however,that a shift of similar magnitude but reduced intensity could beproduced by the ihfA mutant lysate. As a positive control, eachmobility shift assay experiment included a DNA fragment con-taining the IHF site II from the fim invertible element, whichcontrols phase variation of type 1 fimbriation in E. coli. StrongIHF-dependent interaction with this fragment was observed(Fig. 7), consistent with published data (8). A shift resulting ina band of lower intensity at a different location on the gel wasseen reproducibly with the ihfA mutant lysate (Fig. 7). Thenature of this shift is unknown.

Interplay of IHF and H-NS in the regulation of virulencegenes. It is known that H-NS represses transcription of thevirF, virB, and icsA genes, and binding of H-NS to the bp 220to 120 region of the virB promoter has been demonstrated invitro (51). H-NS often interacts with DNA close to the RNApolymerase binding site, thus preventing contact by the poly-merase, and this may also be the case at the virF and icsApromoters. IHF can overcome such H-NS-mediated repressionof the bacteriophage Mu early promoter by interacting withDNA adjacent to the RNA polymerase binding site (53). In anattempt to establish whether a similar mechanism operates atany of the promoters in the S. flexneri virulence gene cascade,

we investigated transcript levels in strains harboring wild-typeor mutant copies of IHF and H-NS.

Strains CJD1122 (hns::Tn10 hnrG::Apr) and CJD1123 (ihfA::Tn10 hns::Tn10 hnrG::Apr) were constructed as described inMaterials and Methods (Table 1). The insertion in hnrG, whichencodes a response regulator (19), was used as a selectablemarker, and while strains harboring this mutation had a noti-cably reduced growth rate, there was no significant effect onvirulence gene expression (Table 4) (43). Cultures of BS184(wild type), CJD1104 (ihfA::Tn10), CJD1122, and CJD1123were grown at either 30 or 37°C, and RNA was isolated atOD600 values of 0.6 and 1.4. At this latter point in the growthcurve, transcription of both virF and icsA is beginning to de-cline, and thus it is not possible to make accurate comparisonsof transcript levels in the various strains for these two genes.Probing the RNA isolated from exponentially growing culturesfor a virF-specific transcript confirmed our observation that inthe ihfA mutant strain, CJD1104, transcription was reducedtwofold (Fig. 8). Furthermore, in agreement with publishedresults (10, 42), the level of transcription in the absence of hns(strain CJD1122) at 30°C was increased compared to that inBS184. However, when the two mutations were combined instrain CJD1123, the effect of the ihf mutation was still seen(Fig. 8). Therefore, IHF still plays a direct activating role evenin the absence of H-NS. A similar pattern was observed whenthe same RNA samples were probed for an icsA-specific tran-script. In contrast, virB was transcribed at similar levels in boththe hns mutant and ihf hns double-mutant strains in the sta-tionary phase; this was greater than in the wild-type strain as aresult of the hns mutation. Therefore, the ihf mutation is epi-static to hns in this case. As shown above, IHF has no effect onvirB transcription during the exponential phase. As a control,the same RNA samples were probed for the hupA transcript:no significant variation was observed (Fig. 8).

When the b-galactosidase activity of the mxiC::lacZ fusionwas assayed in cultures of the above strains grown to an OD600of 1.2, the effect of the ihf mutation was again overcome whenhns was mutated (Table 4); this is consistent with the proposalthat IHF has no input downstream of virB.

DISCUSSION

The results presented here suggest a role for IHF in thepositive regulation of S. flexneri virulence genes. The introduc-tion of an ihfA::Tn10 allele into strain BS184 resulted in a two-to threefold reduction in the expression of a lacZ transcrip-tional fusion to the structural gene mxiC. Moreover, this effectwas confined to stationary-phase cells; during logarithmicgrowth, mxiC expression was only marginally reduced com-pared to the wild type in the absence of IHF. An increase intemperature is the primary inducing signal in S. flexneri, andfurther confirmation of this growth phase-specific requirementfor IHF came from studying the kinetics of the thermal induc-tion of mxiC. We found that following a shift from 30 to 37°C,mxiC expression in cells lacking IHF was induced more slowlythan in the wild-type culture, particularly when the shift oc-curred before the onset of the stationary phase. The virB gene,whose product is required for expression of mxiC, was alsoresponsive to IHF only in the stationary phase. A similar pat-tern was seen for two other VirB-dependent structural genes,icsB and ipgD, the latter of which is cotranscribed with mxiC. Incontrast, the expression of virF, which is located one stepfurther up the regulatory cascade from virB and which encodesthe key regulator of the virulence regulon under study, wasdependent on IHF for maximum expression throughout the

