ArticlePDF Available

Multiple substrates of the Legionella pneumophila Dot/Icm system identified by interbacterial protein transfer

Authors:

Abstract and Figures

Legionella pneumophila is an intracellular pathogen that multiplies in a specialized vacuole within host cells. Biogenesis of this vacuole requires the Dot/Icm type IV protein translocation system. By using a Cre/loxP-based protein translocation assay, we found that proteins translocated by the Dot/Icm complex across the host phagosomal membrane can also be transferred from one bacterial cell to another. The flexibility of this system allowed the identification of several families of proteins translocated by the Dot/Icm complex. When analyzed by immunofluorescence microscopy, a protein identified by this procedure, SidC, was shown to translocate across the phagosomal membranes to the cytoplasmic face of the L. pneumophila phagosome. The identification of large numbers of these substrates, and the fact that the absence of any one substrate rarely results in strong defects in intracellular growth, indicate that there is significant functional redundancy among the Dot/Icm translocation targets. • bacterial pathogenesis • protein translocation • plasmid conjugation
Interbacterial protein translocation by the Dot ͞ Icm system in the absence of DNA transfer. ( A ) Assay for interbacterial protein transfer. Translocation of Cre hybrid protein from a donor bacterial strain is measured by removal of a fl oxed transcriptional terminator located between the trc promoter and the npt II (kan R ) gene on plasmid pZL184 harbored by a recipient bacterial strain. Bacteria harboring the intact reporter are unable to grow on media containing kanamycin and sucrose. The translocation of Cre hybrid protein into the recipient strain leads to the excision, through recombination at the loxP sites, of the DNA fragment that confers sucrose sensitivity ( sacB ) and reconstitution of a functional loxP-npt II translational fusion. S.D., Shine – Dalgarno sequence; npt II , neomycin phosphotransferase; red diamond, transcriptional terminator. Arrows indicate the trc promoter and loxP sites. ( B ) Interbacterial transfer of a fusion derived from the RSF1010 mobA gene. The mobA gene was fused to cre , and RP4-dependent protein translocation into a recipient E. coli strain was measured by using E. coli S17 – 1 (15) as the donor selecting kanamycin resistance and screening for gentamicin sensitivity. Black bars, S17 – 1 Trb ϩ donor; stippled bars, E. coli DH5 ␣ Trb Ϫ donor. ( C ). Plasmids harboring Cre fusions cannot be transferred to recipient cells. Plasmids expressing the designated proteins were harbored in either E. coli S17 – 1 (Trb ϩ ) or L. pneumophila Lp02 (Dot ͞ Icm ϩ ; ref. 11), and the ef fi ciency of plasmid transfer was measured by using either recipient E. coli or L. pneumophila strains, respectively. As a positive control, the identical plasmids having an intact oriT and mob system were used to demonstrate transfer pro fi ciency of donor strains. Black bars, donor strain E. coli S17 – 1 (Trb ϩ ); gray bars, donor strain L. pneumophila Lp02 (Dot ͞ Icm ϩ ; ref. 11). ( D ) Transfer of
… 
SidC is translocated by the L. pneumophila DotIcm system to the host cell and is localized about the phagosomal membrane. (A) Bone marrowderived macrophages from AJ mice were infected with Lp02(doticm intact), Lp03(dotA ), or Lp02(sdcA-sidC) strain expressing GFP, respectively. One hour after infection, cells were fixed as described (10), and SidC was probed with anti-(His) 6-SidC antibodies and Texas red-labeled secondary antibodies. Stained macrophages were scored for translocation of SidC by counting phagosomes that stained positively with anti-(His) 6-SidC. Data shown are from two independent experiments performed in triplicate in which at least 100 phagosomes were scored per coverslip. (B) SidC staining on PNS prepared from L. pneumophila-infected U937 cells in the absence of permeabilization. Sample preparation, immunostaining, and data collection were performed as described in ref. 10 or in A. (C) DotA-dependent translocation of SidC. (Left) Bacteria expressing GFP associated with bone marrow-derived macrophage. Strains used were Lp02(doticm intact; Upper) and Lp03 (dotA ; Lower). (Center) Immunoprobing of infected cells with anti-(His) 6-SidC. (Right) Merged images of GFP and anti-(His) 6-SidC staining. (D) Limited diffusion of SidC from L. pneumophila phagosome. Shown are images of Lp02(doticm intact) with murine bone borrow-derived macrophage. (E) SidC is translocated across the phagosomal membrane. Shown are images of PNSs of Lp02(doticm intact)-infected macrophages. Bacteria and SidC are probed as above, with bacteria marked by GFP and anti-SidC marked in red.
… 
Content may be subject to copyright.
Multiple substrates of the
Legionella pneumophila
DotIcm system identified by interbacterial
protein transfer
Zhao-Qing Luo and Ralph R. Isberg*
Howard Hughes Medical Institute and Department of Molecular Biology and Microbiology, Tufts University School of Medicine, 150 Harrison Avenue,
Boston, MA 02111
Edited by John J. Mekalanos, Harvard Medical School, Boston, MA, and approved November 21, 2003 (received for review August 2, 2003)
Legionella pneumophila is an intracellular pathogen that multiplies
in a specialized vacuole within host cells. Biogenesis of this vacuole
requires the DotIcm type IV protein translocation system. By using
a CreloxP-based protein translocation assay, we found that pro-
teins translocated by the DotIcm complex across the host phago-
somal membrane can also be transferred from one bacterial cell to
another. The flexibility of this system allowed the identification of
several families of proteins translocated by the DotIcm complex.
When analyzed by immunofluorescence microscopy, a protein
identified by this procedure, SidC, was shown to translocate across
the phagosomal membranes to the cytoplasmic face of the L.
pneumophila phagosome. The identification of large numbers of
these substrates, and the fact that the absence of any one substrate
rarely results in strong defects in intracellular growth, indicate that
there is significant functional redundancy among the DotIcm
translocation targets.
bacterial pathogenesis protein translocation plasmid conjugation
A
ctive modification of host cellular functions is essential for
a bacterial pathogen to establish a successful infection. Such
modification often is mediated by injecting effectors into the host
cytoplasm through specialized protein secretion systems (1).
Among these systems, conjugation-adaptive transporters, also
called type IV secretion systems (TFSS), have been identified in
a number of bacterial pathogens (2). Many of these TFSS are
dedicated DNA transfer apparatuses, whereas others allow
Gram-negative bacterial pathogens to translocate protein sub-
strates directly into the cytosol of eukaryotic cells. Only a few
protein substrates of TFSS that are translocated into host cells
have been identified (2).
Legionella pneumophila is an intracellular pathogen that
causes Legionnaire’s disease. After being phagocytosed by mac-
rophages, the bacteria multiply within a specialized vacuole that
is initially isolated from the endocytic pathway (3, 4), possibly by
intercepting early secretory vesicles (5). Biogenesis of this rep-
licative phagosome requires the TFSS transporter called Dot
Icm (6, 7). Substrates transported by this apparatus are believed
to directly promote the targeting pathway of the bacterial
vacuole (8). Two of these substrates, RalF and LidA, have been
identified (9, 10). RalF is a guanine nucleotide exchange factor
for multiple Arf proteins (9), whereas the biochemical activity of
LidA is unknown (10). There are clearly other unidentified
substrates of DotIcm, as mutations that specifically eliminate
LidA cause negligible defects in intracellular growth and muta-
tions in ralF are proficient for intracellular replication (9, 10). A
comprehensive analysis of the identity and function of effectors
translocated by the DotIcm apparatus is crucial in determining
how this bacterium establishes a replication vacuole. We report
here that proteins translocated from bacteria to host cells can
also be transferred between bacterial cells, allowing identifica-
tion of a large cohort of proteins transferred by the DotIcm
apparatus.
Materials and Methods
Bacterial Strains and Growth Conditions. All L. pneumophila strains
used in this study are derivatives of the wild-type strain Lp02 (thyA,
hsdR, and rpsL) (11). Lp03 is an isogenic dotA
mutant (12). All
strains were grown on casamino acids yeast extract thymidine
(CYET) plates or in N-(2-acetamido)-2-aminoethanesulfonic acid
yeast extract (AYE) broth (11). Bone marrow-derived macrophages
were prepared as described (11). To assay for intracellular growth
within macrophages from AJ mice (11) or within the amoebal host
Dictyostelium discoideum (13), L. pneumophila strains were grown
to postexponential phase, as judged by bacterial motility and cell
density (OD
600
3.3–3.7).
