You are on page 1of 13

Molecular Microbiology (2014) 91(6), 1057–1069 ■ doi:10.1111/mmi.

12500
First published online 13 January 2014

Regulation of Salmonella enterica pathogenicity island 1


(SPI-1) by the LysR-type regulator LeuO

Elena Espinosa and Josep Casadesús* ligands that modulate their DNA binding ability upon
Departamento de Genética, Facultad de Biología, binding to a C-terminal domain (Zaim and Kierzek, 2003).
Universidad de Sevilla, Apartado 1095, Sevilla LTTRs can be either activators or repressors of transcrip-
E-41080, Spain. tion (Maddocks and Oyston, 2008).
LeuO is a LTTR originally identified in Salmonella
enterica (Hertzberg et al., 1980; Haughn et al., 1986), and
Summary later found in other genera of gramnegative bacteria
LeuO is a quiescent LysR-type regulator belonging to including Escherichia, Shigella, Vibrio and Yersinia
the H-NS regulon. Activation of leuO transcription (Maddocks and Oyston, 2008; Stratmann et al., 2008).
represses expression of pathogenicity island 1 (SPI-1) Many genes activated by LeuO are repressed by H-NS,
in Salmonella enterica serovar Typhimurium and and for this reason LeuO is often viewed as a transcrip-
inhibits invasion of epithelial cells. Loss of HilE sup- tional antagonist of H-NS (Chen and Wu, 2005; Chen et al.,
presses LeuO-mediated downregulation of SPI-1. Acti- 2005; Hernandez-Lucas et al., 2008; Stratmann et al.,
vation of leuO transcription reduces the level of HilD 2008; Shimada et al., 2009; 2011; Gallego-Hernandez
protein, and loss of HilE restores the wild type HilD et al., 2012) and of its paralogue StpA (De la Cruz et al.,
level. Hence, LeuO-mediated downregulation of SPI-1 2007; Stratmann et al., 2012). Antagonism may involve
may involve inhibition of HilD activity by HilE, a view competition with H-NS for binding to DNA (De la Cruz et al.,
consistent with the fact that HilE is a HilD inhibitor. In 2007; Shimada et al., 2011) or formation of a barrier that
vivo analyses using β-galactosidase fusions indicate hinders H-NS polymerization (Chen and Wu, 2005; Chen
that LeuO activates hilE transcription. In vitro analyses et al., 2005). In other genes, however, LeuO has been
by slot blotting, electrophoretic mobility shift analysis shown to repress transcription (Shi and Bennett, 1995;
and DNase I footprinting show that LeuO binds the hilE Repoila and Gottesman, 2001). In the last few years,
promoter region. Although residual SPI-1 repression identification of LeuO binding sites in the genomes of
by LeuO is observed in the absence of HilE, the LeuO- E. coli and Salmonella has defined DNA motifs bound by
HilE-HilD ‘pathway’ appears to be the major mecha- LeuO, and high throughput analysis of gene expression
nism. Because both leuO and SPI-1 are repressed by has increased the number of loci ascribed to the LeuO
H-NS, activation of leuO transcription may provide a regulon (Shimada et al., 2011; Dillon et al., 2012;
backup mechanism for SPI-1 repression under condi- Hernandez-Lucas and Calva, 2012).
tions that impair H-NS-mediated silencing. Expression of the leuO gene is low during standard
laboratory growth as a consequence of H-NS-mediated
Introduction repression (Klauck et al., 1997). It is thus conceivable that
activation of LeuO synthesis under specific circumstances
LysR-type transcriptional regulators (LTTRs) are members can permit expression of genes that are usually repressed
of a large family of prokaryotic proteins involved in diverse by H-NS and its paralogues (Dorman, 2007; Hernandez-
cellular functions including transport, response to oxidative Lucas and Calva, 2012). Such circumstances remain
stress, nitrogen fixation, biofilm formation and bacterial largely unknown, although leuO expression in E. coli is
virulence (Henikoff et al., 1988; Maddocks and Oyston, known to be moderately higher in minimal low phosphate
2008; Momany and Neidle, 2012). LTTRs are typically medium than in LB, and also in stationary phase compared
300–350 amino acids long and contain an N-terminal with exponential growth (Fang et al., 2000; Majumder
helix–turn–helix DNA binding motif (Maddocks and et al., 2001).
Oyston, 2008). Some LTTRs act in combination with Examples of bacterial functions under LeuO control
Accepted 18 December, 2013. *For correspondence. E-mail
include biofilm formation in Vibrio cholerae (Moorthy and
casadesus@us.es; Tel. (+34) 95542 0881; Fax (+34) 95455 7104. Watnick, 2005) and E. coli (Shimada et al., 2011), and

© 2013 John Wiley & Sons Ltd


1058 E. Espinosa and J. Casadesús ■

synthesis of E. coli fimbriae (Shimada et al., 2011). In might be relevant under circumstances that impair SPI-1
Salmonella enterica serovar Typhi, LeuO activates the silencing by H-NS.
synthesis of quiescent porins OmpS1 and OmpS2
(Fernandez-Mora et al., 2004; Rodriguez-Morales et al., Results
2006; De la Cruz et al., 2007; Hernandez-Lucas et al.,
Activation of leuO transcription represses SPI-1
2008). In Salmonella enterica serovar Typhimurium,
OmpS1 and OmpS2 are required for virulence in To confirm previous evidence indicating that SPI-1 might
the mouse model of typhoid (Lawley et al., 2006; be repressed by LeuO (Dillon et al., 2012), we compared
Rodriguez-Morales et al., 2006). LeuO was also found in a the expression of selected SPI-1 genes in the presence
screen for Salmonella virulence factors using Caenorhab- and in the absence of LeuO. Activation of leuO transcrip-
ditis elegans as host (Tenor et al., 2004) and in a screen for tion was achieved by insertion of a T-POP element
Salmonella genes required for long-term infection of mice upstream of the leuO coding sequence (Fig. S1). Tran-
(Lawley et al., 2006). Lack of LeuO causes attenuation in scription of leuO was activated by addition of autoclaved
the mouse model of typhoid (Rodriguez-Morales et al., chlortetracycline (Rappleye and Roth, 1997; Lee et al.,
2006). 2007; Dillon et al., 2012). This experimental design was
In this study, we show that LeuO downregulates the chosen for the following reasons: (i) In E. coli, leuO expres-
expression of a major virulence determinant of Salmo- sion increases in stationary cultures; in S. enterica,
nella enterica serovar Typhimurium, the gene cluster however, increase of leuO expression under such condi-
known as Salmonella pathogenicity island 1 (SPI-1) tions is small (Fig. S2); (ii) We discarded the use of an
(Galan and Curtiss, 1989). SPI-1 encodes a type 3 Hns− mutant because hns mutations are detrimental in
secretion system and secreted effectors that promote S. enterica ser. Typhimurium, unless accompanied by an
invasion of the intestinal epithelium by interacting with rpoS mutation as in strain LT2 (Wilmes-Riesenberg et al.,
proteins inside epithelial cells (Galan and Curtiss, 1989). 1997); (iii) The combination of hns and leuO mutations
SPI-1 genes are organized in seven or more transcrip- strongly impairs S. enterica viability, and normal growth
tional units under the control of the SPI-1-encoded tran- may require the acquisition of suppressor mutations of
scription factors HilA, HilC, HilD and InvF (Lostroh and unknown nature (data not shown). Use of T-POP to drive
Lee, 2001) and of RtsA, a transcription factor encoded leuO transcription does not cause viability problems, and
outside SPI-1 (Ellermeier and Slauch, 2003). SPI-1 has two additional advantages: (i) the transcription rate can
expression is controlled by additional regulators located be controlled by using different concentrations of tetracy-
outside the island such as the Fur and BarA/SirA acti- cline or chlortetracycline (Lee et al., 2007); (ii) in the strain
vators (Fortune et al., 2006; Ellermeier and Slauch, engineered for this study (SV6141), T-POP prevents tran-
2008) and the HilE repressor (Fahlen et al., 2000). scription from the native leuO promoter, thus avoiding a
Under conditions that permit Salmonella invasion of epi- feedback loop of autogenous activation (Fang and Wu,
thelial cells, SPI-1 expression is bistable (Sturm et al., 1998a,b) that might yield undesirably high levels of LeuO.
2010; Bailly-Bechet et al., 2011). Under the conditions employed in this study, the level of
Our study was inspired by previous findings indicating leuO expression was similar to that of an Hns− mutant
that SPI-1 contains a high density of LeuO binding sites, (Fig. S3).
and that deletion of leuO in serovar Typhimurium Expression of SPI-1 was monitored by measuring the
increases expression of the SPI-1 gene sopA (Dillon β-galactosidase activity of lac fusions in six genes: hilA,
et al., 2012). Indeed, we show that LeuO represses tran- hilC, hilD and invF, which encode transcriptional regulators
scription of SPI-1 and inhibits epithelial cell invasion. of SPI-1 (Altier, 2005; Jones, 2005; Ellermeier and Slauch,
However, we provide evidence that LeuO mainly exerts 2007); invH, which encodes a component of the SPI-1
SPI-1 repression by an indirect mechanism, hitherto secretion apparatus (Ellermeier and Slauch, 2007); and
unsuspected: LeuO activates transcription of the hilE rtsA, a transcriptional regulator of SPI-1 encoded outside
gene, which is located outside SPI-1 and encodes a SPI-1 (Ellermeier and Slauch, 2003; 2007; Jones, 2005).
SPI-1 repressor (Fahlen et al., 2000; Baxter et al., 2003). Activation of leuO transcription reduced the expression of
Because LeuO is a quiescent regulator under laboratory all lac fusions (Fig. 1). As controls, strains carrying the
conditions (Hernandez-Lucas and Calva, 2012), the sig- same lac fusions and a T-POP insertion unlinked to leuO
nificance of LeuO-mediated SPI-1 repression in the Sal- (zzz:T-POP) were used. In four control strains, the
monella lifestyle remains to be established. However, the β-galactosidase activities were similar to that of the wild
fact that both leuO and SPI-1 are repressed by H-NS type, in one strain was slightly higher, and in another strain
(Klauck et al., 1997; Lucchini et al., 2006; Navarre et al., was lower (Fig. S4). This experiment ruled out the possi-
2006) suggests that LeuO might provide a backup bility of an artefact caused by either T-POP or chlortetra-
mechanism for SPI-1 repression. Such a mechanism cycline. We thus concluded that LeuO does repress SPI-1.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1059