TABLE 3. Effect of IHF on expression of the mxiC-lacZ fusionin strains expressing the virF or virB regulatory

genes constitutively at 37°C

Straina Relevant details b-Galactosidaseactivityb

BS184 Wild type 1,029.0 (29.0)CJD1104 ihf 495.7 (10.7)CJD1115 virF pMEP539 17.4 (0.66)CJD1111 virF pMEP537 (virF1) 1,006.1 (10.4)CJD1119 virF ihfA pMEP537 (virF1) 634.6 (30.6)CJD1116 virB pMEP539 17.3 (0.5)CJD1114 virB pMEP538 (virB1) 1,036.8 (0)CJD1120 virB ihfA pMEP538 (virB1) 997.2 (16.2)

a See Table 1 for details of the strains.b Values are given in Miller units. Standard deviations are shown in parentheses.

6544 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 9: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

growth phase, as was an unassociated VirF-regulated gene,icsA.

The observation that an ihfA lesion results in only a two- tothreefold reduction in expression suggests a modulatory rolefor IHF rather than an absolute requirement. These findingsare in agreement with published data on the role of IHF in

other systems. For example, expression of algD in Pseudomo-nas aeruginosa (13, 34) and of the E. coli sodA and ilvPGpromoters (38, 39, 44) is regulated two- to fourfold by IHF. InS. flexneri, expression of the structural genes required for in-vasiveness is dependent upon IHF only in the stationary phase.Since levels of IHF are known to increase by a factor of 5- to

FIG. 5. Mapping of the icsB and ipgD transcription start sites in BS184 (wild type). Start sites are indicated by arrows, and sequencing lanes generated using thesame primer and plasmid pMEP73 as a template are shown alongside. (A) icsB; (B) ipgD; (C) map of the icsB-ipgD intergenic region showing the start sites mappedin panels A and B (angled arrows) and the potential 210 and 235 regions (underlined). The positions of primers used to map the transcription start sites and for gelmobility shift assays (see the text) are indicated by arrows.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6545

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 10: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

10-fold on entry into the stationary phase (14), this suggeststhat only when high levels of IHF are present in the bacteriumdoes it confer its effect. While this is true for the invasivephenotype, we found that maximum expression of the VirFprotein requires IHF at all stages of growth. However, reduc-

tion in the levels of VirF in an ihfA mutant strain is transmittedto the downstream genes only later in the growth phase, whenvirF transcription has started to decline. Therefore, perhapsVirF becomes limiting for the transcription of virB (and thus ofthe downstream genes) only during the stationary phase. It is

FIG. 6. Map of the promoter regions for virB (A), virF (B), and icsA (C) and positions of putative IHF binding sites. For each promoter, the transcriptional startsite is indicated by an angled arrow (see the text for references), and possible 210 and 235 sequences are underlined. The locations of primers used in the mobilityshift assays are indicated by horizontal arrows. Sequences with the best match to the IHF consensus binding site within fragments which were retarded upon incubationwith protein extracts containing IHF are boxed.