Plasmid Constructions. A derivative of pBRR1MCS mob
(Cm
r
)
(14), which was suitable for reporting interbacterial protein
translocation (called pZL184; Fig. 1), was constructed in a
cloning process involving multiple steps. Sequences of oligonu-
cleotides used, sources of gene cassettes, and details of cloning
are found in Supporting Methods, which is published as support-
ing information on the PNAS web site. To express the Cre
fusions, we first constructed pZL180 (Amp
r
and thyA
), which
contains cre on the transfer deficient RSF1010 plasmid pJB908
(J. Vogel, Washington University School of Medicine, St. Louis),
and fusions were constructed as follows: Whole ORFs were
amplified by PCR and were fused to the 3 end of cre on pZL180
for genes smaller than 2 kb, whereas for genes larger than 2 kb,
the 3 regions encoding 500 or 700 amino acids of each ORF was
fused to cre.
Bacterial Two-Hybrid Screening. First, ralF was translationally fused
to the 3 end of fragment T18 of the Bordetella pertussis cya gene,
encoding adenylate cyclase, on pUT18 (15). Sau3AI-generated
genomic DNA fragments from L. pneumophila strain Lp02 (11),
ranging from 800 base pairs to 5 kb, were inserted into pKT25
(15), resulting in a library of L. pneumophila proteins fused to the
C terminus of B. pertussis adenylate cyclase T25 fragment.
Escherichia coli strain BTH101 (15), expressing cya::ralF(C), was
used to identify RalF-interacting proteins by screening either on
LB medium containing 40
gml X-Gal or on the synthetic M63
medium (15) with lactose as the sole carbon source. Strains with
functional adenylate cyclase proteins were identified based on
the presence of detectable Lac
phenotypes on these media.
Proteins that interact with DotF were identified in a similar
manner by using either DotF or DotF (28–123) as the bait.
Bacterial Matings and Intercellular Protein Translocation. For RP4
Trb-mediated translocation of the Cre::MobA hybrid, 2.5
This paper was submitted directly (Track II) to the PNAS office.
Abbreviations: TFSS, type IV secretion systems; Sid, substrate of IcmDot transporter.
Data deposition: The sequence reported in this paper has been deposited in the GenBank
database (accession nos. AY504668–AY504685).
*To whom correspondence should be addressed. E-mail: ralph.isberg@tufts.edu.
© 2004 by The National Academy of Sciences of the USA
www.pnas.orgcgidoi10.1073pnas.0304916101 PNAS
January 20, 2004
vol. 101
no. 3
841–846
MICROBIOLOGY
10
8
cells from saturated cultures of S171 (Trb
) (16) or DH5
(Trb
) expressing the fusion grown in LB broth were mixed with
a 15-fold excess of recipient strain XL1Blue carrying pZL184.
The mixtures were spotted onto 0.45-
m nitrocellulose filters,
placed on LB plates, and incubated at 37°C for 3 h. The excisants
were selected on LB medium containing 5% sucrose and 30
gml kanamycin.
For L. pneumophila to L. pneumophila translocation, Lp02
(11) or Lp03dotA
(12) containing plasmids expressing the
appropriate Cre fusions were grown to postexponential phase in
AYE broth. For matings, 0.45-
m nitrocellulose membranes
were first placed onto CYET medium containing 100 nM
isopropyl
-D-thiogalactoside (IPTG) and the plates were incu-
bated at 37°C for 1 h before spotting with a mixture of 3.5
10
8
donor and a 15-fold excess of recipient Lp03(pZL184). The
mating plates were incubated at 37°C for 14 h and excisants were
selected on CYET 5% sucrose, 30
gml kanamycin, which
kills both parents. Each mating was performed in triplicate and
repeated at least three times. Transfer frequencies were ex-
pressed as number of excisants (resistant to kanamycin and
sucrose but sensitive to gentamicin) per donor bacterium and the
number of donor bacteria was determined by counting colony-
forming units derived from appropriately diluted donor cells
plated onto CYET medium before mating.
For plasmid transfer, transconjugants were detected by plating
the protein translocation mating mixture onto LB containing 100
gml ampicillin and 30
gml chroramphenicol (for E. coli)or
onto charcoal yeast extract containing 5
gml chloramphenicol
(selecting Thy
and Cm
R
) for transfer in L. pneumophila). As a
positive control, the wild-type RSF1010 derivative pKB5 (11)
was used in both E. coli and L. pneumophila matings (6).
Construction of In-Frame Deletions and Complementation of the
Deletion Mutants.
In-frame deletions of L. pneumophila genes
were performed by a two-step allelic exchange strategy as
described (17). In each case, the deletion construct was designed
such that the intact gene was replaced by an ORF predicted to
express a 20-aa polypeptide consisting of the first 10 amino acids
and the last 10 amino acids of the gene. To perform comple-
mentation studies on deletion mutants, the gene of interest was
amplified by PCR and inserted into pJB908 or pBBRMCS2 (14).
Protein Purification and Antibody Preparation. The predicted ORF
of SidC was inserted into pQE30 and the resulting His
6
-SidC
fusion protein was purified from E. coli by using Ni-
nitrilotriacetic acid resin (Qiagen, Valencia, CA). Rabbit poly-
clonal serum was generated by the Pocono Rabbit Farm (Cana-
densis, PA; ref. 10). Antibodies were affinity-purified by using a
matrix containing purified (His)
6
-SidC covalently coupled to
Affigel-10 beads (Bio-Rad) (18).
Western Blot and Immunofluorescence Staining. Total L. pneumo-
phila proteins separated by SDSPAGE were transferred onto
Immobilon-P membranes (Millipore) and probed by Western
blotting, as described (16). Filters were probed with affinity-
purified anti-SidC antibody (diluted 1:5,000), or anti-Bacillus
subtilis isocitrate dehydrogenase polyclonal antibody (diluted
1:5,000; a kind gift from Dr L. Sonenshein, Tufts University
School of Medicine). Immunofluorescence staining was per-
formed with affinity-purified rabbit anti-(His)
6
-SidC antibodies
followed by a Texas red-conjugated goat anti-rabbit antibody
(Molecular Probes). Fixation and probing techniques were per-
formed as described (10). Postnuclear L. pneumophila phago-
somes were isolated as described (10).
Fig. 1. Interbacterial protein translocation by the DotIcm system in the absence of DNA transfer. (A) Assay for interbacterial protein transfer. Translocation
of Cre hybrid protein from a donor bacterial strain is measured by removal of a oxed transcriptional terminator located between the trc promoter and the npt
II (kan
R
) gene on plasmid pZL184 harbored by a recipient bacterial strain. Bacteria harboring the intact reporter are unable to grow on media containing
kanamycin and sucrose. The translocation of Cre hybrid protein into the recipient strain leads to the excision, through recombination at the loxP sites, of the
DNA fragment that confers sucrose sensitivity (sacB) and reconstitution of a functional loxP-npt II translational fusion. S.D., ShineDalgarno sequence; npt II,
neomycin phosphotransferase; red diamond, transcriptional terminator. Arrows indicate the trc promoter and loxP sites. (B) Interbacterial transfer of a fusion
derived from the RSF1010 mobA gene. The mobA gene was fused to cre, and RP4-dependent protein translocation into a recipient E. coli strain was measured
by using E. coli S171 (15) as the donor selecting kanamycin resistance and screening for gentamicin sensitivity. Black bars, S171 Trb
donor; stippled bars, E.
coli DH5
Trb
donor. (C). Plasmids harboring Cre fusions cannot be transferred to recipient cells. Plasmids expressing the designated proteins were harbored
in either E. coli S171 (Trb
)orL. pneumophila Lp02 (DotIcm
; ref. 11), and the efciency of plasmid transfer was measured by using either recipient E. coli
or L. pneumophila strains, respectively. As a positive control, the identical plasmids having an intact oriT and mob system were used to demonstrate transfer
prociency of donor strains. Black bars, donor strain E. coli S171 (Trb
); gray bars, donor strain L. pneumophila Lp02 (DotIcm
; ref. 11). (D) Transfer of
translocated DotIcm substrates between bacterial cells. Protein transfer was performed as described in Materials and Methods, by using either Lp02 (DotIcm
)
or Lp03(dotA
) expressing the designated protein fusions as the donor strains. Gray bars, donor strain L. pneumophila Lp02 (DotIcm
); stippled bars, donor
strain L. pneumophila Lp03 (dotA
).
842
www.pnas.orgcgidoi10.1073pnas.0304916101 Luo and Isberg
Results
Interbacterial Protein Translocation by DotIcm. To identify sub-
strates of the DotIcm translocator, a screening strategy was
devised that allowed us to directly assay for protein transloca-
tion, using the CreloxP system from bacteriophage P1 (for
review, see ref.19). Previous work (20) demonstrated that the
presence of Cre does not interfere with the translocation of the
TFSS substrates VirE2 and VirF in Agrobacterium tumefaciens.