Fig. 1. Regulation of SPI-1 expression by


LeuO. β-Galactosidase activity of hilD::lac,
hilC::lac, invF::lac, rtsA::lac, hilA::lac and
invH::lac fusions in the presence and in the
absence of LeuO. Data are averages and
standard deviations from > 3 independent
experiments.

LeuO downregulates SPI-1 expression via hilE and hilD dependent and HilE-independent mechanisms. However,
activation of leuO transcription in a HilE+ background yields
Because SPI-1 expression is responsive to multiple a homogeneous population of (SPI-1)OFF cells while acti-
regulators (Ellermeier and Slauch, 2007), we devised a vation of leuO transcription in the absence of HilE yields
genetic screen to ascertain whether a single cell function a small subpopulation of SPI-1(OFF) cells. Hence,
might transmit LeuO-mediated regulation to SPI-1. For HilE-dependent downregulation seems to be the major
this purpose, Tn10dCm mutagenesis was performed in ‘pathway’ of SPI-1 repression by LeuO.
a strain that carried the T-POP leuO construct and a Because HilE is a negative regulator of hilD (Baxter
hilC::lac translational fusion (SV6844). When leuO tran- et al., 2003), a tentative interpretation for the occurrence of
scription is activated, strain SV6844 is Lac− and forms HilE-dependent SPI-1 repression was that LeuO might
white colonies on plates containing Xgal. In the screen, increase the HilE level, which in turn might inhibit HilD
Lac+ (blue) colonies were sought among the Cmr isolates activity. Western blot analysis revealed that activation of
generated by Tn10Cm insertion. One such isolate was leuO transcription decreased the HilD level, and the
purified and re-constructed by P22 HT transduction to decrease was suppressed in a ΔleuO background (Fig. 3).
confirm that the Tn10dCm insertion suppressed hilC::lac We also observed that a hilE null mutation increased the
downregulation by LeuO. The isolate was propagated as HilD levels in the presence and in the absence of LeuO
strain SV7036. (Fig. 3), in accordance with two well known facts: the
Amplification by semi-random PCR and sequencing of inhibition of HilD activity by HilE (Baxter et al., 2003) and
the Tn10dCm boundaries indicated that the Tn10dCm the occurrence of autogenous activation of hilD transcrip-
element of SV7036 had disrupted the hilE gene. Use of a tion (Ellermeier et al., 2005). These experiments support
HilE− null mutant constructed ad hoc (strain SV5586) pro- the view that downregulation of SPI-1 by LeuO may involve
vided independent evidence that lack of HilE suppressed HilE-mediated inhibition of HilD activity.
hilC::lac downregulation by LeuO. The involvement of HilD in LeuO-mediated SPI-1 repres-
Single cell analysis of gene expression by flow cytometry sion was further investigated by epistasis analysis. This
(Fig. 2) confirmed that HilE plays a role in LeuO-mediated kind of analysis takes advantage of two well known traits of
repression of SPI-1: (i) activation of leuO expression SPI-1 expression. One is redundancy of certain transcrip-
decreased the activity of a sipB::GFP fusion, abolishing tion factors involved in SPI-1 control (Altier, 2005; Jones,
bistable SPI-1 expression; (ii) lack of LeuO restored the 2005; Ellermeier and Slauch, 2007). Another is that lack of
wild type pattern of sipB::GFP expression, indicating that a single transcription factor does not completely abolish
SPI-1 downregulation was caused by LeuO indeed; (iii) a expression of certain SPI-1 transcriptional units (Ellermeier
hilE null mutation increased sipB::GFP expression and et al., 2005). We thus designed experiments to test the
reduced the size of the SPI-1(OFF) subpopulation, in effect of a hilD null mutation on the expression of rtsA::lac
agreement with the role of HilE as a SPI-1 repressor and hilC::lac fusions in the presence and in the absence of
(Baxter et al., 2003); (iv) activation of leuO expression in a LeuO. If HilD was required for SPI-1 repression by LeuO,
HilE− background yielded a small subpopulation of sipB- we reasoned, downregulation of SPI-1 by LeuO should not
::GFP(OFF) cells; and (v) absence of both LeuO and HilE be observed in the absence of HilD. Results shown in
restored the wild type pattern of sipB::GFP expression. Fig. 4 did not completely fulfil this prediction: moderate
Altogether, these observations suggest that activation of downregulation of SPI-1 by LeuO was observed in a HilD−
LeuO expression downregulates SPI-1 by both HilE- background. Furthermore, even though a hilD mutation

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
1060 E. Espinosa and J. Casadesús ■

Fig. 2. Flow cytometry analysis of SPI-1 expression. Effect of a hilE null mutation on sipB::GFP expression in the presence and in the
absence of LeuO (left column). Data were collected for 40 000 events per sample. In the histograms presented, the cell numbers have been
normalized to 100. Data are also represented by a dot plot (right column).

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1061

others contained individual promoters P1 (pIZ1998), P2


(pIZ1999) and P3 (pIZ2000). Diagrams of the constructs
are shown in Fig. 5B. Measurements of β-galactosidase
activities of the plasmid-borne lac fusions showed that
LeuO regulates expression of the P1, P2 and P3 hilE
promoters in an independent manner (Fig. 4C). However,
LeuO-dependent regulation was found to be stronger
when the three promoters were present (Fig. 5C).