6546 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 11: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

likely that IHF also has a direct input at the virB promoter,although here its role is confined to the stationary phase. Thiscould be a direct result of the increase in IHF levels, or,alternatively, it could be due a change in the requirements ofthe virB promoter late in the growth phase. Several lines ofevidence support a role for DNA topology in the regulation ofvirB—more specifically, a decrease in negative supercoilinginhibits transcription (42, 51). Such changes in supercoilingoccur as the cell enters the stationary phase (15), and it ispossible that IHF induces structural changes in the promoterDNA to counteract this effect. Further evidence for this hy-pothesis comes from the observation that IHF appears to playa role in the temporal activation of mxiC, particularly later inthe growth phase, when the DNA template is more relaxed. Anincrease in temperature leads to enhanced negative supercoil-ing; since this change may be more inefficient in the stationaryphase, IHF could play a more important role in the upregula-

tion of expression then. It is noteworthy that in the log phase,IHF has a slight effect on mxiC transcription at 30°C only (Fig.2) and that its effect is reduced at higher salt concentrations,which result in increased negative supercoiling (data notshown). In addition to activating the virB gene, the VirF pro-tein is required for expression of icsA, an unassociated struc-tural gene required for intercellular spreading. Like virF, icsAis also modulated by IHF throughout the growth phase, andtherefore the phenotypic effect of IHF on virulence gene ex-pression is not confined to the stationary phase.

The nucleoid-associated protein H-NS has been shown torepress transcription from the virB promoter by contacting the220 to 120 region; additionally, the VirF binding site has beenmapped from 217 to 2117 (51). Our genetic evidence fromthe analysis of virB transcript levels in strains harboring wild-type or mutant copies of ihfA or hns shows that activation byIHF appears to be dependent upon H-NS-mediated repres-sion; the ihfA mutation is epistatic to the hns mutation. Such arole for IHF in overcoming repression by H-NS at the earlypromoter of bacteriophage Mu has already been established(53). In this example, IHF is also required for the direct acti-vation of RNA polymerase and binds close to the start site. Atthe virB promoter, however, IHF appears to bind upstream ofboth the H-NS binding site and the region contacted by VirF.The mechanism by which VirF activates this promoter has notbeen established, but it is absolutely required for transcription.It is conceivable that binding of IHF to the upstream siteresults in a conformational change in the DNA, facilitatingcontact by VirF to the downstream region in the presence ofH-NS. When H-NS is absent, VirF may be able to form com-plexes at the virB promoter without assistance from IHF; in-deed, overexpression of VirF can also overcome repression byH-NS (10, 43). Why this mechanism is apparently confined tothe stationary phase is unclear, but it probably involves DNAtopology, as discussed above. Furthermore, the experimentsdescribed here involving the constitutive expression of the virFor virB genes from expression vectors confirm that the bindingof IHF to the virB promoter plays a direct transcriptional rolein vivo and that its effect on virulence gene expression is notsolely due to its action at virF.

Further work is required to elucidate the precise mecha-nism(s) by which IHF modulates S. flexneri virulence geneexpression. By using gel mobility shift assays, we were able toshow that IHF appears to bind to the virF, virB, and icsApromoters but not to those of the structural genes icsB andipgD. Thus, IHF has multiple input sites within the virulencegene regulon. Detailed biochemical analysis is required to de-fine the putative binding sites more accurately.

IHF has previously been found to activate virulence systemsin a variety of organisms. As well as the algD gene of P. aerugi-

FIG. 7. Retardation of fragments of the virB, icsB, and ipgD (upper panel)and virF and icsA (lower panel) promoter regions with crude cell extracts fromBS184 (wild type; lanes 2), CJD1104 (ihfA::Tn10; lanes 3), or CJD1106 (ihfA::Tn10 pBR322-ihfA; lanes 4). Lanes 1 contain unbound probe DNA. The PCRprimers used to amplify the fragments are indicated for each set of lanes; thelocations of the primers are given in Fig. 5 and 6. A fragment of E. coli fim DNAcontaining IHF site II was similarly amplified by PCR and used as a positivecontrol (lower panel).

