ArticlePDF Available

The Legionella pneumophila effector Ceg4 is a phosphotyrosine phosphatase that attenuates activation of eukaryotic MAPK pathways

Authors:

Abstract and Figures

Host colonization by Gram negative pathogens often involves delivery of bacterial proteins called ″effectors″ into the host cell. The pneumonia-causing pathogen Legionella pneumophila delivers more than 330 effectors into the host cell via its type IVB Dot/Icm secretion system. The collective functions of these proteins is the establishment of a replicative niche from which Legionella can recruit cellular materials to grow while evading lysosomal fusion inhibiting its growth. Using a combination of structural, biochemical, and in vivo approaches, we show that one of these translocated effector proteins, Ceg4, is a phosphotyrosine phosphatase harboring a haloacid dehalogenase hydrolase domain. Ceg4 could dephosphorylate a broad range of phosphotyrosine-containing peptides in vitro and attenuated activation of MAPK controlled pathways in both yeast and human cells. Our findings indicate that L. pneumophila's infectious program includes manipulation of phosphorylation cascades in key host pathways. The structural and functional features of the Ceg4 effector unraveled here, provide first insight into its function as a phosphotyrosine phosphatase, paving the way to further studies into L. pneumophila pathogenicity.
Content may be subject to copyright.
Legionella phosphotyrosine phosphatase activity towards MAPKs
1
The Legionella pneumophila effector Ceg4 is a phosphotyrosine phosphatase that attenuates activation of
eukaryotic MAPK pathways
Andrew Quaile1, Peter J Stogios1, Olga Egorova1, Elena Evdokimova1, Dylan Valleau1, Boguslaw
Nocek2, Purnima S Kompella3, Sergio Peisajovich3, Alexander F Yakunin1, Alexander W
Ensminger4, Alexei Savchenko1,5*
1Department of Chemical Engineering and Applied Chemistry, University of Toronto, Toronto, ON,
Canada.
2Structural Biology Center, Advanced Photon Source, Argonne National Laboratory.
3Department of Cell and Systems Biology, University of Toronto, Toronto, ON, Canada,
4Department of Biochemistry, Department of Molecular Genetics, University of Toronto, Toronto, ON,
Canada
5Department of Microbiology, Immunology and Infectious Diseases, Cumming School of Medicine,
University of Calgary, Calgary, AB, Canada
Running title: Legionella phosphotyrosine phosphatase activity towards MAPKs
.
To whom correspondence should be addressed: Dr. Alexei Savchenko. 5Department of Microbiology,
Immunology and Infectious Diseases, Cumming School of Medicine, University of Calgary, 3330 Hospital
Drive NW, Calgary, Alberta, Canada; Telephone: (403)-210-7980; E-mail: alexei.savchenko@ucalgary.ca
Keywords: Legionella pneumophila, crystallography, phosphotyrosine, protein phosphatase, peptide array,
p38 MAPK, bacterial effector.
ABSTRACT
Host colonization by Gram-negative
pathogens often involves delivery of bacterial
proteins called “effectors” into the host cell. The
pneumonia-causing pathogen Legionella
pneumophila delivers more than 330 effectors
into the host cell via its type IVB Dot/Icm
secretion system. The collective functions of
these proteins is the establishment of a replicative
niche from which Legionella can recruit cellular
materials to grow while evading lysosomal fusion
inhibiting its growth. Using a combination of
structural, biochemical, and in vivo approaches,
we show that one of these translocated effector
proteins, Ceg4, is a phosphotyrosine phosphatase
harboring a haloacid dehalogenasehydrolase
domain. Ceg4 could dephosphorylate a broad
range of phosphotyrosine-containing peptides in
vitro and attenuated activation of MAPK-
controlled pathways in both yeast and human
cells. Our findings indicate that L. pneumophila’s
infectious program includes manipulation of
phosphorylation cascades in key host pathways.
The structural and functional features of the Ceg4
effector unraveled here, provide first insight into
its function as a phosphotyrosine phosphatase,
paving the way to further studies into L.
pneumophila pathogenicity.