A reporter suitable for monitoring translocation of Cre fusion
proteins between two prokaryotic cells was constructed in which
expression of the npt II gene depended on the excision of a floxed
cassette consisting of a gentamicin resistance gene, the sacB
gene, and a transcriptional terminator (Fig. 1A). By using this
system, E. coli and L. pneumophila could transfer a hybrid
protein in which cre was fused to the mobA gene of plasmid
RSF1010 (Fig. 1B). When the Cre::MobA hybrid protein was
expressed from a plasmid lacking its origin of transfer (oriT)in
the E. coli strain S171 (16), transfer of the fusion protein into
another E. coli strain could be detected based on excision of the
stopper sequence and expression of kanamycin resistance (Fig.
1B, and Fig. 6, which is published as supporting information on
the PNAS web site). No transfer occurred when the Trb
strain
DH5
, which lacks conjugative transfer functions, was used as
the donor, or when a plasmid having cre alone was tested (Fig.
1B). Moreover, there was no transfer of the plasmid to the
recipient, indicating that the translocation of the fusion protein
occurred in the absence of DNA transfer (Fig. 1C; Cre::MobA
or Cre). L. pneumophila also was able to transfer the hybrid
protein in a DotIcm-dependent manner (Fig. 1D; Cre::MobA),
indicating that the DotIcm system is able to transfer proteins
interbacterially in the absence of DNA transfer (Fig. 1C).
Plasmid Mob proteins such as the MobA of RSF1010 are
believed to be transferred to recipient cells as proteinsingle-
stranded DNA complexes (21). However, our results demon-
strate that conjugation systems such as Trb and DotIcm are able
to transfer MobA (and presumably similar proteins) in the
absence of DNA transfer. This result is similar to what had been
observed with the ColIb-P9 SogL protein (22).
Two proteins known to be transferred from bacterial cells to
mammalian cells could also be translocated interbacterially by
using this strategy. Full-length ralF and the 3 half of lidA [called
lidA(C)] were each translationally fused to cre and the resulting
hybrids were examined for bacterium-bacterium translocation.
We found that wild type L. pneumophila, but not a DotIcm-
deficient strain, could transfer Cre::RalF or Cre::LidA(C), based
on the ability of recipient strain to excise the floxed DNA
fragment and to express kanamycin resistance [Fig. 1D;
Cre::RalF and Cre::LidA(C)]. However, Cre itself could not be
translocated by wild-type L. pneumophila (Fig. 1D). The trans-
location proficiency of Cre::LidA(C) indicates that targeting
information resides in the C terminus of the protein, at least for
interbacterial transfer. We examined whether this finding was
also true for RalF by testing DotIcm-mediated translocation of
Cre::RalF(C) and Cre::RalF(N) respectively. Only Cre:RalF(C)
was translocated at a detectable frequency (Fig. 1D). These
observations are similar to findings regarding A. tumefaciens Vir
proteins, in which secretion signals are localized to the C termini
of the substrates (20, 23).
Identification of Substrates of the DotIcm Transporter. Because the
translocated proteins LidA and RalF could be transferred
interbacterially, we used the bacterial translocation assay to
identify proteins transferred by this transporter. To simplify the
screening of a randomly generated fusion library, we chose to
screen through a preselected pool of candidate substrates. We
hypothesized that some translocated substrates may interact with
specific components of the DotIcm complex. Therefore, a
bacterial two-hybrid screen (15) was used to identify L. pneu-
mophila proteins that specifically interacted with the carboxyl
portion of RalF. From 41 positive clones sequenced, DotF,
predicted to be an inner membrane component of the DotIcm
complex, was identified nine times independently. Interestingly,
in all cases, only the portion of DotF that spans amino acids
28123 was obtained. The full-length DotF also gave a positive
two-hybrid readout, although at a somewhat lower level (Fig. 2).
A bacterial two-hybrid study was then performed by using Cya
fusions to the 3 ends of L. pneumophila genes and baits of DotF
or its derivative, DotF (28123). From 150,000 candidates, we
isolated 148 clones that showed some level of interaction higher
than background, and these were sequenced and analyzed.
Although it was clear that some strains gave barely detectable
-galactosidase activity and were of questionable significance,
this strategy greatly reduced the number of strains to be ana-
lyzed. Complete ORFs of these genes were retrieved from the
L. pneumophila genome database (http:兾兾genome3.cpmc.
columbia.edulegion) and 68 genes were identified. Of these
genes, 20 candidates were eliminated based primarily on their
high hydrophobicity, which may have caused nonspecific inter-
actions with DotF. Forty-eight candidate genes were subse-
quently tested for interbacterial transfer (see Table 2, which is
published as supporting information on the PNAS web site, for
data on 17 such candidates from a single experiment). Table 1
shows eight clones that gave a positive signal by using the
CreloxP reporter system. We designated these proteins sub-
strate of IcmDot transporter (Sid) (Table 1). The efficiency of
interbacterial transfer varied, ranging from 10
6
to 10
5
ex-
cisants per input donor cell. Moreover, transfer of all of these
proteins depended on the DotIcm apparatus, because mating
with dotA
strains expressing these fusions failed to cause the
excision of the floxed DNA fragment and the subsequent
expression of the kanamycin resistance marker (Table 1).
Characteristics of
sid
Genes. Sequence analyses revealed that with
the exception of sidB, the sid genes have no significant orthologs
present in the GenBank NR database (Table 1). SidB contains
Fig. 2. Interactions between DotF and the C termini of RalF and SidC,
respectively. E. coli strain BTH101 (14) containing the indicated plasmids was
grown overnight at 28°C in LB and the cultures were diluted 20-fold in the
same medium containing 100 nM IPTG. Cultures were grown for 14 h before
an appropriate volume was withdrawn for
-galactosidase assay. Tested
strains are as follows: (column A) BTH101(pKT25dotF, pUT18C); (column B)
BTH101(pKT25, pUT18CRalF(C); (column C) BTH101(pKT25dotF,
pUT18CRalF(C); (column D) BTH101(pDotF (28123), pKT25RalF(C); (column E)
BTH101 [pDotF (28123), pSidC300(C)]k; and (column F) the leucine zipper
domain of the yeast GCN4 protein (14) was used as positive control in
BTH101(pKT25Zip, pUT18CZip).
-galactosidase activity was expressed as
Miller units. Experiments were performed in triplicate for three independent
times. Data shown are from one representative experiment.
Luo and Isberg PNAS
January 20, 2004
vol. 101
no. 3
843
MICROBIOLOGY
a putative active site found in some lipases and shows a region
similar to a portion of the Rtx toxin from Vibrio cholerae (24).
In contrast, many of these proteins have one or more paralogs
present in the L. pneumophila genome (Table 1). The similarities
among the paralogs range from almost identical predicted
proteins (E 0) to rather loose similarity (E 5e
2
to 2e
8
;
Table 1). Interestingly, in some cases, subsets of paralogs are
organized into contiguous ORFs. For instance, sidC and its
homolog sdcA are separated by only 150 base pairs, whereas
sdeA, sdeB, and sdeC are closely clustered (Fig. 3). In addition,
a paralog of sidE (sdeD) is located directly upstream from what
appears to be the operon encoding sidC (Fig. 3B). Interestingly,
two putative genes, orf1 and orf3, located near the sidA and sdeC
regions are highly similar (E 0; Fig. 3 A and B), although these
genes were not identified in the original two-hybrid assay.
SidC Is Translocated by DotIcm Into the Cytosol of Host Cells. To
verify that proteins identified in our assay are translocated into
mammalian cells, affinity-purified antiserum against His
6
-SidC
was used to analyze expression and translocation of SidC by L.
pneumophila. As predicted, a protein of 110 kDa was detected
in both a doticm
strain and a dotA
mutant of L. pneumophila,
but not in a mutant missing sidC and its paralog sdcA [(sdcA-
sidC)] (Fig. 3C). Moreover, expression of SidC was induced
7-fold in cells grown to post exponential phase (Fig. 3C). The
translocated substrates RalF and LidA also show induced ex-
pression in post exponential phase (ref. 9 and G. M. Conover and
R.R.I., unpublished observations), and a large body of evidence
indicates that factors critical for intracellular survival of L.
pneumophila are similarly regulated (25, 26). When infected
macrophages were probed for SidC by indirect immunofluores-
cence microscopy, we found that1hafteruptake, 86% of L.
pneumophila phagosomes stained positively for SidC (Fig. 4A),
with the protein localizing about the phagosomal membrane
(Fig. 4 C and D). In some cases, there was asymmetric diffusion
of the protein emanating from the phagosome surface, with most
infected cells showing protein only in the region near the
phagosome (Fig. 4D).