TABLE 4. Expression of the mxiC::lacZ fusion in strains harboringwild-type or mutant copies of ihfA or hns after growth

at 30 or 37°C to an OD600 of 1.2

Straina Relevant detailsb-Galactosidase activityb at:

30°C 37°C

BS184 Wild type 20.3 (0.64) 435.4 (16.7)CJD1121 hnrG::Apr 25.21 (0.85) 399.0 (6.3)CJD1104 ihfA::Tn10 8.3 (0.1) 218.2 (10.9)CJD1122 hnrG::Apr hns::Tn10 93.51 (3.0) 590.2 (14.3)CJD1123 ihfA::Tn10 hnrG::Apr hns::Tn10 71.93 (1.2) 585.4 (12.5)

a See Table 1 for details of the strains.b Values are given in Miller units. Standard deviations are shown in parentheses.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6547

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 12: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

nosa, pilus production in Neisseria gonorrhoeae (25) and thesynthesis of type 1 fimbriae in E. coli (16) are regulated by IHFin response to various environmental stimuli. For the S. flexnerivirulence genes, the primary inducing signal is a growth tem-perature of 37°C, and our results show that IHF is importantfor the upregulation of expression following an increase in

temperature, particularly in cells entering the stationary phaseof growth. The conditions of stationary phase in the laboratorymay be similar to those encountered by the organism duringinfection in vivo, and the ihfA gene of Salmonella typhimuriumhas been shown to be expressed in vivo during mouse infection(31). Enteric bacteria encounter multiple stresses both inside

FIG. 8. Interplay of the ihfA mutation with an hns mutation. Cultures of BS184 (wild type), CJD1104 (ihfA::Tn10), CJD1122 (hns::Tn10), and CJD1123 (ihfA::Tn10hns::Tn10) were grown at 30 or 37°C, and RNA was extracted at OD600 values of 0.6 and 1.4. (A) Northern blots of the virF and icsA transcripts from log-phase culturesand of the virB and hupA (control) transcripts from log-phase and stationary-phase cultures as indicated. (B) The virF, icsA, virB, and hupA bands in panel A werescanned densitometrically, and the data are expressed as a percentage of the most intense band for each transcript. The experiment was repeated three times, and typicaldata are presented.

6548 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 13: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

and outside the human host, and since expression of the inva-sion genes is required only when S. flexneri is inside the humanintestine, it is not surprising that the bacteria have evolved acomplex system to respond to a range of environmental inputs.It is very likely that the interplay among DNA supercoiling,H-NS, and IHF presumably contributes to the fine-tuning ofthe virulence response to meet these varying conditions.

ACKNOWLEDGMENTS

We are grateful to Andrew Free for a critical reading of the manu-script.

This work was supported by grant 042380/Z/94/Z from the Well-come Trust (United Kingdom).

REFERENCES

1. Adler, B., C. Sasakawa, T. Tobe, S. Makino, K. Komatsu, and M. Yoshikawa.1989. A dual transcriptional activation system for the 230 kb plasmid genescoding for virulence-associated antigens of Shigella flexneri. Mol. Microbiol.3:627–635.

2. Allaoui, A., J. Mounier, M.-C. Prevost, P. J. Sansonetti, and C. Parsot. 1992.icsB, a Shigella flexneri virulence gene necessary for the lysis of protrusionsduring intercellular spread. Mol. Microbiol. 6:1605–1616.

3. Allaoui, A., P. J. Sansonetti, and C. Parsot. 1992. MxiJ, a lipoprotein in-volved in secretion of Shigella Ipa invasins, is homologous to YscJ, a secre-tion factor of the Yersinia Yop proteins. J. Bacteriol. 174:7661–7669.

4. Allaoui, A., P. J. Sansonetti, and C. Parsot. 1993. MxiD: an outer membraneprotein necessary for the secretion of the Shigella flexneri Ipa invasins. Mol.Microbiol. 7:59–68.

5. Andrews, G. P., and A. T. Maurelli. 1992. mxiA of Shigella flexneri 2a, which

FIG. 8—Continued.

VOL. 179, 1997 IHF AND SHIGELLA VIRULENCE GENES 6549

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from

Page 14: Positive Regulation of Shigella flexneri Virulence Genes by ... · Positive Regulation of Shigella flexneri Virulence Genes ... in virulence gene expression. We found that tran-scription

facilitates export of invasion plasmid antigens, encodes a homolog of thelow-calcium response protein, LcrD, of Yersinia pestis. Infect. Immun. 60:3287–3295.