Binding of LeuO to the hilE promoter region


Fig. 3. Effect of a hilE mutation on the levels of HilD protein in the
presence and in the absence of LeuO. DnaK was used as loading To test whether LeuO is able to bind the hilE promoter
control in Western blots.
region, a slot blot binding assay was performed. A 449 bp
DNA fragment containing the P1, P2 and P3 hilE promot-
was epistatic over hilE, moderate repression of SPI-1 ers was incubated with increasing concentrations of LeuO
by LeuO was still observed in a HilD− HilE− background protein. Binding was unambiguously detected (Fig. 6A).
(Fig. 4). Hence, LeuO seems to repress SPI-1 by a major Quantitative analysis of binding (Fig. 6B) indicated that,
‘pathway’ involving HilD and HilE, and also by minor, HilD- under the conditions of the assay, LeuO bound the hilE
and HilE-independent mechanisms. This conclusion is DNA fragment with an approximate Kd of 0.37 μM. As a
coherent with the detection of a subpopulation of SPI- negative control, a binding assay with the rtsA promoter
1(OFF) cells when LeuO expression was activated in a was performed, and LeuO was unable to bind the DNA
HilE− background (Fig. 2). fragment (data not shown).
Binding of LeuO to the hilE promoter region was also
tested by electrophoretic mobility shift assays. DNA frag-
LeuO activates hilE transcription
ments that contained the entire promoter region or the
A tentative model for LeuO-mediated downregulation of individual P1, P2 and P3 promoters were used. LeuO
SPI-1 via HilE and HilD is that LeuO may activate hilE binding was observed upstream of P2 and P3 but not
expression, and that HilE-mediated inhibition of HilD upstream of P1 (Fig. 6C). This observation is consistent
activity (Baxter et al., 2003) may contribute to SPI-1 with two facts: (i) the weak level of P1-driven hilE tran-
downregulation. To test whether LeuO is an activator of scription detected with the promoter-probe plasmid
hilE expression, a hilE::lacZ translational fusion was con- (Fig. 5C); (ii) the existence of putative LeuO binding sites
structed on the Salmonella chromosome. Comparison of upstream of P2 and P3 but not upstream of P1 (Fig. S5).
β-galactosidase activities in the presence and in the Further evidence of LeuO binding to the hilE promoter
absence of LeuO (strains SV7327 and SV7328) indicated region was obtained by DNA footprinting using the cus-
that hilE expression is activated by LeuO (Fig. 5A). tomary 449 bp DNA fragment as target. Protection from
Transcription of hilE is known to be driven by three DNase I digestion was observed in the presence of LeuO
promoters (Lim et al., 2007). To identify the hilE promot- (Fig. 6D). Protection was stronger at a region that over-
er(s) under LeuO control, the P1, P2 and P3 promoters laps the P3 promoter. Because LTTRs can act at a dis-
were cloned on the promoter-probe vector pIC552 (Macian tance (Maddocks and Oyston, 2008; Momany and Neidle,
et al., 1994) to generate transcriptional lac fusions. A con- 2012), LeuO binding at the P3 region may be sufficient to
struct contained the three promoters (pIZ1997), and the permit regulation of the downstream promoter P2 (and

Fig. 4. Epistasis analysis of SPI-1


expression. Effect of LeuO expression on the
β-galactosidase activity of rtsA::lac and
hilC::lac translational fusions in HilD− and
HilD− HilE− backgrounds. Data are averages
and standard deviations from > 3 independent
experiments.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
1062 E. Espinosa and J. Casadesús ■

Fig. 5. Regulation of hilE expression by


LeuO.
A. β-Galactosidase activity of an hilE::lac
translational fusion in the presence and in the
absence of LeuO.
B. Diagram of the hilE promoter region, drawn
using information from (Lim et al., 2007). The
DNA fragments cloned on pIC552 are
represented below. The −505 to −56 fragment
contains the three hilE promoters (plasmid
pIZ1997). The −159 to −56 fragment contains
the P1 promoter (pIZ1998). The −334 to −161
fragment contains the P2 promoter (pIZ1999).
The −505 to −336 fragment contains the P3
promoter (pIZ2000).
C. β-Galactosidase activities of pIZ1997,
pIZ1998, pIZ1999 and pIZ2000 in the
presence and in the absence of LeuO. The
pIC552 vector was included as control. Data
are averages and standard deviations from
> 3 independent experiments.

perhaps P1). It is also possible that lower affinity binding Discussion


to P2 may boost transcription.
The existence of LeuO has been known for more than two
decades (Hertzberg et al., 1980), and its recognition as a
Activation of leuO transcription inhibits
member of the LysR family was published fifteen years ago
epithelial cell invasion
(Henikoff et al., 1988). Studies in a variety of gramnegative
During infection of animals, Salmonella Typhimurium bacteria have shown that LeuO regulates transcription of
invades epithelial cells using the SPI-1 type 3 secretion genes involved in diverse physiological processes, either
system. The observation that LeuO downregulates SPI as an activator or as a repressor (Shi and Bennett, 1995;
expression thus raised the question of whether activation Repoila and Gottesman, 2001; Chen and Wu, 2005;
of hilE transcription by LeuO might inhibit epithelial cell Stratmann et al., 2008; 2012; Shimada et al., 2009; 2011;
invasion. To test this possibility, assays of epithelial cell Dillon et al., 2012; Hernandez-Lucas and Calva, 2012). In
invasion were performed in vitro, and the invasion rates of serovars Typhi and Typhimurium of Salmonella enterica
T-POP leuO and T-POP ΔleuO strains were compared. As LeuO has been shown to regulate virulence-related
a control, a SPI-1 deletion mutant was included in the genes (Fernandez-Mora et al., 2004; Tenor et al., 2004;
assays. Activation of leuO transcription decreased inva- Rodriguez-Morales et al., 2006; Hernandez-Lucas et al.,
sion > 100-fold, and lack of LeuO restored the wild type 2008). Given these antecedents, the identification of LeuO
invasion rate (Fig. 7A). A HilE− mutant was more invasive binding sites in Salmonella pathogenicity island 1 (Dillon
than the wild type (Fig. 7B), an observation coherent with et al., 2012), combined with the observation that deletion of
the role of HilE as a negative regulator of SPI-1 (Fahlen leuO increased expression of the SPI-1 gene sopA (Dillon
et al., 2000; Baxter et al., 2003). In the absence of HilE, et al., 2012), suggested that LeuO might regulate SPI-1, a
activation of leuO transcription reduced epithelial cell inva- major determinant of Salmonella virulence (Galan and
sion three- to fourfold (Fig. 7B), thereby providing further Curtiss, 1989). Multiple controls adjust SPI-1 expression
evidence for HilE-independent downregulation of SPI-1 by to conditions that permit invasion of epithelial cells
LeuO. We thus conclude that LeuO inhibits invasion of (Altier, 2005; Jones, 2005; Ellermeier and Slauch, 2007).
epithelial cells by Salmonella enterica serovar Typhimu- Because activation of SPI-1 expression requires relief from
rium, and that the main inhibition ‘pathway’ requires HilE. H-NS-mediated silencing (Lucchini et al., 2006; Navarre
These conclusions are in agreement with the ability of et al., 2006) and LeuO acts often as an H-NS antagonist
LeuO to activate hilE transcription (Fig. 5). (Hernandez-Lucas et al., 2008; Shimada et al., 2009;

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1063

Fig. 6. Binding of LeuO-His6x to the hilE promoter region.


A. Slot blot binding assay. A 449 bp DNA fragment containing the P1, P2 and P3 hilE promoters was incubated with increasing concentrations
of LeuO-His6x.
B. Quantification of the binding assay. The fluorescence of the DNA bound to the membrane was represented versus the concentration of
LeuO-His6x protein.
C. Electrophoretic mobility shift assays of LeuOxHis6x binding to the hilE promoter region. PhilE designates a DNA fragment containing the
entire hilE promoter region, while P1, P2 and P3 designate DNA fragments containing individual hilE promoters. A DNA fragment containing
the S. enterica gene STM4318 was used as control. The concentrations of LeuO-His6x are indicated above each lane.
D. DNase I footprinting of LeuOxHis6x binding to the hilE promoter region using end-labelled linear DNA fragments. The size and quantity of
6-FAM-labelled digestion products were measured using a capillary electrophoresis DNA sequencing instrument.

Fig. 7. Inhibition of S. enterica invasion by LeuO.