Because the above strategy only detects secretion of SidC and
does not demonstrate translocation across the phagosomal mem-
brane, a second approach was pursued (10). To demonstrate
SidC translocation across the phagosomal membrane by the
DotIcm transporter, we prepared postnuclear supernatants
(PNS) from macrophages incubated with L. pneumophila and
probed intact phagosomes for SidC in the absence of perme-
abilization reagents. Approximately 85% of these phagosomes
stained positively for SidC (Fig. 4 B and E), whereas 5%
stained positively with anti-L. pneumophila serum, demonstrat-
ing that the phagosomal membranes were intact. Less than 0.3%
of macrophages containing a L. pneumophila dotA
strain
stained positively for SidC (Fig. 4 A and C), which was consistent
with our data showing that interbacterial translocation of this
protein requires the DotIcm apparatus. Similarly, in PNS, no
isolated phagosome containing a dotA
strain stained positively
for SidC (Fig. 4B). The antibody reactivity was specific, because
no macrophages harboring a mutant lacking sidC and its up-
Table 1. L. pneumophila proteins identied in this study that are translocated between
bacterial cells by the DotIcm system
Gene
name
Accession
no.
Protein size,
aa
L. pneumophila
paralogs
E values of
paralogs
Orthologs in other
species (E value)
Translocation
frequency*
sidA AY504668 474 NA 7.2 1.2 10
6
sidB AY504669 417 4 7e
26
to 5e
09
Rtx toxinlipase (e
5
) 6.7 0.2 10
6
sidC AY504673 918 2 0
3.3 0.7 10
6
sidD AY504675 472 NA 2.3 0.2 10
5
sidE AY504676 1,495 5 02e
05
1.2 0.5 10
6
sidF AY504681 912 NA 6.2 0.2 10
5
sidG AY504682 965 NA 8.2 0.3 10
6
sidH AY504683 2,225 3 7e
08
to 3e
04
3.2 0.4 10
5
SdeC
AY504679 1,538 5 0 to 2e
05
2.2 0.9 10
6
Translocation assay was performed as described in Materials and Methods. NA, not applicable.
*Translocation frequency was expressed as numbers of excisants (resistant to both kanamycin and sucrose and
sensitive to gentamicin) per donor cell. No translocation occurred when the dotA
strain Lp03 was used as the
donor. The number of donor cells was determined by plating the appropriate dilutions of the donor culture onto
solid CYE medium. The rates of spontaneous mutants resistant to kanarnycin and sucrose of the reporter strain
appeared on the selective medium used to select excisants were approximately 2.5 10
9
per recipient.
An E value of 0 indicates that the paralogs are almost identical proteins.
Paralogs are named for the gene identied in DotF two-hybrid assays. For instance, paralog A of sidE is called
sdeA.
Fig. 3. Clustering of sid genes, their paralogs into operon-like structures, and
growth phase regulation of sidC.(A) sidC and its homolog sdcA are separated
by only 150 bp, and the two genes are closely linked to a paralog of sidE(sdeD).
(B) Three paralogs of sidE are part of a contiguous region of the chromosome
that contains ve signicant ORFs. (C) Growth phase regulation of sidC.
Bacteria were grown in AYE broth to either an OD
600
1.8 (exponential) or 3.7
(postexponential), harvested, and analyzed by SDSPAGE and immunoblot-
ting with afnity-puried anti-(His)
6
-SidC. Displayed are Lp02(doticm intact;
ref. 11), Lp03 (dotA
; ref. 12), and Lp02(sdcA, sidC), an Lp02 derivative
deleted for sidC and its upstream paralog. The isocitrate dehydrogenase
(ICDH) protein was used as a loading control by probing with antiserum raised
against B. subtilis ICDH.
844
www.pnas.orgcgidoi10.1073pnas.0304916101 Luo and Isberg
stream paralog stained positively for this protein [Fig. 4A;
(sdcA-sidC)], nor did macrophages infected with a simple
deletion of sidC (data not shown). These results demonstrate
that during infection, SidC is translocated into mammalian cells
by the DotIcm system, and that translocation occurs across the
phagosomal membrane. Because the transfer frequencies of
other proteins in our CreloxP system were comparable to that
of SidC, we postulate that these proteins are also targeted to the
host cell by the DotIcm transporter.
SdeC Is Required for Efficient Intracellular Growth by
L. pneumophila
.
To examine the importance of the Sid proteins in L. pneumophila
pathogenesis, we constructed in-frame deletions in some of these
genes and tested the resulting mutants for intracellular growth in
bone marrow-derived macrophages (10). Individual deletions of
sidA, sidD, sidF,orsidG, which have no detectable paralogs in
the available L. pneumophila database, resulted in strains with
intracellular growth properties that were difficult to distinguish
from the wild-type strain. Furthermore, a mutant lacking sidC
and its upstream paralog sdcA grew proficiently (data not
shown). Finally, a quadruple mutant lacking sidB and its three
paralogs has a defect in intracellular growth, but this result was
not significantly different from a mutant missing sdbA alone
(data not shown), suggesting that functional redundancy extends
beyond specific substrate families. We then constructed in-frame
deletion mutants lacking individual paralogs of sidE and exam-
ined intracellular growth of these mutants in bone marrow-
derived or in D. discoideum. Unexpectedly, only deletion of the
single paralog sdeC had a detectable growth defect (Fig. 5). In
D. discoideum, by using L. pneumophila strains harboring the
plasmid vector, the yield of viable counts after 48 h after
incubation was depressed 3-fold for the mutant lacking sdeC
(Fig. 5). Defective growth could be complemented in trans by a
single ORF containing sdeC (Fig. 5). A similar growth defect of
this mutant also was observed in macrophage (data not shown).
In contrast, mutants missing sidE, sdeA,orsdeB, respectively,
had no detectable growth defect in bone marrow-derived mac-
rophages (data not shown). As expected, a Cre::SdeC fusion was
found to be translocated between bacterial cells (Table 1).
Discussion
By developing a genetic assay for monitoring protein transloca-
tion using the creloxP system, we conclusively demonstrated
that the DotIcm TFSS transporter can perform interbacterial
transfer of proteins known to be targeted to mammalian cells, an
observation that we have exploited to identify a large number of
translocated proteins. The majority of the proteins we identified
have no significant orthologs in the database, suggesting that
biogenesis of the L. pneumophila replication compartment may
involve mechanisms that differ from other organisms, such as
Chlamydia trachomatis and Mycobacterium spp., which reside in
similar intracellular vacuoles. Alternatively, proteins with similar
functions in different microbial species may have evolved inde-
pendent of each other, and show little sequence similarity to each
Fig. 4. SidC is translocated by the L. pneumophila DotIcm system to the host
cell and is localized about the phagosomal membrane. (A) Bone marrow-
derived macrophages from AJ mice were infected with Lp02(doticm intact),
Lp03(dotA
), or Lp02(sdcA-sidC) strain expressing GFP, respectively. One
hour after infection, cells were xed as described (10), and SidC was probed
with anti-(His)
6
-SidC antibodies and Texas red-labeled secondary antibodies.
Stained macrophages were scored for translocation of SidC by counting
phagosomes that stained positively with anti-(His)
6
-SidC. Data shown are from
two independent experiments performed in triplicate in which at least 100
phagosomes were scored per coverslip. (B) SidC staining on PNS prepared from
L. pneumophila-infected U937 cells in the absence of permeabilization. Sam-
ple preparation, immunostaining, and data collection were performed as
described in ref. 10 or in A.(C) DotA-dependent translocation of SidC. (Left)
Bacteria expressing GFP associated with bone marrow-derived macrophage.
Strains used were Lp02(doticm intact; Upper) and Lp03 (dotA
; Lower).
(Center) Immunoprobing of infected cells with anti-(His)
6
-SidC. (Right)
Merged images of GFP and anti-(His)
6
-SidC staining. (D) Limited diffusion of
SidC from L. pneumophila phagosome. Shown are images of Lp02(doticm
intact) with murine bone borrow-derived macrophage. (E) SidC is translocated
across the phagosomal membrane. Shown are images of PNSs of Lp02(doticm
intact)-infected macrophages. Bacteria and SidC are probed as above, with
bacteria marked by GFP and anti-SidC marked in red.
Fig. 5. sdeC is required for efcient intracellular growth. D. discoideum cells
were infected with a multiplicity of infection of 0.05, and growth of bacteria
was monitored as described (13). Total bacterial cells were washed from
individual microtiter wells at designated times, and the appropriate dilutions
were plated on charcoal yeast extract plates to obtain the colony-forming
units. Fold of growth was obtained by dividing colony-forming units at a given
time point by the input bacterial cell numbers. Strains tested are as follows:
black bars, Lp02(intact doticm); gray bars, Lp02(sdeC); striped bars,
Lp02(sdeC) harboring pZL192 that carries the ORF of sdeC. Data shown are
from two independent experiments performed in triplicate.
Luo and Isberg PNAS
January 20, 2004
vol. 101
no. 3
845
MICROBIOLOGY
other. Many of the translocated proteins have paralogs within L.
pneumophila, often encoded in chromosomal regions devoted to
expression of translocated substrates of DotIcm (Fig. 3 A and
B). These genes may have been organized to ensure proper
expression during exposure to environmental conditions that are
optimal for initiating intracellular growth.