Host colonization by Gram-negative
pathogens often involves delivery of specific sets
of bacterial proteins called “effectors” into the
host cell by specialized secretion systems. Acting
in concert, effectors secure the pathogen’s
survival and replication, via manipulation of host
processes and the establishment of subcellular
environments that favor the pathogen. Disruption
of effector translocation (1-3) results in
attenuation of intracellular growth, highlighting
the essential role of effectors in pathogen-host
interactions. Despite significant progress in
identification of effector arsenals in bacterial
genomes and in defining the molecular functions
of individual effectors, many remain
uncharacterized, necessitating further studies into
their role in pathogenesis. The composition and
number of effectors varies dramatically between
http://www.jbc.org/cgi/doi/10.1074/jbc.M117.812727The latest version is at
JBC Papers in Press. Published on January 4, 2018 as Manuscript M117.812727
Copyright 2018 by The American Society for Biochemistry and Molecular Biology, Inc.
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
2
different pathogens. For example, plague and
intestinal disease-causing Yersinia species
encode less than ten effectors delivered by the
type III secretion system (4) while the similar
secretion system in Shigella spp. is involved in
translocation of close to 30 effectors (5). Notably,
sequence-related effectors are found in pathogens
with very diverse invasion strategies suggesting
that these effector families are involved in
common host manipulation tactics.
A causative agent of severe pneumonia in
humans, the Legionella pneumophila genome
encodes over 330 effector proteins, that are
translocated via the Dot/Icm (defect in organelle
trafficking/ intracellular multiplication), type
IVB secretion system (e.g. (6-11), or see (12) for
a recent review). Legionella’s effectors account
for more than 10% of its proteome (13) and
represents the largest effector set known to the
bacterial world. In their natural habitat of fresh
water reservoirs Legionella spp. invade diverse
amoebae by preventing formation of the
phagolysosome (14) followed by modification of
the newly co-opted compartment into an
organelle ideal for intracellular replication of the
bacteria, the Legionella containing vacuole
(LCV) (15). The ability to apply the same
invasion strategy to invade human alveolar
macrophages raises the intriguing possibility that
Legionella’s effectors target host processes that
are conserved between distant eukaryotic phyla.
The sheer number of Legionella effectors and
their apparent functional redundancy makes their
functional characterization particularly
challenging (16).
Studies into the function of bacterial
effectors suggested that these pathogenic factors
demonstrate unparalleled abilities to manipulate
a wide spectrum of host cell’s processes
including facilitating alterations in cytoskeletal
rearrangement (17,18), vesicular trafficking
(19,20), signal transduction (21,22) and
transcription regulation (23,24). To achieve these
effects, effectors are often involved in post-
translational modification (PTM) of specific host
proteins. A key regulation mechanism in both
eukaryotic and bacterial cells, PTMs typically
involve enzymatic covalent modification of
targeted proteins at specific residues, which
affect that protein’s activity, localization or
interactions thus triggering a change in protein
function. Effector proteins can possess not only
the PTM activities found in the bacterial world
but also ostensibly exclusively eukaryotic ones.
Of particular prevalence, bacterial effectors have
evolved to mimic the activity of ubiquitin protein
ligases, which control the final step in the
eukaryote-specific PTM involving attachment of
the ubiquitin polypeptide to targeted proteins,
usually resulting in this protein’s degradation by
the proteasome (25). Effectors with this PTM
activity have been identified in the arsenals of
many bacterial pathogens, including E. coli
(26,27), Shigella flexneri (24,28), Salmonella
spp. (26,29) and Legionella pneumophila (30,31).
The most common PTM in eukaryotic
cells is phosphorylation, in which a phosphate
group from a donor molecule such as ATP onto
hydroxyl functional groups on residues (serine,
threonine or tyrosine) of the targeted protein
(32,33). This PTM is catalyzed by kinases and the
human genome encodes several hundred protein
kinases divided into tyrosine and serine/threonine
specific enzymes (34). This PTM mechanism is
involved in most if not all known human cell
processes.
One of the best-studied examples of
phosphorylation-controlled signaling is mediated
by Mitogen-Activated Protein Kinases (MAPKs).