A. Invasion of HeLa cells by the wild type (WT), a strain carrying a SPI-1 deletion (ΔSPI-1), a strain with leuO transcription driven by T-POP
(T-POP-leuO), and a leuO deletion mutant (T-POP ΔleuO).
B. Invasion of HeLa cells by the wild type (WT), a strain carrying a SPI-1 deletion (ΔSPI-1), a HilE− null mutant, (ΔhilE), a HilE− null mutant
with leuO transcription driven by T-POP (T-POP-leuO ΔhilE), and a strain lacking both LeuO and HilE (T-POP ΔleuO ΔhilE). The invasion rate
of the wild type was 0.1–0.3%, and was normalized to 1. Data are averages and standard deviations from > 6 independent experiments.
Because of the disparate invasion rates of the strains under study, different scales are used in each graph.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
1064 E. Espinosa and J. Casadesús ■

2011; Stratmann et al., 2012), the possibility that LeuO SPI-1 repressor, a tentative model is that activation of hilE
might repress SPI-1 was intriguing. transcription by LeuO might boost SPI-1 repression,
LeuO is a quiescent LTTR under standard laboratory perhaps under conditions in which H-NS fails to do so.
conditions because leuO transcription is repressed by Such hypothetical conditions can be expected to activate
H-NS (Klauck et al., 1997). To study LeuO-dependent leuO expression because leuO transcription is also
regulation of SPI-1 in S. enterica serovar Typhimurium, repressed by H-NS (Klauck et al., 1997). The existence of
transcription of the leuO gene was freed from H-NS a subordinate machinery to secure SPI-1 silencing might
repression by introducing a T-POP element upstream of have selective value because SPI-1 expression causes
leuO (Fig. S2). Activation of leuO transcription was found growth retardation (Sturm et al., 2010). Tight repression of
to downregulate genes belonging to independent tran- SPI-1 might have also selective value in environments
scriptional units within SPI-1, as well as the SPI-1- different from the animal intestine. For instance, coloniza-
controlled rtsA gene located outside SPI-1 (Fig. 1). These tion of plants by Salmonella is more efficient in the absence
experiments confirmed that LeuO represses SPI-1 of SPI-1 components, perhaps because the presence of
expression. A consequence of SPI-1 repression by LeuO the SPI-1 secretion apparatus in the Salmonella envelope
is reduced invasion of epithelial cells in vitro (Fig. 7). triggers a defence response by the plant (Iniguez et al.,
A genetic screen for loss-of-function mutations that 2005).
restored SPI-1 expression in the presence of LeuO pro- The view that LeuO may back up or replace H-NS to
vided evidence that the hilE gene is necessary for LeuO- silence SPI-1 is consistent with the existence of overlap-
mediated repression of SPI-1, an hypothesis confirmed ping controls that contribute to silencing of horizontally
upon directed construction of a HilE− mutant. The hilE acquired genes. For instance, the E. coli nucleoid-
gene is located outside SPI-1, and encodes a repressor of associated proteins Hha and YdgT enhance silencing of
SPI-1 expression (Baxter et al., 2003). HilE inactivates foreign genes by H-NS (Vivero et al., 2008; Banos et al.,
the transcriptional activator HilD (Baxter et al., 2003). 2009). Interestingly, YdgT and Hha appear to be redun-
HilD inactivation disrupts a positive feedback loop for dant, and YdgT has been proposed to act as a backup
hilD autogenous activation and causes SPI-1 repression molecule (Paytubi et al., 2004). The intricacy of accommo-
(Ellermeier et al., 2005). We thus considered that LeuO dating horizontally acquired genes in the host regulatory
might activate hilE transcription, and that HilE might network (Ochman et al., 2000; Lercher and Pal, 2008;
repress SPI-1 via HilD inactivation. Price et al., 2008) may confer adaptive value to redundant
The existence of a LeuO-HilE-HilD ‘pathway’ of SPI-1 control.
repression in S. enterica serovar Typhimurium is sup-
ported by several lines of evidence: (i) lack of HilE relieves
LeuO-mediated repression of SPI-1 (Fig. 2); (ii) activation Experimental procedures
of leuO transcription decreases the level of HilD protein Bacterial strains, media and culture conditions
(Fig. 3); (iii) the HilD protein decrease caused by LeuO is
suppressed by a hilE mutation (Fig. 3); and (iv) LeuO Strains of S. enterica serovar Typhimurium (henceforth abbre-
viated as S. enterica) used in this study are listed in Table 1,
activates transcription of hilE (Fig. 5).
and derive from ATCC 14028. An exception is SV6829 which
In the absence of HilE, however, LeuO remains able to
derives from SL1344. Strain SV6413 carries a hilD::lac
downregulate expression of SPI-1 (Fig. 4) and to reduce transcriptional fusion downstream from the hilD stop codon,
epithelial cell invasion (Fig. 7). HilD is likewise dispensa- constructed in such a way that the strain remains HilD+
ble for HilE-independent downregulation of SPI-1 by (Lopez-Garrido and Casadesus, 2012). The remaining chro-
LeuO (Fig. 4). However, HilE-independent SPI-1 repres- mosomal lac fusions are translational. E. coli DH5∝ [endA1
sion appears to be weak in comparison with the HilE- hsdR17 supE44 thi1 recA1 gyrA96 relA1 ΔlacU189 (Φ80
lacZΔM15)] (Hanahan, 1983) was used as the host of plas-
dependent ‘pathway’ (Figs 2 and 4). This conclusion is in
mids. E. coli BL21 [F− dcm ompT hsdS(rB− mB−) gal [malB+]K-
agreement with the observation that activation of leuO
12(λS)] is a product of Stratagene, La Jolla, CA.
transcription decreases epithelial cell invasion > 100-fold Transductional crosses using phage P22 HT 105/1 int201
in the presence of HilE and only three- to fourfold in the (Schmieger, 1972) were used for S. enterica strain construc-
absence of HilE (Fig. 7). tion operations involving chromosomal markers, and for
Our ignorance of natural conditions that permit leuO transfer of small plasmids between S. enterica strains. The
expression advises against interpretation of the physiologi- transduction protocol has been described elsewhere (Garzon
et al., 1995). To obtain phage-free isolates, transductants
cal significance of SPI-1 regulation by LeuO. However, a
were purified by streaking on green plates (Chan et al., 1972),
conceivable scenario is that LeuO might either backup
using methyl blue (Sigma-Aldrich, St Louis, MO) instead of
or relieve H-NS repression of certain loci, in a fashion aniline blue. Phage sensitivity was tested by cross-streaking
reminiscent of the HN-S-repressed VirT-VirB regulatory with the clear-plaque mutant P22 H5. Strain SV6141 was
cascade in Shigella (Tobe et al., 1993). Because HilE is a constructed by transducing the T-POP insertion of SV7424