The lack of intracellular growth defects observed in L. pneu-
mophila mutants lacking translocated substrates is reminiscent
of a L. pneumophila ralF mutant, which is indistinguishable from
the parental wild-type strain in regards to intracellular growth
(9). Nevertheless, a ralF mutant is unable to recruit Arf1 to the
surface of the replicative phagosome, a property that may play
some unknown role in the lifestyle of the organism (27). It is
possible that elimination of some of the genes here identified
similarly results in the formation of phagosomes with alterations
in either morphology or host protein content that have little
consequence with respect to growth in cultured cells. Alterna-
tively, the roles played by these proteins may be substituted by
other substrates of DotIcm, or some of these proteins may be
important for growth in some untested host cell.
It is common that a bacterial pathogen codes for numerous
effectors but the presence of multiple paralogs of a specific
translocated effector in the same organism is only occasionally
found (2830). The close similarity among proteins in a family
points to functional redundancy, which may provide an expla-
nation for the failure to identify these proteins in previous
genetic screens for bacterial mutants defective in intracellular
growth. In the case of the sidB family, however, functional
redundancy clearly extends beyond the identified paralogs, and
proteins of little similarity in sequence or function may be
redundant. It is plausible that each paralog may be adapted to
promote growth only in specific host cells. Because L. pneumo-
phila is a versatile pathogen that interacts with very diverse hosts
in the environment, the establishment of a successful intracel-
lular replication niche may require only a subset of translocated
substrates, with a single subset adapted for a particular host cell
type.
Sorting out the minimal complement of translocated proteins
necessary for intracellular replication will be a challenge for the
future, because the number of such proteins may be quite large,
based on the data displayed in Table 1. In addition to the
products of the genes identified in the original CreloxP assay
and sdeC, we have tested four paralogs of Sid proteins for
interbacterial transfer, and all can promote transfer of Cre (data
not shown). Based on this finding, we speculate that it is likely
that all of the paralogs are capable of translocation. Further-
more, in another study we found an additional family of four
proteins that can be transferred (S. M. VanRheenen, Z.-Q.L.,
and R.R.I., unpublished results). Altogether, this brings the
minimum number of translocated proteins encoded by L. pneu-
mophila to 24. In fact, there are probably many more proteins
translocated by DotIcm, as the DotF interaction screen was not
carried to saturation. The gene bank required in-frame fusions
to a Sau3AI restrictions site, which may not be present in all
genes encoding translocated substrates. Furthermore, the screen
described here focused on only the subset of DotF interactors,
and it appears that not all translocated substrates bind DotF at
detectable levels in this assay.
The other salient feature of the DotIcm translocation system
that was uncovered here is the large size of the proteins that were
translocated. Many of the proteins identified in this study are
90 kDa, with one predicted to be 240 kDa. It appears likely
that TFSS have evolved to transport large molecules, such as
DNAprotein complexes, and these transfer systems may be the
preferred translocators for high molecule weight substrates.
In summary, the L. pneumophila TFSS is a transporter system
of striking flexibility. It is capable of promoting bacterial con-
jugation and translocation of proteins from bacteria into mam-
malian cells, as well as transporting these same proteins between
bacterial strains. Although the efficiency of interbacterial pro-
tein transfer appears to be many orders of magnitude lower than
the transfer from bacteria to macrophages (Table 1 and Figs. 1D
and 4 A and B), this property has allowed us to identify substrates
of the DotIcm system. Understanding the functions of these
proteins should allow elucidation of the mechanisms underlying
the biogenesis of the L. pneumophila-replicative phagosome.
Finally, the methods we described here may be generalized to
identify mammalian effectors transferred by all TFSS.
We thank M. Tang for assistance in protein purification; S. Farrand, D.
Ladant, A. Vergunst, and J. Vogel for supplying plasmids; the Isberg
laboratory for helpful discussions; and Drs. Carol Kumamoto, Michael
Malamy, Susan VanRheeven, Marion Shonn, Isabelle Derre, and Mat-
thias Machner for review of the text. This work was supported by the
Howard Hughes Medical Institute. Z.-Q.L. is a Howard Hughes Medical
Institute Fellow of the Life Sciences Research Foundation, and R.R.I. is
a Howard Hughes Medical Institute Investigator.
1. Staskawicz, B. J., Mudgett, M. B., Dangl, J. L. & Galan, J. E. (2001) Science
292, 22852289.
2. Christie, P. J. (2001) Mol. Microbiol. 40, 294305.
3. Horwitz, M. A. (1983) J. Exp. Med. 158, 21082126.
4. Sturgill-Koszycki, S. & Swanson, M. S. (2000) J. Exp. Med. 192, 12611272.
5. Kagan, J. C. & Roy, C. R. (2002) Nat. Cell Biol. 4, 945954.
6. Vogel, J. P., Andrews, H. L., Wong, S. K. & Isberg, R. R. (1998) Science 279,
873876.
7. Segal, G., Purcell, M. & Shuman, H. A. (1998) Proc. Natl. Acad. Sci. USA 95,
16691674.
8. Vogel, J. P. & Isberg, R. R. (1999) Curr. Opin. Microbiol. 2, 3034.
9. Nagai, H., Kagan, J. C., Zhu, X., Kahn, R. A. & Roy, C. R. (2002) Science 295,
679682.
10. Conover, G. M., Derre, I. I., Vogel, J. P. & Isberg, R. R. (2003) Mol. Microbiol.
48, 305321.
11. Berger, K. H. & Isberg, R. R. (1993) Mol. Microbiol. 7, 719.
12. Berger, K. H., Merriam, J. J. & Isberg, R. R. (1994) Mol. Microbiol. 14,
809822.
13. Solomon, J. M., Rupper, A., Cardelli, J. A. & Isberg, R. R. (2000) Infect.
Immun. 68, 29392947.
14. Kovach, M. E., Elzer, P. H., Hill, D. S., Robertson, G. T., Farris, M. A., Roop,
R. M. & Peterson, K. M. (1995) Gene 166, 175176.
15. Karimova, G., Pidoux, J., Ullmann, A. & Ladant, D. (1998) Proc. Natl. Acad.
Sci. USA 95, 57525756.
16. Simon, R., Priefer, U. & Puhler, A. (1983) BioTechnology 1, 3745.
17. Dumenil, G. & Isberg, R. R. (2001) Mol. Microbiol. 40, 11131127.
18. Harlow, E. & Lane, D. (1999) in Using Antibodies, A Laboratory Manual (Cold
Spring Harbor Lab. Press, Plainview, NY), pp. 311343.
19. Van Duyne, G. D. (2001) Annu. Rev. Biophys. Biomol. Struct. 30, 87104.
20. Vergunst, A. C., Schrammeijer, B., den Dulk-Ras, A., de Vlaam, C. M.,
Regensburg-Tuink, T. J. & Hooykaas, P. J. (2000) Science 290, 979982.
21. Christie, P. J. (1997) J. Bacteriol. 179, 30853094.
22. Wilkins, B. M. & Thomas, A. T. (2000) Mol. Microbiol. 38, 650657.
23. Simone, M., McCullen, C. A., Stahl, L. E. & Binns, A. N. (2001) Mol. Microbiol.
41, 12831293.
24. Lin, W., Fullner, K. J., Clayton, R., Sexton, J. A., Rogers, M. B., Calia, K. E.,
Calderwood, S. B., Fraser, C. & Mekalanos, J. J. (1999) Proc. Natl. Acad. Sci.
USA 96, 10711076.
25. Byrne, B. & Swanson, M. S. (1998) Infect. Immun. 66, 30293034.
26. Hammer, B. K. & Swanson, M. S. (1999) Mol. Microbiol. 33, 721731.
27. Roy, C. R. & Tilney, L. G. (2002) J. Cell Biol. 158, 415419.
28. Hartman, A. B., Venkatesan, M., Oaks, E. V. & Buysse, J. M. (1990) J.
Bacteriol. 172, 19051915.
29. Stender, S., Friebel, A., Linder, S., Rohde, M., Mirold, S. & Hardt, W. D. (2000)
Mol. Microbiol. 36, 12061221.
30. Miao, E. A., Scherer, C. A., Tsolis, R. M., Kingsley, R. A., Adams, L. G.,
Baumler, A. J. & Miller, S. I. (1999) Mol. Microbiol. 34, 850864.