MAPKs are highly conserved Ser/Thr protein
kinases that have been extensively studied for
their central roles in mediating signal
transduction of extracellular stimuli to the
appropriate biological response (35). MAPKs are
activated by dual phosphorylation of threonine
and tyrosine residues in a Thr-X-Tyr motif
located in their activation loops by upstream
kinases (MAPK kinases, or MAPKKs) (36). In
turn, MAPKs are responsible for phosphorylating
and activating downstream MAPK activated
protein kinases (MAPKAPKs or MKs). These
downstream targets elicit activation of processes
including regulation of stress response,
proliferation, differentiation and apoptosis. Their
control of processes highly relevant to bacterial
infections as well as their careful regulation and
conservation among eukaryotes have made
MAPKs attractive targets for bacterial effectors;
Icm/Dot substrates LegK1-4 share
homology to eukaryotic protein kinases and
activation of MAPKs in response to Legionella is
well documented (37-39). Furthermore, while
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
3
phosphorylation of SAP/JNK, ERK1/2 and p38
are seen at early time-points even in translocation
deficient mutants (39), sustained activation of
SAP/JNK and p38 is reliant on effector
translocation (38).
Although protein phosphatases have not
been previously detected amongst Legionella’s
complement of effectors (40) precedent for this
mechanism amongst other effectors of other
species undoubtedly exists; Yersinia pestis YopH
is a phosphotyrosine phosphatase capable of
removing the pTyr signal through hydrolysis,
thereby muting its activity (41), while Shigella
flexneri OspF is a phosphothreonine lyase that
irreversibly dephosphorylates the activating
threonine of MAPKs (42). A recent analysis of
the effector repertoire in the causative agent of Q
fever, Coxiella burnetii also identified effectors
Cbu1676 and Cbu0885 as phosphatases targeting
the MAP kinase pathway in the eukaryotic host
surrogate S. cerevisiae (43). S. cerevisiae
possesses five main MAP kinase pathways
regulating filamentation, cell wall integrity,
sporulation, mating and hyperosmosis, with the
latter two pathways, controlled by Fus3 and Hog1
respectively, being most broadly conserved
among diverse eukaryotic organisms (44). The
closest human homologues of Fus3 and Hog1, are
ERK2 and p38. The first indication of Cbu1676
and Cbu0885 function came from in silico
analysis that pointed to the presence of the
haloacid dehalogenase-like (HAD-like) domain
in these bacterial proteins (43). Proteins
containing HAD-like domains have a broad range
of activities including dehalogenase,
phosphonatase and phosphomutase activity and
although the majority characterized thus far are
phosphatases and ATPases (45,46), protein
phosphatase activity has only been observed in
eukaryotes (47-50).
All HAD-like domains share a common
overall fold featuring a core Rossmann fold,
consisting of at least two pairs of α-helices that
sandwich the core five-stranded parallel β-sheet
in the order ‘54123’, with a squiggle and flap
motif at the end of β1-strand (45,46). The
common molecular architecture between HAD
superfamily members includes four (I to IV)
highly conserved sequence motifs co-localized to
the active site. Motifs I and IV feature conserved
aspartate residues that coordinate the Mg2+ ion
required for catalysis. In addition to squiggle and
flap motifs, HAD-like domains typically contain
an insertion to the catalytic core domain called a
cap domain, which controls active site access and
is involved in substrate binding (51-53). Despite
these recognizable sequence motifs, significant
variation in substrate specificity and activity of
the HAD-like protein family necessitates detailed
structural analysis and rigorous substrate
specificity studies.
In addition to the Coxiella effectors
mentioned above, putative HAD-like domains
have been identified in the uncharacterized
effectors Lpg0096 (also known as Ceg4 / co-
regulated with the effector encoding genes 4”),
Lpg1101 and Lpg2555 from L. pneumophila
(43). Here, using x-ray crystallography and
biochemical activity screening, we show that
Ceg4 is an atypical HAD-like phosphotyrosine
phosphatase able to attenuate the activation of
MAP kinases in both human and yeast cells.
These results indicate that L. pneumophila
facilitates its infectious program by manipulation
of phosphorylation cascades of key pathways in
its host cells.