6. Bernardini, M. L., J. Mounier, H. d’Hauteville, M. Coquis-Rondon, and P. J.Sansonetti. 1989. Identification of icsA, a plasmid locus of Shigella flexnerithat governs intra- and inter-cellular spread through the interaction withF-actin. Proc. Natl. Acad. Sci. USA 86:3867–3871.

7. Bernardini, M. L., A. Fontaine, and P. J. Sansonetti. 1990. The two com-ponent regulatory system OmpR-EnvZ controls the virulence of Shigellaflexneri. J. Bacteriol. 172:6274–6281.

8. Blomfield, I. C., D. H. Kulasekara, and B. I. Eisenstein. 1997. Integrationhost factor stimulates both FimB- and FimE-mediated site-specific DNAinversion that controls phase variation of type 1 fimbriae expression inEscherichia coli. Mol. Microbiol. 23:705–717.

9. Bolivar, F., R. L. Rodriguez, M. V. Greene, H. L. Betlach, H. L. Heynecker,H. W. Boyer, J. H. Crosa, and S. Falkow. 1977. Construction and character-ization of new cloning vehicles. II. A multipurpose cloning system. Gene2:95–113.

10. Colonna, B., M. Casalino, P. A. Fradiani, C. Zagaglia, S. Naitza, L. Leoni, G.Prosseda, A. Coppo, P. Ghelardini, and M. Nicoletti. 1995. H-NS regulationof virulence gene expression in enteroinvasive Escherichia coli harboring thevirulence plasmid integrated into the host chromosome. J. Bacteriol. 177:4703–4712.

11. Craig, N. L., and H. A. Nash. 1984. E. coli integration host factor binds tospecific sites in DNA. Cell 39:707–716.

12. Dagberg, B., and B. E. Uhlin. 1992. Regulation of virulence associatedplasmid genes in enteroinvasive Escherichia coli. J. Bacteriol. 174:7606–7612.

13. Delic-Attree, I., B. Toussaint, A. Froger, J. C. Wilson, and P. M. Vignais.1996. Isolation of an IHF-deficient mutant of a Pseudomonas aeruginosamucoid isolate and evaluation of the role of IHF in algD gene expression.Microbiology 142:2785–2793.

14. Ditto, M. D., D. Roberts, and R. A. Weisberg. 1994. Growth phase variationof integration host factor level in Escherichia coli. J. Bacteriol. 176:3738–3748.

15. Dorman, C. J., G. C. Barr, N. Nı Bhriain, and C. F. Higgins. 1988. DNAsupercoiling and the anaerobic and growth phase regulation of tonB expres-sion. J. Bacteriol. 170:2816–2826.

16. Dorman, C. J., and C. F. Higgins. 1987. Fimbrial phase variation in Esche-richia coli: dependence on integration host factor and homologies with othersite specific recombinases. J. Bacteriol. 169:3840–3843.

17. Dorman, C. J., N. Nı Bhriain, and C. F. Higgins. 1990. DNA supercoilingand environmental regulation of virulence gene expression in Shigella flex-neri. Nature 344:789–792.

18. Falconi, M., N. P. Higgins, R. Spurio, C. L. Pon, and C. O. Gualerzi. 1993.Expression of the gene encoding the major bacterial nucleoid protein H-NSis subject to transcriptional autorepression. Mol. Microbiol. 10:273–282.

19. Free, A. 1995. H-NS and the regulation of transcription in Escherichia coli.Ph.D. thesis. University of Dublin, Dublin, Republic of Ireland.

20. Freundlich, M., N. Ramani, E. Mathew, A. Sirko, and P. Tsui. 1992. The roleof integration host factor in gene expression in Escherichia coli. Mol. Micro-biol. 6:2557–2563.

21. Goodrich, J. A., M. L. Schwartz, and W. R. McClure. 1990. Searching for andpredicting the activity of sites for DNA binding proteins: compilation andanalysis of the binding sites for Escherichia coli integration host factor (IHF).Nucleic Acids Res. 18:4993–5000.