846
www.pnas.orgcgidoi10.1073pnas.0304916101 Luo and Isberg

Supplementary resources (18)

... Our analysis indicated that nine L. pneumophila effector models contained the ɑ/β hydrolase domain (ECOD T group "alpha/beta-hydrolases"), and three effector models contained domains reminiscent of SGNH hydrolases (ECOD T group "SGNH hydrolase") ( Table 6). Only three of these effectors -Lpg1642/SidB, Lpg1907, and Lpg2911 -have been previously reported to possess an ɑ/β hydrolase domain (Gomez-Valero et al., 2011;Gomez-Valero et al, 2014;Luo & Isberg, 2004), The structure of Lpg2422 has been experimentally determined (PDB 4m0m), confirming the presence of this domain; however, to the best of our knowledge, none of these effector proteins have been experimentally characterized to possess hydrolase activity. Along the same lines, we did not find any previous reports describing Legionella effectors possessing the SGNH hydrolase domain. ...
Preprint
Full-text available
Legionella pneumophila utilizes the Dot/Icm type IVB secretion system to deliver hundreds of effector proteins inside eukaryotic cells to ensure intracellular replication. Our understanding of the molecular functions of the largest pathogenic arsenal known to the bacterial world remains incomplete. By leveraging advancements in 3D protein structure prediction, we provide a comprehensive structural analysis of 368 L. pneumophila effectors, representing a global atlas of predicted functional domains summarized in a database (https://pathogens3d.org/legionella-pneumophila). Our analysis identified 157 types of diverse functional domains in 287 effectors, including 159 effectors with no prior functional annotations. Furthermore, we identified 35 unique domains in 30 effector models that have no similarity with experimentally structurally characterized proteins, thus, hinting at novel functionalities. Using this analysis, we demonstrate the activity of thirteen domains, including three unique folds, predicted in L. pneumophila effectors to cause growth defects in the Saccharomyces cerevisiae model system. This illustrates an emerging strategy of exploring synergies between predictions and targeted experimental approaches in elucidating novel effector activities involved in infection.
... The Gram-negative pathogen Legionella pneumophila is the causative agent for Legionnaires' disease, a severe type of pneumonia resulting from infection of alveolar macrophages. Upon entry into its host cells, the bacterium utilizes its Dot/Icm type-IV secretion system (T4SS) to translocate at least 330 effectors into the host cytosol for its survival and replication [1][2][3] . Among the variety of strategies employed by these effectors, ADP-ribosylation, in which ADP-ribose (ADPR) from nicotinamide adenine dinucleotide (NAD + ) is added onto target proteins, has recently emerged as a means for host subjugation. ...
Article
Full-text available
ADP-ribosylation is a reversible post-translational modification involved in various cellular activities. Removal of ADP-ribosylation requires (ADP-ribosyl)hydrolases, with macrodomain enzymes being a major family in this category. The pathogen Legionella pneumophila mediates atypical ubiquitination of host targets using the SidE effector family in a process that involves ubiquitin ADP-ribosylation on arginine 42 as an obligatory step. Here, we show that the Legionella macrodomain effector MavL regulates this pathway by reversing the arginine ADP-ribosylation, likely to minimize potential detrimental effects caused by the modified ubiquitin. We determine the crystal structure of ADP-ribose-bound MavL, providing structural insights into recognition of the ADP-ribosyl group and catalytic mechanism of its removal. Further analyses reveal DUF4804 as a class of MavL-like macrodomain enzymes whose representative members show unique selectivity for mono-ADP-ribosylated arginine residue in synthetic substrates. We find such enzymes are also present in eukaryotes, as exemplified by two previously uncharacterized (ADP-ribosyl)hydrolases in Drosophila melanogaster. Crystal structures of several proteins in this class provide insights into arginine specificity and a shared mode of ADP-ribose interaction distinct from previously characterized macrodomains. Collectively, our study reveals a new regulatory layer of SidE-catalyzed ubiquitination and expands the current understanding of macrodomain enzymes.
... In contrast, one LP effector protein called SdhA was shown to be essential for the replication of LP inside mouse macrophages 8 . SdhA belongs to the SdhA family of effectors sharing~40% sequence similarity with its paralogs SdhB and SidH in the N-terminal region (~700 amino acids) of the protein sequence 9 . Interestingly, deletion of SdhA alone results in defective growth in mouse macrophages while deletion of the two paralogues SidH and SdhB does not cause any significant growth phenotype indicating that these paralogs may have different functions during infection. ...
Article
Full-text available
Legionella pneumophila (LP) secretes more than 300 effectors into the host cytosol to facilitate intracellular replication. One of these effectors, SidH, 253 kDa in size with no sequence similarity to proteins of known function is toxic when overexpressed in host cells. SidH is regulated by the LP metaeffector LubX which targets SidH for degradation in a temporal manner during LP infection. The mechanism underlying the toxicity of SidH and its role in LP infection are unknown. Here, we determined the cryo-EM structure of SidH at 2.7 Å revealing a unique alpha helical arrangement with no overall similarity to known protein structures. Surprisingly, purified SidH came bound to a E. coli EF-Tu/t-RNA/GTP ternary complex which could be modeled into the cryo-EM density. Mutation of residues disrupting the SidH-tRNA interface and SidH-EF-Tu interface abolish the toxicity of overexpressed SidH in human cells, a phenotype confirmed in infection of Acanthamoeba castellani. We also present the cryo-EM structure of SidH in complex with a U-box domain containing ubiquitin ligase LubX delineating the mechanism of regulation of SidH. Our data provide the basis for the toxicity of SidH and into its regulation by the metaeffector LubX.
... This transition is linked to large-scale transcriptomic alterations mediated by multiple bacterial two-component systems in response to various environmental signals (43). In particular, many effector-encoding genes were found to be expressed in the post-exponential phase in bacteriological media, a profile equivalent to that of the transmissive form of intracellular L. pneumophila (44). To obtain insights into the virulence traits of Lug15, we investigated its expres sion profile in bacteria grown in broth with Lug15-specific antibodies. ...
Article
Full-text available
Legionella pneumophila is a facultative intracellular pathogen that causes legionellosis. The key to its virulence is the delivery of hundreds of effector proteins into host cells via the d efective in o rganelle t rafficking/ i ntra c ellular m ultiplication type IV secretion system. These effectors modulate numerous host signaling pathways to create a niche called the L egionella - c ontaining v acuole (LCV) permissive for its intracellular replication. Previous investigation revealed that exploitation of the host ubiquitin system is among the most important strategies used by L. pneumophila to coopt host processes for its benefit. Here, we show that the effector Legionella ubiquitin ligase gene 15 (Lug15) (Lpg2327), which has no detectable homology with any enzyme involved in ubiquitin signaling, is an E3 ligase. In L. pneumophila -infected cells, Lug15 is localized on the LCV and impacts its association with polyubiquitinated proteins. We also demonstrate that Sec22b is ubiquitinated and recruited to the LCV by Lug15. Thus, our results establish Lug15 as a novel E3 ligase that functions to recruit a SNARE protein to remodel the L. pneumophila phagosome. IMPORTANCE Protein ubiquitination is one of the most important post-translational modifications that plays critical roles in the regulation of a wide range of eukaryotic signaling pathways. Many successful intracellular bacterial pathogens can hijack host ubiquitination machinery through the action of effector proteins that are injected into host cells by secretion systems. Legionella pneumophila is the etiological agent of legionellosis that is able to survive and replicate in various host cells. The d efective in o rganelle t rafficking (Dot)/ i ntra c ellular m ultiplication (Icm) type IV secretion system of L. pneumophila injects over 330 effectors into infected cells to create an optimal environment permissive for its intracellular proliferation. To date, at least 26 Dot/Icm substrates have been shown to manipulate ubiquitin signaling via diverse mechanisms. Among these, 14 are E3 ligases that either cooperate with host E1 and E2 enzymes or adopt E1/E2-independent catalytic mechanisms. In the present study, we demonstrate that the L. pneumophila effector Legionella ubiquitin ligase gene 15 (Lug15) is a novel ubiquitin E3 ligase. Lug15 is involved in the remodeling of LCV with polyubiquitinated species. Moreover, Lug15 catalyzes the ubiquitination of host SNARE protein Sec22b and mediates its recruitment to the LCV. Ubiquitination of Sec22b by Lug15 promotes its noncanonical pairing with plasma membrane-derived syntaxins (e.g., Stx3). Our study further reveals the complexity of strategies utilized by L. pneumophila to interfere with host functions by hijacking host ubiquitin signaling.