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1065

Table 1. Strain list. (Dillon et al., 2012) to ATCC 14208. Tagging of proteins with 3×
FLAG or HA epitopes (strains SV6929 and SV7313) was
Strain Genotype performed as described by Lionello Bossi and co-workers
(Uzzau et al., 2001). Luria-Bertani (LB) broth was used as
SV5282a hilA::lacZ
standard liquid medium. Solid LB contained agar at 1.5%.
SV5295a invF::lacZ
SV5301a invH::lacZ Antibiotics were used at the concentrations described previ-
SV5384a hilC::lacZ ously (Torreblanca et al., 1999). The indicator for monitoring
SV5386a ΔhilD hilC::lacZ β-galactosidase activity in plate tests was 5-bromo-4-chloro-
SV5546 ΔhilD rtsA::lacZ 3-indolyl-β-D-galactopyranoside (‘Xgal’, Sigma Chemical Co.,
SV5584a rtsA::lacZ 40 μg ml−1). To activate leuO transcription, chlortetracycline
SV5586 hilE::Km
(5 μg ml−1) was added to the culture.
SV6055b SPI-1::Km
SV6141 T-POP-leuO
SV6142 T-POP-ΔleuO Construction of a T-POP ΔleuO mutant (SV6142)
SV6413e hilD930::lacZ
SV6829c leuO::3xFLAG Disruption and replacement of leuO with a chloramphenicol
SV6841 T-POP-leuO hilA::lacZ
resistance cassette was performed by lambda Red recom-
SV6842 T-POP-ΔleuO hilA::lacZ
SV6844 T-POP-leuO hilC::lacZ bineering (Datsenko and Wanner, 2000). Briefly, the Cmr
SV6845 T-POP-ΔleuO hilC::lacZ resistance cassette from plasmid pKD3 was PCR-amplified
SV6847 T-POP-leuO invF::lacZ with primers leuOP1 and leuOP2 (Table S1). The PCR
SV6848 T-POP-ΔleuO invF::lacZ product was used to transform an ATCC 14028 derivative
SV6850 T-POP-leuO hilD930::lacZ carrying the Red recombinase plasmid pKD46. Transfor-
SV6851 T-POP-ΔleuO hilD930::lacZ
mants were selected on LB with chloramphenicol.
SV6853 T-POP-leuO rtsA::lacZ
SV6854 T-POP-ΔleuO rtsA::lacZ
SV7034 T-POP-leuO invH::lacZ Construction of a sipB::GFP fusion (strain SV7975)
SV7035 T-POP-ΔleuO invH::lacZ
SV7036 T-pop leuO hilE::Tnd10dCm hilC::lacZ A fragment containing the promoterless green fluorescent
SV7037 T-POP-leuO ΔhilE protein (gfp) and the chloramphenicol resistance cassette
SV7038 T-POP-ΔleuO ΔhilE
SV7044 T-POP-leuO ΔhilD hilC::lacZ
was amplified from the pZEP07 plasmid using primers sipB-
SV7045 T-POP-ΔleuO ΔhilD hilC::lacZ GFP+-For and sipB-GFP+-Rev (Table S1). The 5′ regions of
SV7048 T-POP-leuO ΔhilD rtsA::lacZ these primers are homologous to the 3′ untranslated region of
SV7049 T-POP-ΔleuO ΔhilD rtsA::lacZ the S. enterica sipB gene, so that the fusion is formed down-
SV7312 T-POP-leuO hilD::HA stream of the sipB stop codon. The construct was integrated
SV7313 T-POP-ΔleuO hilD::HA into the chromosome of S. enterica using the Lambda Red
SV7315 T-POP-leuO ΔhilE hilD::HA
SV7316 T-POP-ΔleuO ΔhilE hilD::HA
recombination system.
SV7317 ΔhilD ΔhilE rtsA::lacZ
SV7318 T-POP-leuO ΔhilD ΔhilE rtsA::lacZ Construction of a hilE mutant (SV5586)
SV7319 T-POP-ΔleuO ΔhilD ΔhilE rtsA::lacZ
SV7320 ΔhilD ΔhilE hilC::lacZ Disruption and replacement of hilE with a Kmr cassette was
SV7321 T-POP-leuO ΔhilD ΔhilE hilC::lacZ
SV7322 T-POP-ΔleuO ΔhilD ΔhilE hilC::lacZ
performed by lambda Red recombineering (Datsenko and
SV7327 T-POP-leuO hilE::lacZ Wanner, 2000). The Kmr resistance cassette of plasmid
SV7328 T-POP-ΔleuO hilE::lacZ pKD13 was PCR-amplified with primers hilEPS1 and hilEPS4
SV7741 T-POP leuO/pIC552 (Table S1). The PCR product was used to transform the
SV7742 T-POP ΔleuO/pIC552 wild-type strain carrying pKD46. Transformants were
SV7743 T-POP leuO/pIZ1997 selected on LB plates with kanamycin. The antibiotic resist-
SV7744 T-POP ΔleuO/pIZ1997
SV7745 T-POP leuO/pIZ1998
ance casette was excised by recombination with plasmid
SV7746 T-POP ΔleuO/pIZ1998 pCP20 (Datsenko and Wanner, 2000). For the construction of
SV7747 T-POP leuO/pIZ1999 a translational hilE::lac fusion on the Salmonella chromo-
SV7748 T-POP ΔleuO/pIZ1999 some (strain SV7327), FRT sites generated by excision of
SV7749 T-POP leuO/pIZ2000 Kmr cassette were used to integrate plasmid pCE40
SV7750 T-POP ΔleuO/pIZ2000 (Ellermeier et al., 2002).
SV7975d sipB::GFP
SV7976 ΔhilE sipB::GFP
SV7977 T-POP leuO sipB::GFP Screen for suppressors of SPI-1 downregulation in a
SV7978 T-POP ΔleuO sipB::GFP
LeuO-constitutive strain
SV7979 ΔhilE T-POP leuO sipB::GFP
SV7980 ΔhilE T-POP ΔleuO sipB::GFP
Strain SV6842 (hilC::lac T-POP-leuO) was transduced with a
a. Constructed by J. López-Garrido. Tn10dCm pool, and transductants were selected on LB
b. Constructed by F. Baisón-Olmo. plates supplemented with chloramphenicol and Xgal. Inde-
c. SL1344 derivative. pendent Lac+ transductants were sought and purified on
d. Constructed by I. Cota and S. B. Hernández. green plates. To characterize Tn10dCm insertions, a semi-
e. Strain described by (Lopez-Garrido and Casadesus, 2012). random, two step PCR protocol (Chun et al., 1997) was used
to amplify genomic regions adjacent to the Tn10dCm

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
1066 E. Espinosa and J. Casadesús ■

element. The reactions were carried out as described else- events per sample, and were analysed with CXP and FlowJo
where (Baison-Olmo et al., 2012), using primers listed in 8.7 Software.
Table S1. Amplification products were sequenced using
primers ST-PCR-Cm2-Int and st1. The Tn10dCm insertion
Quantitative reverse transcriptase PCR
had occurred at a TGCT/GCCA motif in the hilE coding
sequence. RNA was extracted from stationary phase cultures
(OD600 ∼ 2). Cells were centrifuged at 13 000 rpm at 4°C and
Plasmids resuspended in 100 μl of lysozyme (Sigma Chemical)
(3 mg ml−1). RNA was extracted using Trizol (Invitrogen,
Plasmid pZEP07 (Hautefort et al., 2003) was used to Carlsbad, CA), according to manufacturer’s instructions. The
construct a sipB::GFP fusion. Plasmid pIZ1871 is a pET21a quality and quantity of the extracted RNA was determined
derivative (Novagen EMD4 Biosciences, Darmstadt, using a ND-1000 spectrophotometer (Nanodrop Technolo-
Germany) containing leuO-His6x (Dillon et al., 2012). Plas- gies, Wilmington, DE). To diminish genomic DNA contamina-
mids carrying transcriptional lacZ fusions are derivatives of tion and to obtain cDNA, 1 μg RNA was treated with the
the promoter-probe vector pIC552 (Macian et al., 1994). To Quantitec Reverse Transcription Kit (Quiagen, Venlo, The
construct plasmid-borne lac fusions on pIC552, DNA from Netherlands). Quantitative RT-PCR reactions were per-
strain ATCC 14028 was used as template for PCR amplifica- formed in LightCycler 480 II (Roche). Each reaction was
tion with primers listed in Table S1. The amplified fragments carried out in a total volume of 10 μl on a 480-well optical
were digested with BglII and XhoI and ligated with BglII/XhoI- reaction plate (Roche, Basel, Switzerland) containing 5 μl
digested pIC552. The DNA fragments from the hilE promoter SYBR, 0.5 DYE II (Takara Bio, Otsu, Japan), 4 μl cDNA (1/50)
region carried by pIC552 derivatives are as follows: pIZ1997, and two gene-specific primers at a final concentration of
−505 to −56; pIZ1998, −159 to −56; pIZ1999, −334 to −161; 0.2 mM each. Real-time cycling conditions were as follows:
and pIZ2000, −505 to −336. (i) 95°C for 10 min; (ii) 40 cycles at 95°C for 15 s, 60°C for
1 min. A non-template control was included for each primer
β-Galactosidase assays set. Melting curve analysis verified that each reaction con-
tained a single PCR product. Gene expression levels were
Bacterial cultures were grown in LB + chlortetracycline until normalized to transcripts of gmk that served as an internal
stationary phase (OD600 ≤ 2). Levels of β-galactosidase activ- control. Gene-specific primers were designed with PRIMER3
ity were assayed using the CHCl3-sodium dodecyl sulphate software (http://primer3.sourceforge.net) and are listed in
permeabilization procedure (Miller, 1972). Table S1.