22. Goosen, N., and P. van de Putte. 1995. The regulation of transcriptioninitiation by integration host factor. Mol. Microbiol. 16:1–7.

23. Hale, T. L. 1991. Genetic basis of virulence in Shigella species. Microbiol.Rev. 55:206–224.

24. Higgins, C. F., C. J. Dorman, and N. Nı Bhriain. 1990. Environmentalinfluences on DNA supercoiling: a novel mechanism for the regulation ofgene expression, p. 421–432. In M. Riley and K. Drlica (ed), The bacterialchromosome. American Society for Microbiology, Washington, D.C.

25. Hill, S. A., D. S. Samuels, J. H. Carlson, J. Wilson, D. Hogan, L. Lubke, andR. J. Belland. 1997. Integration host factor is a transcriptional cofactor ofpilE in Neisseria gonorrhoeae. Mol. Microbiol. 23:649–656.

26. Hromockyj, A. E., and A. T. Maurelli. 1989. Identification of Shigella inva-sion genes by isolation of temperature-regulated inv::lacZ operon fusions.Infect. Immun. 57:2963–2970.

27. Hromockyj, A. E., S. C. Tucker, and A. T. Maurelli. 1992. Temperatureregulation of Shigella virulence: identification of the repressor gene virR, ananalogue of hns, and partial complementation by tyrosyl transfer RNA(tRNA1

tyr). Mol. Microbiol. 6:2113–2124.28. Karem, K., and J. W. Foster. 1993. The influence of DNA topology on the

environmental regulation of a pH-regulated locus in Salmonella typhi-murium. Mol. Microbiol. 10:75–86.

29. Lambert de Rouvroit, C., C. Sluiters, and G. R. Cornelis. 1992. Role of thetranscriptional activator, VirF, and temperature in the expression of the

pYV plasmid genes of Yersinia enterocolitica. Mol. Microbiol. 6:395–409.30. Lett, M., C. Sasakawa, N. Okada, T. Sakai, S. Makino, M. Yamada, K.

Komatsu, and M. Yoshikawa. 1989. virG, a plasmid-encoded virulence geneof Shigella flexneri: identification of the virG protein and determination of thecomplete coding sequence. J. Bacteriol. 171:353–359.

31. Mahan, M. J., J. M. Slauch, and J. J. Mekalanos. 1993. Selection of bacterialvirulence genes that are specifically induced in host tissues. Science 259:686–688.

32. Maurelli, A. T., B. Blackmon, and R. Curtiss III. 1984. Temperature-depen-dent expression of virulence genes in Shigella flexneri. Infect. Immun. 43:195–201.

33. Maurelli, A. T., and P. J. Sansonetti. 1988. Identification of a chromosomalgene controlling temperature-regulated expression of Shigella virulence.Proc. Natl. Acad. Sci. USA 85:2820–2824.

34. May, T. B., and A. M. Chakrabarty. 1994. Pseudomonas aeruginosa: genesand enzymes of alginate synthesis. Trends Microbiol. 2:151–157.

35. Miller, H. I. 1984. Primary structure of the himA gene of Escherichia coli:homology with DNA binding protein HU and association with the phenyl-alanine tRNA synthetase operon. Cold Spring Harbor Symp. Quant. Biol.49:691–698.

36. Miller, J. H. 1972. Experiments in molecular genetics. Cold Spring HarborLaboratory Press, Cold Spring Harbor, N.Y.

37. Nakayama, S.-I., and H. Watanabe. 1995. Involvement of cpxA, a sensor ofa two-component regulatory system in the pH-dependent regulation of ex-pression of the Shigella sonnei virF gene. J. Bacteriol. 177:5062–5069.

38. Pagel, J. M., J. W. Winkelman, C. W. Adams, and G. W. Hatfield. 1992. DNAtopology-mediated regulation of transcription initiation from the tandempromoters of the ilvGMEDA operon of Escherichia coli. J. Mol. Biol. 224:919–935.