Preprint
Full-text available
Legionella pneumophila grows within membrane-bound vacuoles in phylogenetically diverse hosts. Intracellular growth requires the function of the Icm/Dot type-IVb secretion system, which translocates more than 300 proteins into host cells. A screen was performed to identify L. pneumophila proteins that stimulate MAPK activation, using Icm/Dot translocated proteins ectopically expressed in mammalian cells. In parallel, a second screen was performed to identify L. pneumophila proteins expressed in yeast that cause growth inhibition in MAPK pathway-stimulatory high osmolarity medium. LegA7 was shared in both screens, a protein predicted to be a member of the bacterial cysteine protease family that has five carboxyl-terminal ankyrin repeats. Three conserved residues in the predicted catalytic triad of LegA7 were mutated. These mutations abolished the ability of LegA7 to inhibit yeast growth. To identify other residues important for LegA7 function, a generalizable selection strategy in yeast was devised to isolate mutants that have lost function and no longer cause growth inhibition on high osmolarity medium. Mutations were isolated in the two amino-terminal ankyrin repeats, as well as an inter-domain region located between the cysteine protease domain and the ankyrin repeats. These mutations were predicted by AlphaFold modeling to localize to the face opposite from the catalytic site, arguing that they interfere with the positive regulation of the catalytic activity. Based on our data, we present a model in which LegA7 harbors a cysteine protease domain with an inter-domain and two amino-terminal ankyrin repeat regions that modulate the function of the catalytic domain.
Preprint
Full-text available
To promote intracellular survival and infection, Legionella spp. translocate hundreds of effector proteins into eukaryotic host cells using a type IV b protein secretion system (T4bSS). T4bSS are well known to translocate soluble as well as transmembrane domain-containing effector proteins (TMD-effectors) but the mechanisms of secretion are still poorly understood. Herein we investigated the secretion of hydrophobic TMD-effectors, of which about 80 were previously reported to be encoded by L. pneumophila. A proteomic analysis of fractionated membranes revealed that TMD-effectors are targeted to and inserted into the bacterial inner membranes of L. pneumophila independent of the presence of a functional T4bSS. While the T4bSS chaperones IcmS and IcmW were critical for secretion of all tested TMD-effectors, they did not influence inner membrane targeting of these proteins. As for soluble effector proteins, translocation of TMD-effectors into host cells depended on a C-terminal secretion signal and this signal needed to be presented towards the cytoplasmic side of the inner membrane. A different secretion behavior of TMD- and soluble effectors and the need for small periplasmic loops within TMD-effectors provided strong evidence that TMD-effectors are secreted in a two-step secretion process: Initially, an inner membrane intermediate is formed, that is extracted towards the cytoplasmic side, possibly by the help of the type IV coupling protein complex and subsequently secreted into eukaryotic host cells by the T4bSS core complex. Overall, our study highlights the amazing versatility of T4bSS to secrete soluble and TMD-effectors from different subcellular locations of the bacterial cell.
Article
Legionella are freshwater Gram‐negative bacteria that in their normal environment infect protozoa. However, this adaptation also allows Legionella to infect human alveolar macrophages and cause pneumonia. Central to Legionella pathogenesis are more than 330 secreted effectors, of which there are 9 core effectors that are conserved in all pathogenic species. Despite their importance, the biochemical function of several core effectors remains unclear. To address this, we have taken a structural approach to characterize the core effector of unknown function LceB, or Lpg1356, from Legionella pneumophila. Here we solve an X‐ray crystal structure of LceB using an AlphaFold model for molecular replacement. The experimental structure shows that LceB adopts a Sel1‐like repeat fold as predicted. However, the crystal structure captured multiple conformations of LceB all of which differed from the AlphaFold model. Comparison of the predicted model and the experimental models suggests that LceB is highly flexible in solution. Additionally, molecular analysis of LceB using its close structural homologues reveals sequence and structural motifs of known biochemical function. Specifically, LceB harbors a repeated KAAEQG motif that both stabilizes the Sel1‐like repeat fold and is known to participate in protein‐protein interactions with eukaryotic host proteins. We also observe that LceB forms several higher‐order oligomers in solution. Overall, our results have revealed that LceB has conformational flexibility, self‐associates, and contains a molecular surface for binding a target host‐cell protein. Additionally, our data provides structural insights into the Sel1‐like repeat family of proteins that remain poorly studied. This article is protected by copyright. All rights reserved.
Article
The intracellular bacterial pathogen Legionella pneumophila ( L.p.) manipulates eukaryotic host ubiquitination machinery to form its replicative vacuole. While nearly 10% of L.p.’s ∼330 secreted effector proteins are ubiquitin ligases or deubiquitinases, a comprehensive measure of temporally resolved changes in the endogenous host ubiquitinome during infection has not been undertaken. To elucidate how L.p hijacks host cell ubiquitin signaling, we generated a proteome-wide analysis of changes in protein ubiquitination during infection. We discover that L.p. infection increases ubiquitination of host regulators of subcellular trafficking and membrane dynamics, most notably ∼40% of mammalian Ras superfamily small GTPases. We determine that these small GTPases undergo non-degradative ubiquitination at the Legionella-containing vacuole membrane. Finally, we find that the bacterial effectors SidC/SdcA play a central role in cross-family small GTPase ubiquitination, and that these effectors function upstream of SidE-family ligases in the poly-ubiquitination and retention of GTPases in the LCV membrane. This work highlights the extensive reconfiguration of host ubiquitin signaling by bacterial effectors during infection and establishes simultaneous ubiquitination of small GTPases across the Ras superfamily as a novel consequence of L.p. infection. Our findings position L.p. as a tool to better understand how small GTPases can be regulated by ubiquitination in uninfected contexts.
Article
In the tug-of-war between host and pathogen, both evolve to combat each other's defence arsenals. Intracellular phagosomal bacteria have developed strategies to modify the vacuolar niche to suit their requirements best. Conversely, the host tries to target the pathogen-containing vacuoles towards the degradative pathways. The host cells use a robust system through intracellular trafficking to maintain homeostasis inside the cellular milieu. In parallel, intracellular bacterial pathogens have coevolved with the host to harbour strategies to manipulate cellular pathways, organelles, and cargoes, facilitating the conversion of the phagosome into a modified pathogen-containing vacuole (PCV). Key molecular regulators of intracellular traffic, such as changes in the organelle (phospholipid) composition, recruitment of small GTPases and associated effectors, soluble N-ethylmaleimide-sensitive factor-activating protein receptors (SNAREs), etc., are hijacked to evade lysosomal degradation. Legionella, Salmonella, Coxiella, Chlamydia, Mycobacterium, and Brucella are examples of pathogens which diverge from the endocytic pathway by using effector-mediated mechanisms to overcome the challenges and establish their intracellular niches. These pathogens extensively utilise and modulate the end processes of secretory pathways, particularly SNAREs, in repurposing the PCV into specialised compartments resembling the host organelles within the secretory network; at the same time, they avoid being degraded by the host’s cellular mechanisms. Here, we discuss the recent research advances on the host–pathogen interaction/crosstalk that involves host SNAREs, conserved cellular processes, and the ongoing host–pathogen defence mechanisms in the molecular arms race against each other. The current knowledge of SNAREs, and intravacuolar bacterial pathogen interactions, enables us to understand host cellular innate immune pathways, maintenance of homeostasis, and potential therapeutic strategies to combat ever-growing antimicrobial resistance.
Article
Full-text available
Legionella pneumophila is a gram‐negative bacteria found in natural and anthropogenic aquatic environments such as evaporative cooling towers, where it reproduces as an intracellular parasite of cohabiting protozoa. If L. pneumophila is aerosolized and inhaled by a susceptible person, bacteria may colonize their alveolar macrophages causing the opportunistic pneumonia Legionnaires' disease. L. pneumophila utilizes an elaborate regulatory network to control virulence processes such as the Dot/Icm Type IV secretion system and effector repertoire, responding to changing nutritional cues as their host becomes depleted. The bacteria subsequently differentiate to a transmissive state that can survive in the environment until a replacement host is encountered and colonized. In this review, we discuss the lifecycle of L. pneumophila and the molecular regulatory network that senses nutritional depletion via the stringent response, a link to stationary phase‐like metabolic changes via alternative sigma factors, and two‐component systems that are homologous to stress sensors in other pathogens, to regulate differentiation between the intracellular replicative phase and more transmissible states. Together, we highlight how this prototypic intracellular pathogen offers enormous potential in understanding how molecular mechanisms enable intracellular parasitism and pathogenicity.