RESULTS
Legionella effector Ceg4 demonstrates
phosphotyrosine specific phosphatase activity in
vitroThe Dot/Icm-dependent translocation of L.
pneumophila Ceg4 protein encoded by the
lpg0096 gene was previously demonstrated using
CyaA fusion translocation assays (8). Our
sequence analysis using Phobius (54) suggested
that in addition to the N-terminal haloacid
dehalogenase-like (HAD-like) domain
mentioned above, this effector contains two C-
terminal transmembrane (TM) helices (residues
266-289 and 295-320). As no TM signatures were
detected in Lpg1101 or Lpg2555 we hypothesize
that Lpg0096/Ceg4 may be a member of a
functionally diversified family that relies of
localization to appropriately direct its activity
toward the host. Using data from two recent large
scale comparative genomics studies of Legionella
species (55), we performed phylogenetic analysis
of Ceg4 with the 33 other putative HAD domain-
containing effectors (Figure 1). Ceg4, Lpg1101
and additional 12 homologues formed a distinct
clade corresponding to Legionella orthologue
group LOG_02908, while Lpg2555 and five
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
4
other effectors were localized in a distinct group.
Comparative sequence analysis across the Ceg4
containing clade indicated that sequence
conservation among these effectors was highest
within the HAD-like domain, with significant
sequence variation in their C-terminal domains.
Lani_0822 from L. anisa was the sole exception
to this observation, possessing an additional 50
residues at both the N and C termini. Lpg1101 is
also unique within LOG_02908, having
homology to the phosphatidylinositol 4-
phosphate binding domain of SidM/DrrA-like
(56,57) at its C-terminal. Most notably however,
eight out of the 14 Ceg4 homologues possessed
two TM domains in their C-terminal portion,
suggesting that localization to the host membrane
is an important and common feature to this subset
of HAD-like effectors in Legionella.
To confirm the general activity of the
Ceg4 HAD-like domain, we purified an N-
terminal fragment spanning residues 1 to 193 of
L. pneumophila Ceg4 (see Materials and Methods
for details). In line with predictions, the Ceg4[1-
193] fragment demonstrated robust phosphatase
activity against the generic phosphatase substrate
para-nitrophenylphosphate (pNPP) (58). Highest
reaction rates were observed between pH 6.5 and
8, with activity dropping markedly above pH 8.
Ceg4 also demonstrated a strict requirement for
Mg2+ metal ions (Supplementary data, Figure
S1B,C,D). Expanding on these results, we tested
Ceg4[1-193] for activity against a library of 94
phosphorylated metabolic substrates allowing for
querying a broad range of possible specificities
(58). In these assays, Ceg4[1-193] demonstrated
the highest activity toward phosphotyrosine
(Figure 2C, see also Supplemental Figure S1E).
To determine if Ceg4[1-193] is active against
protein substrates, we screened for its ability to
remove the phosphate group from a selection of
53 phosphopeptides, chosen for their importance
to signaling in Saccharomyces cerevisiae, which
has been successfully used as a model eukaryotic
system in characterization of bacterial effector
functions including those from Legionella
(30,59). This set included pSer-, pThr- and pTyr-
containing sequences. Consistent with our
previous results, Ceg4[1-193] demonstrated
robust phosphatase activity against nine diverse
peptides comprising the full pool of pTyr-
containing sequences in this substrate array
(Figure 1E, Supplemental Figure S1F), and
moreover, at levels 4 to 5 times higher than for
pSer or pThr peptides. Combined, this data
showed that Ceg4 is a phosphotyrosine specific
phosphatase active against peptide substrates.
Crystal structure of HAD-like domain
provides molecular insight into phosphatase
activity of Ceg4To gain further insight into the
molecular function of Ceg4, we determined the
crystal structure of the Ceg4[1-208] fragment to
1.88 Å by the single wavelength anomalous
dispersion (SAD) method (see Table 1 for x-ray
crystallographic statistics). The final structural
model spanned Ceg4 residues 1-204 and a portion
of the N-terminal fusion tag sequence
(GQENLYFQG) corresponding to the TEV
protease cleavage site (Figure 3A). In addition to
a core Rossmann-like fold, the HAD-like domain
of Ceg4 included a cap subdomain consisting of
three α-helices and two long loops inserted
between the D9-X-D11 “squiggle motif” and the
α1 helix (Figures 2A and 2B). The location of the
cap subdomain insertion in relation to the
conserved motifs I and II of the core Rossman
fold classified it as a C1 cap (46). Such C1 caps
were previously identified in cN-III nucleotidase
(60), Eya2 protein tyrosine phosphatase (61) and
MDP-1 sugar phosphatase (62,63). However,
according to our analysis, the Ceg4 cap
subdomain did not show any significant structural
similarity with these or any other C1 caps of
structurally-characterized HAD-like domains
(Supplementary Figure S2).