Western immunoblot assays Purification of LeuO protein

Whole cell protein extracts were prepared from bacterial cul- For LeuO purification, plasmid pIZ1871 was used (Dillon et al.,
tures in LB broth containing chlortetracycline. Cultures were 2012). E. coli BL21/pIZ1871 was grown in LB broth, and
grown at 37°C until stationary phase (OD600 ∼ 2). Bacterial expression of leuO was induced with 1 mM isopropylβ-D-thio-
cells contained in 1 ml of culture were collected by centrifu- galactopyranoside (IPTG). After 4 h induction, cells were cen-
gation (16 000 g, 2 min, 4°C) and suspended in 50 μl of trifuged and resuspended in lysis buffer [20 mM Tris, 300 mM
Laemmli sample buffer [1.3% sodium dodecyl sulphate, 10% NaCl, 10 mM imidazole, 1 mM phenylmethylsulphonyl fluoride
(v/v) glycerol, 50 mM tris(hydroxymethyl)aminomethane-HCl (PMSF), 5 μl ml−1 protein inhibitor cocktail (Sigma-Aldrich, St
(Tris-HCl), 1.8% β-mercaptoethanol, 0.02% bromophenol Louis, MO)], and lysed by sonication. The suspension was
blue, pH 6.8]. Proteins were resolved by Tris-Tricine-PAGE centrifuged at 10 000 rpm for 30 min. LeuO-His6x was purified
using 12% gels. Conditions for protein transfer have been by nickel affinity chromatography (GE Healthcare). The
described elsewhere (Balbontin et al., 2006). Primary anti- column was washed once with 5 ml of washing buffer (20 mM
bodies were anti-FLAG M2 monoclonal antibody (1:5000, Tris, 300 mM NaCl, 50 mM imidazole), and twice with the
Sigma Chemical Co, St Louis, MO) and anti-DnaK monoclo- same buffer containing 75 mM and 100 mM of imidazole.
nal antibody (1:5000, MBL International, MA). Goat anti- Protein elution was performed with 4 ml of elution buffer
mouse horseradish peroxidase-conjugated antibody (1:5000, (20 mM Tris, 300 mM NaCl, 300 mM imidazole). Imidazole
Bio-Rad, Hercules, CA) was used as secondary antibody. was removed and the protein was concentrated by washing
Proteins recognized by the antibodies were visualized by with storage buffer (20 mM Tris, 300 mM NaCl, 10% glycerol)
chemiluminiscence using luciferin–luminol. and centrifuged using Amicon® Ultra centrifugal filters. LeuO-
His6x protein was stored at −80°C.
Flow cytometry
Slot blot and gel mobility shift assays of LeuO binding
Bacterial cultures were grown at 37°C in LB + chlortetracy-
cline until stationary phase (OD600 ∼ 2). Cells were washed DNA probes labelled with 6-caroxyfluorescein (FAM) were
and re-suspended in PBS to a final concentration of prepared by PCR amplification using primer pairs listed in
5 × 106 cells ml−1. Data acquisition and analysis were per- Table S1. PCR products were purified with the Wizard® SV
formed using a Cytomics FC500-MPL cytometer (Beckman Clean-Up-System (Promega, Madison, WI). Ten nanograms
Coulter, Brea, CA). Approximately 5 × 106 cells were ana- of FAM-labelled probe were used for slot blot assays, and
lysed for GFP expression. Data were collected for 40 000 25 ng for electrophoretic gel mobility shift assays. In both

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1067

types of experiments, the FAM-labelled probes were incu- lar of bacteria were obtained after lysis of infected cells with
bated at room temperature for 30 min with increasing con- 1% Triton X-100 and plating on appropriate media. Strain
centrations of LeuO-6xHis in a final volume of 20 μl. The binding discrimination was performed as for the input. Infections were
buffer L10× contained 20 mM HEPES, 100 mM KCl, 2 mM carried out in triplicate. The invasion rate was defined as the
MgCl2, 0,1 mM EDTA and 20% glycerol (De la Cruz et al., number of bacteria in the output (intracellular bacteria recov-
2007). For slot blot binding assays, protein-DNA complexes ered after 2 h of infection) divided by the number of bacteria
were diluted 1:10 with PBS and blotted onto nitrocellulose in the input (initial inoculum).
filters (AmershamTM HybondTM-ECL, GE Healthcare) using a
PR 600 Slot Blot Manifold (Hoefer Scientific Instruments, San
Francisco, CA) connected to a portable vacuum/pressure Acknowledgements
pump (Merck Millipore, Darmstadt, Germany). Wells were
then washed five times with PBS and membranes were air- This study was supported by grants BIO2010–15023 and
dried. DNA fragments were visualized with a FLA-5100 CSD2008-00013 from the Spanish Ministry of Economy and
Imaging system (Fujifilm, Tokyo, Japan). For gel shift assays, Competitivity (MINECO) and the European Regional Fund.
protein-DNA complexes were subjected to electrophoresis at We are grateful to Charles Dorman, Shane Dillon, Francisco
4°C in a 6% non-denaturing acrylamide : bisacrylamide Ramos-Morales and Elena Cardenal-Muñoz for discussions,
(29:1) gel in 0.5 Tris-borate-EDTA buffer. to Javier López-Garrido, Fernando Baisón-Olmo, Ignacio
Cota, and Sara B. Hernández for strain constructions, to
María Antonia Sánchez-Romero for advice in flow cytometry
DNase I footprinting of LeuO binding to experiments, and to Modesto Carballo, Laura Navarro, and
PCR-amplified DNA Cristina Reyes of the Servicio de Biología, Centro de Inves-
tigación, Tecnología e Innovación de la Universidad de
A PhilE DNA probe labelled with 6-caroxyfluorescein (FAM) Sevilla (CITIUS) for help with experiments performed at the
was prepared by PCR amplification using primers listed in facility.
Table S1. DNase I footprinting was performed as described
elsewhere (Cameron and Dorman, 2012) with minor modifi-
cations. DNase I footprinting reactions were performed in References
15 μl reaction volumes containing 1× LeuO binding buffer
(2 mM HEPES, 10 mM KCl, 0.2 mM MgCl2, 0.01 mM EDTA Altier, C. (2005) Genetic and environmental control of salmo-
and 2% glycerol), 0.01 mM DTT, 100 ng μl−1 bovine serum nella invasion. J Microbiol 43 (Spec. No.): 85–92.
albumin (BSA), 50 nM bait DNA, and 4 μM LeuO-His6x. LeuO- Bailly-Bechet, M., Benecke, A., Hardt, W.D., Lanza, V.,
DNA binding was allowed to equilibrate at room temperature Sturm, A., and Zecchina, R. (2011) An externally modu-
for 30 min. One μl (0.05 units) of DNase I (Roche) was then lated, noise-driven switch for the regulation of SPI1 in
added, mixed gently, and incubated at room temperature for Salmonella enterica serovar Typhimurium. J Math Biol 63:
5 min. Reactions were stopped by addition of 2 μl EDTA 637–662.
(100 mM) followed by vigorous vortexing and heat denatura- Baison-Olmo, F., Cardenal-Munoz, E., and Ramos-Morales,
tion at 95°C for 10 min. Digestion products were desalted F. (2012) PipB2 is a substrate of the Salmonella patho-
using MicroSpin G-25 columns (GE Healthcare, Sevilla, genicity island 1-encoded type III secretion system.
Spain), and were analysed on an ABI 3730 DNA Analyzer Biochem Biophys Res Commun 423: 240–246.
along with GeneScan 500-LIZ size standards (Applied Bio- Balbontin, R., Rowley, G., Pucciarelli, M.G., Lopez-Garrido,
systems, Alcobendas, Spain). J., Wormstone, Y., Lucchini, S., et al. (2006) DNA adenine
methylation regulates virulence gene expression in Salmo-
nella enterica serovar Typhimurium. J Bacteriol 188: 8160–
Invasion assays in HeLa epithelial cells 8168.
Banos, R.C., Vivero, A., Aznar, S., Garcia, J., Pons, M.,
HeLa human epithelial cells (ATCC CCL2) were grown in Madrid, C., and Juarez, A. (2009) Differential regulation of
Dulbecco’s modified Eagle medium (DMEM) containing 10% horizontally acquired and core genome genes by the bac-
fetal calf serum and 1 mM glutamine (Life Technologies, Prat terial modulator H-NS. PLoS Genet 5: e1000513.
de Llobregat, Spain). HeLa cells were seeded in 24-well Baxter, M.A., Fahlen, T.F., Wilson, R.L., and Jones, B.D.
plates (Costar, Corning, New York, NY) the day before infec- (2003) HilE interacts with HilD and negatively regulates
tion, and were grown at 37°C with 5% CO2. Bacteria were hilA transcription and expression of the Salmonella
grown overnight in LB broth containing NaCl 0.3 M and chlo- enterica serovar Typhimurium invasive phenotype. Infect
rtetracycline 5 μg ml−1 at 37°C without shaking. An aliquot Immun 71: 1295–1305.
from the bacterial culture was added to HeLa cells to reach a Cameron, A.D., and Dorman, C.J. (2012) A fundamental
multiplicity of infection of 50 bacteria per eukaryotic cell. regulatory mechanism operating through OmpR and DNA
Bacterial dilutions were prepared in DMEM. The colony- topology controls expression of Salmonella pathogenicity
forming units of the strains in the input were enumerated by islands SPI-1 and SPI-2. PLoS Genet 8: e1002615.
plating a dilution series of the inoculum, using the appropriate Chan, R.K., Botstein, D., Watanabe, T., and Ogata, Y. (1972)
antibiotic to distinguish the strains. Thirty minutes after inva- Specialized transduction of tetracycline by phage P22 in
sion, the cells were washed twice with phosphate-buffered Salmonella typhimurium. II. Properties of a high frequency
saline (PBS) and incubated for 90 min in fresh DMEM con- transducing lysate. Virology 50: 883–898.
taining 100 μg ml−1 gentamicin. Numbers of viable intracellu- Chen, C.C., and Wu, H.Y. (2005) LeuO protein delimits the