39. Parekh, B. S., and G. W. Hatfield. 1996. Transcriptional activation by pro-tein-induced DNA bending: evidence for a DNA structural transmissionmodel. Proc. Natl. Acad. Sci. USA 93:1173–1177.

40. Porter, M. E., and C. J. Dorman. 1994. A role for H-NS in the thermo-osmotic regulation of virulence gene expression in Shigella flexneri. J. Bac-teriol. 176:4187–4191.

41. Porter, M. E., and C. J. Dorman. 1997. Virulence gene deletion frequency isincreased in Shigella flexneri following conjugation, transduction and trans-formation. FEMS Microbiol. Lett. 147:163–172.

42. Porter, M. E., and C. J. Dorman. Differential regulation of the plasmid-encoded genes in the Shigella flexneri virulence regulon. Mol. Gen. Genet., inpress.

43. Porter, M. E., and C. J. Dorman. Unpublished data.44. Presutti, D. G., and H. M. Hassan. 1995. Binding of integration host factor

(IHF) to the Escherichia coli sodA gene and its role in the regulation of asodA-lacZ fusion gene. Mol. Gen. Genet. 246:228–235.

45. Rice, P. A., S. Yang, K. Mizuuchi, and H. Nash. 1996. Crystal structure of anIHF-DNA complex: a protein-induced U-turn. Cell 87:1295–1306.

46. Sakai, T., C. Sasakawa, and M. Yoshikawa. 1988. Expression of four viru-lence antigens of Shigella flexneri is positively regulated at the transcriptionallevel by the 30 kilodalton virF protein. Mol. Microbiol. 2:589–597.

47. Sasakawa, C., K. Komatsu, T. Tobe, T. Suzuki, and M. Yoshikawa. 1993.Eight genes in region 5 that form an operon essential for invasion of epi-thelial cells by Shigella flexneri 2a. J. Bacteriol. 175:2334–2346.

48. Silhavy, T. J., M. L. Berman, and L. W. Enquist. 1984. Experiments withgene fusions. Cold Spring Harbor Laboratory Press, Cold Spring Harbor,N.Y.

49. Sternberg, N. L., and R. Mauer. 1991. Bacteriophage-mediated generalizedtransduction in Escherichia coli and Salmonella typhimurium. Methods En-zymol. 204:18–43.

50. Tobe, T., S. Nagai, N. Okada, B. Adler, M. Yoshikawa, and C. Sasakawa.1991. Temperature-regulated expression of invasion genes in Shigella flexneriis controlled through the transcriptional activation of the virB gene on thelarge plasmid. Mol. Microbiol. 5:887–893.

51. Tobe, T., M. Yoshikawa, T. Mizuno, and C. Sasakawa. 1993. Transcriptionalcontrol of the invasion regulatory gene virB of Shigella flexneri: activation byVirF and repression by H-NS. J. Bacteriol. 175:6142–6149.

52. Tobe, T., M. Yoshikawa, and C. Sasakawa. 1995. Thermoregulation of virBtranscription in Shigella flexneri by sensing of changes in local DNA super-helicity. J. Bacteriol. 177:1094–1097.

53. van Ulsen, P., M. Hillebrand, L. Zulianello, P. van de Putte, and N. Goosen.1996. Integration host factor alleviates the H-NS-mediated repression of theearly promoter of bacteriophage Mu. Mol. Microbiol. 21:567–578.

54. Velterop, J. S., M. A. Dijkhuizen, R. van’t Hof, and P. W. Postma. 1995. Aversatile vector for controlled expression of genes in Escherichia coli andSalmonella typhimurium. Gene 153:63–65.

55. Venatesan, M. M., J. M. Buysse, and E. V. Oaks. 1992. Surface presentationof Shigella flexneri invasion plasmid antigens requires the products of the spalocus. J. Bacteriol. 174:1990–2001.

6550 PORTER AND DORMAN J. BACTERIOL.

on July 1, 2020 by guesthttp://jb.asm

.org/D

ownloaded from