In addition to overall structural similarity
to other Rossmann-like folds, the core fold of
Ceg4 possessed several conserved features
consistent with the canonical motifs of other
HAD-like phosphatases, such as D9 and D11
(motif I), T103 (motif II), K135 (motif III), D157
and D158 (motif IV) (Figure 3C). Collectively,
these residues formed a small pocket ~350 Å3 in
volume, with N20 from the cap subdomain
forming a “lid” over the pocket (Figure 3D).
Inspection of the active site appears to confirm
the correct positioning of D9 and D11 for their
putative functions as the nucleophile and general
acid/base residues respectively. The active site
also contained a Mg2+ ion that was coordinated by
D9, D158 and three ordered water molecules; the
position of the magnesium ion is conserved with
those of mono- and divalent ions bound to other
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
5
HAD-like phosphatases (60-62). The active site
also contained a Cl- ion associated with the
sidechain of K135, the backbone of K104 and two
ordered water molecules; its position is conserved
with the position of sulfate or phosphate ions
trapped in the active site of other crystallized
HAD-like phosphatases (60-62). Finally, the
active site also contained a highly-coordinated
water that interacted with the F8, D9, D11, T103
and the Cl- ion.
To confirm the involvement of these
residues with Ceg4 catalytic activity, we
performed site-directed mutagenesis and tested
the resultant Ceg4[1-193] mutants in vitro for
activity towards pNPP and pTyr substrates as
described above. In accordance with their
predicted contributions to catalysis, the D9A,
D11A, D11N, D157A, D158A and D162A
mutations abrogated phosphatase activity,
validating their essentiality for the catalytic
function of this protein (Supplementary Figure
S3). Inspection of the molecular packing in
the Ceg4[1-208] crystal lattice did not reveal
significantly extensive contacts indicative of
oligomerization. However, by size exclusion
chromatography (Supplementary Figure S4A),
we observed a mixture of monomeric and dimeric
species. Consistent with this, we observed that
the nine residues that we resolved of the N-
terminal fusion tag comprising the TEV protease
cleavage sequence (G(-8)Q(-7)E(-6)N(-5)L(-
4)Y(-3)F(-2)Q(-1)G(0)) contacted the active site
of the adjacent molecule in the crystal lattice
(Supplementary Figure S4B). This peptide
adopted an extended, nearly β-strand-like
conformation (Supplementary Figure S4C).
Notably, the tyrosine residue in the tag sequence
deeply bound in the active site pocket, with its
hydroxyl group forming a network of interactions
including hydrogen bonds with the sidechain of
D11, the bound Cl- ion, and with two ordered
water molecules (Supplementary Figure S4C).
Other interactions included hydrogen bonds
between the sidechain of Q(-1) of the tag and
E40, between the backbone amide of F(-2)
residue of the tag and N20 of the C1 cap, between
the sidechain of K104 and the backbone carbonyl
of Q(-6), and R(-7) of the tag formed two
interactions (with E165 and the backbone
carbonyl of G133). The C1 cap also interacts
with the fusion tag peptide via hydrophobic
interactions between Y43 and Y(-3) and F24 and
Q(-1) and a hydrogen bond between E40 and Q(-
1). Given Ceg4 specificity toward pTyr
peptides, we hypothesized that our observed
binding of the tyrosine from the N-terminal
fusion tag may be representative of Ceg4
interactions with its substrate. To further
examine this, we modified the N-terminal fusion
TEV cleavage sequence to match the sequence
one of pTyr carrying peptides Q(-7)M(-6)T(-
5)G(-4)Y(-3)V(-2)S(-1)T(0) identified as a Ceg4
substrate in our peptide array screening and
representing the activation loop sequence of the
yeast Hog1 MAP kinase (64). Hog1 is involved
in a signaling pathway regulating yeast
hyperosmotic adaptation (64) and is a close
homologue of human p38 MAP kinase, a
pathway previously implicated in Legionella
pathogenesis (38,39,65). As mentioned
previously, Coxiella HAD-like effectors
efficiently modulated the activity of other yeast
MAP kinases prompting us to hypothesize that
our in vitro activity results may also be indicative
of Ceg4 activity against MAPK kinases.