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
1068 E. Espinosa and J. Casadesús ■

transcriptionally active and repressive domains on the bac- Fortune, D.R., Suyemoto, M., and Altier, C. (2006) Identifica-
terial chromosome. J Biol Chem 280: 15111–15121. tion of CsrC and characterization of its role in epithelial cell
Chen, C.C., Chou, M.Y., Huang, C.H., Majumder, A., and Wu, invasion in Salmonella enterica serovar Typhimurium.
H.Y. (2005) A cis-spreading nucleoprotein filament is Infect Immun 74: 331–339.
responsible for the gene silencing activity found in the Galan, J.E., and Curtiss, R., 3rd (1989) Cloning and molecu-
promoter relay mechanism. J Biol Chem 280: 5101–5112. lar characterization of genes whose products allow Salmo-
Chun, K.T., Edenberg, H.J., Kelley, M.R., and Goebl, M.G. nella typhimurium to penetrate tissue culture cells. Proc
(1997) Rapid amplification of uncharacterized transposon- Natl Acad Sci USA 86: 6383–6387.
tagged DNA sequences from genomic DNA. Yeast 13: Gallego-Hernandez, A.L., Hernandez-Lucas, I., De la Cruz,
233–240. M.A., Olvera, L., Morett, E., Medina-Aparicio, L., et al.
Datsenko, K.A., and Wanner, B.L. (2000) One-step inactiva- (2012) Transcriptional regulation of the assT-dsbL-dsbI
tion of chromosomal genes in Escherichia coli K-12 using gene cluster in Salmonella enterica serovar Typhi IMSS-1
PCR products. Proc Natl Acad Sci USA 90: 6640–6645. depends on LeuO, H-NS, and specific growth conditions. J
De la Cruz, M.A., Fernandez-Mora, M., Guadarrama, C., Bacteriol 194: 2254–2264.
Flores-Valdez, M.A., Bustamante, V.H., Vazquez, A., and Garzon, A., Cano, D.A., and Casadesus, J. (1995) Role of Erf
Calva, E. (2007) LeuO antagonizes H-NS and StpA- recombinase in P22-mediated plasmid transduction.
dependent repression in Salmonella enterica ompS1. Mol Genetics 140: 427–434.
Microbiol 66: 727–743. Hanahan, D. (1983) Studies on transformation of Escherichia
Dillon, S.C., Espinosa, E., Hokamp, K., Ussery, D.W., coli with plasmids. J Mol Biol 166: 557–580.
Casadesus, J., and Dorman, C.J. (2012) LeuO is a global Haughn, G.W., Wessler, S.R., Gemmill, R.M., and Calvo,
regulator of gene expression in Salmonella enterica J.M. (1986) High A+T content conserved in DNA
serovar Typhimurium. Mol Microbiol 85: 1072–1089. sequences upstream of leuABCD in Escherichia coli and
Dorman, C.J. (2007) H-NS, the genome sentinel. Nat Rev Salmonella typhimurium. J Bacteriol 166: 1113–1117.
Microbiol 5: 157–161. Hautefort, I., Proenca, M.J., and Hinton, J.C. (2003) Single-
Ellermeier, C.D., and Slauch, J.M. (2003) RtsA and RtsB copy green fluorescent protein gene fusions allow accurate
coordinately regulate expression of the invasion and flagel- measurement of Salmonella gene expression in vitro and
lar genes in Salmonella enterica serovar Typhimurium. J during infection of mammalian cells. Appl Environ Microbiol
Bacteriol 185: 5096–5108. 69: 7480–7491.
Ellermeier, C.D., Janakiram, A., and Slauch, J.M. (2002) Henikoff, S., Haughn, G.W., Calvo, J.M., and Wallace, J.C.
Construction of targeted single copy lac fusions using (1988) A large family of bacterial activator proteins. Proc
lambda Red and FLP-mediated site-specific recombination Natl Acad Sci USA 85: 6602–6606.
in bacteria. Gene 290: 153–161. Hernandez-Lucas, I., and Calva, E. (2012) The coming of age
Ellermeier, C.D., Ellermeier, J.R., and Slauch, J.M. (2005) of the LeuO regulator. Mol Microbiol 85: 1026–1028.
HilD, HilC and RtsA constitute a feed forward loop that Hernandez-Lucas, I., Gallego-Hernandez, A.L., Encarnacion,
controls expression of the SPI1 type three secretion S., Fernandez-Mora, M., Martinez-Batallar, A.G., Salgado,
system regulator hilA in Salmonella enterica serovar Typh- H., et al. (2008) The LysR-type transcriptional regulator
imurium. Mol Microbiol 57: 691–705. LeuO controls expression of several genes in Salmonella
Ellermeier, J.R., and Slauch, J.M. (2007) Adaptation to the enterica serovar Typhi. J Bacteriol 190: 1658–1670.
host environment: regulation of the SPI1 type III secretion Hertzberg, K.M., Gemmill, R., Jones, J., and Calvo, J.M.
system in Salmonella enterica serovar Typhimurium. Curr (1980) Cloning of an EcoRI-generated fragment of the
Opin Microbiol 10: 24–29. leucine operon of Salmonella typhimurium. Gene 8: 135–
Ellermeier, J.R., and Slauch, J.M. (2008) Fur regulates 152.
expression of the Salmonella pathogenicity island 1 type III Iniguez, A.L., Dong, Y., Carter, H.D., Ahmer, B.M., Stone,
secretion system through HilD. J Bacteriol 190: 476–486. J.M., and Triplett, E.W. (2005) Regulation of enteric endo-
Fahlen, T.F., Mathur, N., and Jones, B.D. (2000) Identification phytic bacterial colonization by plant defenses. Mol Plant
and characterization of mutants with increased expression Microbe Interact 18: 169–178.
of hilA, the invasion gene transcriptional activator of Sal- Jones, B.D. (2005) Salmonella invasion gene regulation: a
monella typhimurium. FEMS Immunol Med Microbiol 28: story of environmental awareness. J Microbiol 43 (Spec.
25–35. No.): 110–117.
Fang, M., and Wu, H.Y. (1998a) A promoter relay mechanism Klauck, E., Bohringer, J., and Hengge-Aronis, R. (1997) The
for sequential gene activation. J Bacteriol 180: 626–633. LysR-like regulator LeuO in Escherichia coli is involved in
Fang, M., and Wu, H.Y. (1998b) Suppression of leu-500 the translational regulation of rpoS by affecting the expres-
mutation in topA+ Salmonella typhimurium strains. The pro- sion of the small regulatory DsrA-RNA. Mol Microbiol 25:
moter relay at work. J Biol Chem 273: 29929–29934. 559–569.
Fang, M., Majumder, A., Tsai, K.J., and Wu, H.Y. (2000) Lawley, T.D., Chan, K., Thompson, L.J., Kim, C.C., Govoni,
ppGpp-dependent leuO expression in bacteria under G.R., and Monack, D.M. (2006) Genome-wide screen for
stress. Biochem Biophys Res Commun 276: 64–70. Salmonella genes required for long-term systemic infection
Fernandez-Mora, M., Puente, J.L., and Calva, E. (2004) of the mouse. PLoS Pathog 2: e11.
OmpR and LeuO positively regulate the Salmonella Lee, C., Wozniak, C., Karlinsey, J.E., and Hughes, K.T.
enterica serovar Typhi ompS2 porin gene. J Bacteriol 186: (2007) Genomic screening for regulatory genes using the
2909–2920. T-POP transposon. Methods Enzymol 421: 159–167.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069
Regulation of SPI-1 by LeuO 1069