Article
Full-text available
A lambda gt11 expression library of Tn5-tagged invasion plasmid pWR110 (from Shigella flexneri serotype 5, strain M90T-W) contained a set of recombinants encoding a 60-kilodalton protein (designated IpaH) recognized by rabbit antisera raised against S. flexneri invasion plasmid antigens (J. M. Buysse, C. K. Stover, E. V. Oaks, M. M. Venkatesan, and D. J. Kopecko, J. Bacteriol. 169:2561-2569, 1987). Southern blot analysis of wild-type S. flexneri serotype 5 invasion plasmid DNA (pWR100) digested with various combinations of five restriction enzymes and hybridized with defined ipaH probes showed complex hybridization patterns resulting from multiple copies of the ipaH gene on pWR100. DNA sequence analysis of a 2.9-kilobase (kb) EcoRI fragment directing IpaH antigen synthesis in plasmid recombinant pWR390 revealed an open reading frame coding for a 532-amino-acid protein (60.8 kilodaltons); this size matched well with the estimated size of IpaH determined by Western blot analysis of M90T-W cells and maxicell analysis of Escherichia coli HB101(pWR390) transformants. Examination of the amino acid sequence of IpaH revealed a hydrophilic protein with six evenly spaced 14-residue (L-X2-L-P-X-L-P-X2-L-X2-L) repeat motifs in the amino-terminal end of the molecule. Southern blot analysis of HindIII-digested pWR100 DNA probed with defined segments of the pWR390 2.9-kb insert demonstrated that the multiple band hybridization pattern resulted from repeats of a significant portion of the ipaH structural gene in five distinct HindIII fragments (9.8, 7.8, 4.5, 2.5, and 1.4 kb). Affinity-purified IpaH antibody, used to monitor the expression of the antigen in M90T-W cells grown at 30 and 37 degrees C, showed that IpaH synthesis was not regulated by growth temperature.
Article
Legionella pneumophila, the causative agent of Legionnaires' pneumonia, replicates within alveolar macrophages by preventing phagosome-lysosome fusion. Here, a large number of mutants called dot (defective for organelle trafficking) that were unable to replicate intracellularly because of an inability of the bacteria to alter the endocytic pathway of macrophages were isolated. The dot virulence genes encoded a large putative membrane complex that functioned as a secretion system that was able to transfer plasmid DNA from one cell to another.
Article
Salmonella typhimurium translocates effector proteins into host cells via the SPI1 type III secretion system to induce responses such as membrane ruffling and internalization by non-phagocytic cells. Activation of the host cellular RhoGTPase Cdc42 is thought to be a key event during internalization. The translocated Salmonella protein SopE is an activator for Cdc42. Because SopE is absent from most S. typhimurium strains it remains unclear whether all S. typhimurium strains rely on activation of Cdc42 to invade host cells. We have identified SopE2, a translocated effector protein common to all S. typhimurium strains. SopE2 is a guanine nucleotide exchange factor for Cdc42 and shows 69% sequence similarity to SopE. Analysis of S. typhimurium mutants demonstrated that SopE2 plays a role in recruitment of the actin-nucleating Arp2/3 complex to the membrane ruffles and in efficient host cell invasion. Transfection experiments showed that SopE2 is sufficient to activate host cellular Cdc42, to recruit the actin-nucleating Arp2/3 complex and to induce actin cytoskeletal rearrangements and internalization. In conclusion, as a result of SopE2 all S. typhimurium strains tested have the capacity to activate Cdc42 signalling inside host cells which is important to ensure efficient entry.
Article
The Agrobacterium VirB/D4 transport system mediates the transfer of a nucleoprotein T complex into plant cells, leading to crown gall disease. In addition, several Virulence proteins must somehow be transported to fulfill a function in planta. Here, we used fusions between Cre recombinase and VirE2 or VirF to directly demonstrate protein translocation into plant cells. Transport of the proteins was monitored by a Cre-mediated in planta recombination event resulting in a selectable phenotype and depended on the VirB/D4 transport system but did not require transferred DNA.
Article
We have developed a new vector strategy for the insertion of foreign genes into the genomes of gram negative bacteria not closely related to Escherichia coli. The system consists of two components: special E. coli donor strains and derivatives of E. coli vector plasmids. The donor strains (called mobilizing strains) carry the transfer genes of the broad host range IncP–type plasmid RP4 integrated in their chromosomes. They can utilize any gram negative bacterium as a recipient for conjugative DNA transfer. The vector plasmids contain the P–type specific recognition site for mobilization (Mob site) and can be mobilized with high frequency from the donor strains. The mobilizable vectors are derived from the commonly used E. coli vectors pACYC184, pACYC177, and pBR325, and are unable to replicate in strains outside the enteric bacterial group. Therefore, they are widely applicable as transposon carrier replicons for random transposon insertion mutagenesis in any strain into which they can be mobilized but not stably maintained. The vectors are especially useful for site–directed transposon mutagenesis and for site–specific gene transfer in a wide variety of gram negative organisms.
Article
Salmonellae encode two virulence-associated type III secretion systems (TTSS) within Salmonella pathogenicity islands 1 and 2 (SPI1 and SPI2). Two Salmonella typhimurium genes, sspH1 and sspH2, that encode proteins similar to the Shigella flexneri and Yersinia species TTSS substrates, IpaH and YopM, were identified. SspH1 and SspH2 are proteins containing leucine-rich repeats that are differentially targeted to the SPI1 and SPI2 TTSS. sspH2 transcription was induced within RAW264.7 macrophages, and was dependent upon the SPI2-encoded regulator ssrA/ssrB. In contrast, sspH1 transcription is independent of SPI2, and is not induced after bacterial phagocytosis by eukaryotic cells. Infection of eukaryotic cells with strains expressing a SspH2–CyaA fusion protein resulted in SPI2 TTSS-dependent cAMP increases. In contrast, SspH1–CyaA-mediated cAMP increases were both SPI1 and SPI2 TTSS dependent. sspH2-like sequences were found in most Salmonella serotypes examined, whereas sspH1 was detected in only one S. typhimurium isolate, indicating that the copy number of sspH genes can be variable within Salmonella serotypes. S. typhimurium deleted for both sspH1 and sspH2 was not able to cause a lethal infection in calves, indicating that these genes participate in S. typhimurium virulence for animals.
Article
Legionella pneumophila survives in aquatic environments, but replicates within amoebae or the alveolar macrophages of immunocompromised individuals. Here, the signal transduction pathway that co-ordinates L. pneumophila virulence expression in response to amino acid depletion was investigated. To facilitate kinetic and genetic studies, a phenotypic reporter of virulence was engineered by fusing flaA promoter sequences to a gene encoding green fluorescent protein. When subjected to amino acid depletion, L. pneumophila accumulated ppGpp and converted from a replicative to a virulent state, as judged by motility and sodium sensitivity. ppGpp appeared to initiate this response, as L. pneumophila induced to express the Escherichia coli RelA ppGpp synthetase independently of nutrient depletion accumulated ppGpp, exited the exponential growth phase and expressed flaAgfp, motility, sodium sensitivity, cytotoxicity and infectivity, five traits correlated with virulence. Although coincident with the stationary phase, L. pneumophila virulence expression appeared to require an additional factor: mutant Lp120 accumulated ppGpp and acquired two stationary phase traits but none of six virulence phenotypes analysed. We propose that, when nutrients are limiting, ppGpp acts as an alarmone, triggering the expression of multiple traits that enable L. pneumophila to escape its spent host, to survive and disperse in the environment and to re-establish a protected intracellular replication niche.
Article
Four new antibiotic-resistant derivatives of the broad-host-range (bhr) cloning vector pBBR1MCS have been constructed. These new plasmids have several advantages over many of the currently available bhr vectors in that: (i) they are relatively small (< 5.3 kb), (ii) they possess an extended multiple cloning site (MCS), (iii) they allow direct selection of recombinant plasmid molecules in Escherichia coli via disruption of the LacZα peptide, (iv) they are mobilizable when the RK2 transfer functions are provided in trans and (v) they are compatible with IncP, IncQ and IncW group plasmids, as well as with ColE1- and P15a-based replicons.
Article
The interactions between the L. pneumophila phagosome and monocyte lysosomes were investigated by prelabeling the lysosomes with thorium dioxide, an electron-opaque colloidal marker, and by acid phosphatase cytochemistry. Phagosomes containing live L. pneumophila did not fuse with secondary lysosomes at 1 h after entry into monocytes or at 4 or 8 h after entry by which time the ribosome-lined L. pneumophila replicative vacuole had formed. In contrast, the majority of phagosomes containing formalin-killed L. pneumophila, live Streptococcus pneumoniae, and live Escherichia coli had fused with secondary lysosomes by 1 h after entry into monocytes. Erythromycin, a potent inhibitor of bacterial protein synthesis, at a concentration that completely inhibits L. pneumophila intracellular multiplication, had no influence on fusion of L. pneumophila phagosomes with secondary lysosomes. However, coating live L. pneumophila with antibody or with antibody and complement partially overcame the inhibition of fusion. Also activating the monocytes promoted fusion of a small proportion of phagosomes containing live L. pneumophila with secondary lysosomes. Acid phosphatase cytochemistry revealed that phagosomes containing live L. pneumophila did not fuse with either primary or secondary lysosomes. In contrast to phagosomes containing live bacteria, the majority of phagosomes containing formalin-killed L. pneumophila were fused with lysosomes by acid phosphatase cytochemistry. The capacity of L. pneumophila to inhibit phagosome-lysosome fusion may be a critical mechanism by which the bacterium resists monocyte microbicidal effects.