The structure of the tag-modified
Ceg4[1-208]HOG1p fragment was solved to 1.9 Å
by Molecular Replacement (Table 1). This crystal
structure was almost identical to the original
Ceg4[1-208]TEV site structure described above, and
superimposed with RMSD of less than 0.3 Å
across the entire protein backbone. In this crystal
structure, we resolved the positions of the six
Hog1-derived residues T(-5)G(-4)Y(-3)V(-2)S(-
1)T(0) from the modified fusion tag. As with the
Ceg4[1-208]TEV site structure, the crystal packing
of Ceg4[1-208]HOG1p showed the N-terminal
peptide interacted with the active site of the
adjacent Ceg4[1-208]HOG1p molecule (Figure 4A),
with the Y(-3) residue bound deeply in the pocket
(Figure 4B). The HOG1p peptide also adopted
an extended β-strand like conformation and its
structure was strikingly similar with that of the
TEV cleavage site peptide, most especially across
residues -5 through 0 (RMSD 0.5 Å over the six
matching atoms) (Figure 4C). As we could
resolve only a shorter region of this peptide, we
were able to identify fewer interactions between
this peptide and Ceg4[1-208], and those were
similar the ones observed in the Ceg4[1-208]TEV
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
6
site structure (i.e. the interactions between Y(-3)
and the active site, N20 and the backbone amide
of residue (-2) of the tag, and hydrophobic
interactions between Y43 and Y(-3) plus F24 and
residue -1).
Next, we compared the position of the
tyrosine from the expression tags and the active
site configuration of the Ceg4[1-208] active site
with other structurally characterized HAD
phosphatases. This analysis showed that the
fusion tag tyrosine hydroxyl group adopts a
position that is 3.4 Å from the bound chlorine
atom, which itself occupied the same general
position as phosphate or phosphate analogs (i.e.
phosphate bound to E. coli YrbI (PDB 3I6B, (66))
or beryllium trifluoride bound to Eya2 (PDB
3HB0, (61)) (Supplementary Figure S2B). The
position of the Ceg4[1-208]-bound magnesium
ions were also conserved in position with other
HAD-like phosphatases (Supplementary Figure
S2B). This analysis indicates that the active site
configuration observed in the Ceg4[1-208]
crystal structures is a good approximation of the
position of a phosphotyrosine substrate bound to
this enzyme.
Overall, our structural analysis revealed
a compact active site in the HAD-like domain of
Ceg4[1-208] restricted by a unique cap motif.
This active site is able to accommodate the
phosphotyrosine residue from a peptide substrate,
the position of which can be gleaned from the
conformation of the fusion tag in the Ceg4[1-208]
crystal lattice.
Ceg4 shows activity against Hog1 and
Fus3 MAP kinases in Saccharomyces cerevisiae
According to our phosphopeptide library
screening, Ceg4[1-193] demonstrated equally
robust activity against peptides QMTGpYVSTR
and GMTEpYVATR representing the activation
loops of yeast Hog1 and Fus3 MAP kinases,
respectively (Fig. 1C). To test if in vitro activity
of the Ceg4[1-193] fragment against yeast MAP
kinase phosphopeptides is representative of in
vivo activity of this effector, we tested the ability
of full-length Ceg4 to affect the activation of
Hog1- and Fus3-controlled pathways in a yeast
model system. For this we used two S. cerevisiae
strains engineered to express fluorescent reporter
proteins (Stl2-BFP or Fus1-GFP) upon activation
of Hog1 or Fus3 controlled pathways,
respectively (see Material and Methods for
details). We expressed full-length Ceg4[1-397] or
the Ceg4[1-208] fragment in these S. cerevisiae
strains and measured the overall fluorescence
signal from 10,000 cells. S. cerevisiae strains
overexpressing Ceg4 showed significant
reduction of activation of both Hog1 and Fus3
activated pathways (42% and 56%, respectively),
as compared to the control strain carrying an
empty vector (Figure 5A). The strain expressing
the Ceg4[1-208] fragment showed a reduced
ability to suppress MAPK activation in the case
of Fus3-controlled activation compared to the
strain expressing the full-length effector. In
contrast, the expression of the Ceg4[1-208]
fragment resulted in further decrease in Hog1-
controlled activation compared to the strain
expressing the full-length effector. Based on
these results, Legionella Ceg4 is implicated in
regulation of host MAPK controlled pathways as
demonstrated by reduced expression of
fluorescent reporters for both Fus3 and Hog1. In
addition, our data suggested that Ceg4 activity
against the Fus3 pathway is dependent on the C-
terminal region of the effector that contains the
membrane-spanning elements, pointing to the
potential role of this domain for in vivo specificity
of Ceg4.