Lercher, M.J., and Pal, C. (2008) Integration of horizontally ompS2 mutants are attenuated for virulence in mice. Infect
transferred genes into regulatory interaction networks Immun 74: 1398–1402.
takes many million years. Mol Biol Evol 25: 559–567. Schmieger, H. (1972) Phage P22 mutants with increased or
Lim, S., Yun, J., Yoon, H., Park, C., Kim, B., Jeon, B., et al. decreased transducing abilities. Mol Gen Genet 119:
(2007) Mlc regulation of Salmonella pathogenicity island I 75–88.
gene expression via hilE repression. Nucleic Acids Res 35: Shi, X., and Bennett, G.N. (1995) Effects of multicopy LeuO
1822–1832. on the expression of the acid-inducible lysine decarboxy-
Lopez-Garrido, J., and Casadesus, J. (2012) Crosstalk lase gene in Escherichia coli. J Bacteriol 177: 810–814.
between virulence loci: regulation of Salmonella enterica Shimada, T., Yamamoto, K., and Ishihama, A. (2009) Involve-
pathogenicity island 1 (SPI-1) by products of the std fim- ment of the leucine response transcription factor LeuO in
brial operon. PLoS ONE 7: e30499. regulation of the genes for sulfa drug efflux. J Bacteriol
Lostroh, C.P., and Lee, C.A. (2001) The Salmonella patho- 191: 4562–4571.
genicity island-1 type III secretion system. Microbes Infect Shimada, T., Bridier, A., Briandet, R., and Ishihama, A. (2011)
3: 1281–1291. Novel roles of LeuO in transcription regulation of E. coli
Lucchini, S., Rowley, G., Goldberg, M.D., Hurd, D., Harrison, genome: antagonistic interplay with the universal silencer
M., and Hinton, J.C. (2006) H-NS mediates the silencing of H-NS. Mol Microbiol 82: 378–397.
laterally acquired genes in bacteria. PLoS Pathog 2: e81. Stratmann, T., Madhusudan, S., and Schnetz, K. (2008)
Macian, F., Perez-Roger, I., and Armengod, M.E. (1994) An Regulation of the yjjQ-bglJ operon, encoding LuxR-type
improved vector system for constructing transcriptional transcription factors, and the divergent yjjP gene by H-NS
lacZ fusions: analysis of regulation of the dnaA, dnaN, recF and LeuO. J Bacteriol 190: 926–935.
and gyrB genes of Escherichia coli. Gene 145: 17–24. Stratmann, T., Pul, U., Wurm, R., Wagner, R., and Schnetz,
Maddocks, S.E., and Oyston, P.C. (2008) Structure and func- K. (2012) RcsB-BglJ activates the Escherichia coli leuO
tion of the LysR-type transcriptional regulator (LTTR) family gene, encoding an H-NS antagonist and pleiotropic regu-
proteins. Microbiology 154: 3609–3623. lator of virulence determinants. Mol Microbiol 83: 1109–
Majumder, A., Fang, M., Tsai, K.J., Ueguchi, C., Mizuno, T., 1123.
and Wu, H.Y. (2001) LeuO expression in response to star- Sturm, A., Heinemann, M., Arnoldini, M., Benecke, A.,
vation for branched-chain amino acids. J Biol Chem 276: Ackermann, M., Benz, M., et al. (2010) The cost of viru-
19046–19051. lence: retarded growth of Salmonella Typhimurium cells
Miller, J.H. (1972) Experiments in Molecular Genetics. Cold expressing type III secretion system 1. PLoS Pathog 7:
Spring Harbor, NY: Cold Spring Harbor Laboratory Press. e1002143.
Momany, C., and Neidle, E.L. (2012) Defying stereotypes: Tenor, J.L., McCormick, B.A., Ausubel, F.M., and Aballay, A.
the elusive search for a universal model of LysR-type regu- (2004) Caenorhabditis elegans-based screen identifies
lation. Mol Microbiol 83: 453–456. Salmonella virulence factors required for conserved host-
Moorthy, S., and Watnick, P.I. (2005) Identification of novel pathogen interactions. Curr Biol 14: 1018–1024.
stage-specific genetic requirements through whole Tobe, T., Yoshikawa, M., Mizuno, T., and Sasakawa, C.
genome transcription profiling of Vibrio cholerae biofilm (1993) Transcriptional control of the invasion regulatory
development. Mol Microbiol 57: 1623–1635. gene virB of Shigella flexneri: activation by virF and repres-
Navarre, W.W., Porwollik, S., Wang, Y., McClelland, M., sion by H-NS. J Bacteriol 175: 6142–6149.
Rosen, H., Libby, S.J., and Fang, F.C. (2006) Selective Torreblanca, J., Marques, S., and Casadesus, J. (1999) Syn-
silencing of foreign DNA with low GC content by the H-NS thesis of FinP RNA by plasmids F and pSLT is regulated by
protein in Salmonella. Science 313: 236–238. DNA adenine methylation. Genetics 152: 31–45.
Ochman, H., Lawrence, J.G., and Groisman, E.A. (2000) Uzzau, S., Figueroa-Bossi, N., Rubino, S., and Bossi, L.
Lateral gene transfer and the nature of bacterial innovation. (2001) Epitope tagging of chromosomal genes in Salmo-
Nature 405: 299–304. nella. Proc Natl Acad Sci USA 98: 15264–15269.
Paytubi, S., Madrid, C., Forns, N., Nieto, J.M., Balsalobre, C., Vivero, A., Banos, R.C., Mariscotti, J.F., Oliveros, J.C.,
Uhlin, B.E., and Juarez, A. (2004) YdgT, the Hha paralogue Garcia-del Portillo, F., Juarez, A., and Madrid, C. (2008)
in Escherichia coli, forms heteromeric complexes with Modulation of horizontally acquired genes by the Hha-
H-NS and StpA. Mol Microbiol 54: 251–263. YdgT proteins in Salmonella enterica serovar Typhimu-
Price, M.N., Dehal, P.S., and Arkin, A.P. (2008) Horizontal rium. J Bacteriol 190: 1152–1156.
gene transfer and the evolution of transcriptional regulation Wilmes-Riesenberg, M.R., Foster, J.W., and Curtiss, R., 3rd
in Escherichia coli. Genome Biol 9: R4. (1997) An altered rpoS allele contributes to the avirulence
Rappleye, C.A., and Roth, J.R. (1997) A Tn10 derivative of Salmonella typhimurium LT2. Infect Immun 65: 203–210.
(T-POP) for isolation of insertions with conditional Zaim, J., and Kierzek, A.M. (2003) The structure of full-length
(tetracycline-dependent) phenotypes. J Bacteriol 179: LysR-type transcriptional regulators. Modeling of the full-
5827–5834. length OxyR transcription factor dimer. Nucleic Acids Res
Repoila, F., and Gottesman, S. (2001) Signal transduction 31: 1444–1454.
cascade for regulation of RpoS: temperature regulation of
DsrA. J Bacteriol 183: 4012–4023. Supporting information
Rodriguez-Morales, O., Fernandez-Mora, M., Hernandez-
Lucas, I., Vazquez, A., Puente, J.L., and Calva, E. (2006) Additional supporting information may be found in the online
Salmonella enterica serovar Typhimurium ompS1 and version of this article at the publisher’s web-site.

© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 91, 1057–1069

You might also like