Next, to test the link between Ceg4
phosphatase activity and MAPK regulation we
probed the effect of individual Ceg4 active site
residue substitutions on the ability of this effector
to dampen the activation of Hog1-controlled
pathways in yeast. The Ceg4 residues targeted by
this analysis were chosen based on their direct
involvement in phosphatase catalytic activity
(D11, K135, D158, D162 and), and their
participation in forming the active site by the cap
subdomain (K17, S18, N20, V23, F24, E26 and
Y43) and the core HAD-like domain (K104,
E108, E131, T161 and N186) (Figure 5B and C).
In line with our in vitro activity results,
alanine substitution of the conserved D11, K135,
D158 and D162 residues directly involved in
catalytic activity of the HAD-like domain
resulted in abrogation of Ceg4 suppression of the
hyperosmotic stress response. This observation
confirmed our hypothesis that Ceg4-triggered
dampening of MAPK activation is directly linked
to the phosphatase activity of its HAD-like
domain. Substitution of active site pocket
residues V23, F24, Y43 and K104 also had a
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
7
negative impact on the ability of Ceg4 to control
MAPK activation with V23D substitution
resulting in complete loss of activity. In contrast,
substitutions of Ceg4 K17, E26, E108, E131 and
T161 which are located distally from the active
site and N20 which partially covers the entrance
to the catalytic pocket did not significantly affect
its activity against Hog1 activated pathway.
These observations are consistent with previously
observed mechanisms of catalysis of HADs and
furthermore indicate the involvement of certain
cap domain residues in substrate recognition.
Combined, these results clearly linked
Ceg4’s in vivo activity as a regulator of MAPK
pathways with its HAD-like domain phosphatase
active site and identified individual residues
involved in catalysis and substrate interactions.
Ceg4 localizes to HeLa endoplasmic
reticulum and attenuates MAPK p38 activation in
vivoHaving determined that Ceg4 is able to act
on conserved eukaryotic MAP kinases, and
having also established a possible role of C-
terminal region of the Ceg4 effector in this
activity, we were interested in identifying the
subcellular localization of Ceg4 in human cells
and in determining if the MAPK dampening
activity extended to human MAP kinases.
To that end, human HeLa cells were
transfected with constructs expressing wild type
Ceg4, the catalytically inactive variant Ceg4D9A,
Ceg4[1-207] and Ceg4[208-397] fragments each
fused to GFP. In keeping with the presence of the
predicted transmembrane regions in the C-
terminal portion of Ceg4, the full length (both
wild type and D9A mutant) and Ceg4[208-397]
fragments demonstrated specific perinuclear
localization, while the construct Ceg4[1-207]
which lacks the C-terminal region containing the
TM domain showed diffuse localization
throughout the cell (Figure 6A). Given the
differential importance of the C-terminal domain
to dampening S. cerevisiae pheromone and
hyperosmolarity responses we posit that this
portion of Ceg4 is critically important to its
function in the host cell. We therefore sought to
more specifically determine the sub-cellular
localization of Ceg4. HeLa cells transfected with
N-terminally GFP tagged Ceg4D9A,
counterstained with ER-tracker dye demonstrated
that Ceg4 co-localizes predominantly with
endoplasmic reticular structures.
With both structural and biochemical
analyses indicating MAPK activation loop
sequences as potential targets of Ceg4
phosphatase activity, we also tested its ability to
perform this function on the closest human
homologue of Hog1, the p38 MAPK. HEK293t
cells transfected with either Ceg4 or Ceg4D9A
were tested for changes in the phosphorylation
state of human p38 MAPK upon stimulation with
either TPA or anisomycin. Western blot analysis
for both total p38 and phospho-p38 (Figure 6C)
showed that while the total levels of p38 are the
same in cells expressing the wild type and
catalytically-inactive mutant, cells harboring
wildtype Ceg4 demonstrated a significant
reduction of signal corresponding to
phosphorylated p38 compared to the same signal
in cells carrying the Ceg4D9A mutant.
Combined, our structural and functional
analysis has identified Legionella Ceg4 as a
bacterial HAD protein tyrosine phosphatase that
is able to attenuate the MAP kinase responses in
both human and yeast cells in vivo, and does so
via removal of the phosphate moiety from
phospho-tyrosine in their activation loops that are
critical to their activation.
DISCUSSION
Translocation of effector proteins inside
the host cell is an important and common strategy
adopted by many Gram-negative bacteria
including important human pathogens. This
necessitates the functional characterization of
specific effectors as a necessary step in
understanding of host-pathogen interactions and
for the development of novel anti-bacterial
therapies. Here, we show the conserved
Legionella effector Ceg4 can modulate the
phosphorylation state of eukaryotic MAP kinases
through its HAD-like phosphatase domain and
we clarify the molecular structure of this domain,
providing key molecular details into this effector.
Ceg4 is one of three predicted HAD-like
domain containing effectors in the Legionella
pneumophila genome that are known to be
translocated by the Dot/Icm system. According to
our analysis, HAD-like effectors are also found in
other Legionella species suggesting that this
functional domain is widely used by these
pathogens for manipulation of host signaling
pathways. Despite commonality among HAD-
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
8
like effectors, Ceg4 represents a sequence-
distinct group containing eight Legionella
effectors that feature the combination of an N-
terminal HAD-like domain and a C-terminal
region carrying two transmembrane helices. The
C-terminal region of the Ceg4 effector plays an
important part in interactions of this effector with
eukaryotic MAP kinases as deletion of this region
had a significant effect on the ability of this
effector to dampen the signaling by the yeast
Fus3p MAPK. Furthermore, alanine substitution
of catalytic residues D11 and D158 in the Ceg4
HAD-like domain active site abrogated
phosphatase activity and ability to dampen
MAPK activation. Combined with the broad
substrate specificity of the Ceg4 HAD-like
domain toward phosphotyrosine peptides
demonstrated by our in vitro assays, this
observation prompted us to suggest that the C-
terminal region may be responsible for tailoring
the general phosphatase activity of Ceg4 toward
its specific host target.
The crystal structure of the Ceg4[1-208]
fragment showed strong similarity with
previously structurally-characterized members of
the HAD-like protein family including specific
features of the active site such as conservation of
key catalytic residues and coordination of a Mg2+
ion, known to be essential for catalysis (51).
Probing the Ceg4 active site cavity with site-
directed mutagenesis not only confirmed the role
of residues predicted to be directly involved in
catalysis but also revealed the subset of residues
important for Ceg4 activity against MAP kinase
substrates. Substitution of residues such as D11
which acts as a general acid/base, drastically
reduced or completely abrogated the activity of
this effector against Fus3 MAPK in a yeast model
system. Notably, previous prediction of specific
residues important for Ceg4 activity based on
similarity to Coxiella HAD-like effectors (43)
was only partially confirmed by our structure and
subsequent mutagenesis. Specifically, key
residues of motifs II and III were previously
predicted to correspond to Ser81 and Lys111.
However, our structural analysis pointed instead
to Thr103 and Lys135 fulfilling this role. This
observation highlights the limits of primary
sequence based analysis applied to highly diverse
HAD-like proteins, and effectors in general and
reiterates the necessity in molecular and
structural data to complement functional
diversity across large protein families.
The substrate specificity of HAD-like
domains is often defined by the ‘cap’ subdomain
insertion that controls the access to the catalytic
center (67). The Ceg4 HAD-like domain
structure features an unusual α-helical cap motif
never before described for this domain. This
novel cap motif is compatible with the broad
specificity of this domain against
phosphotyrosine peptides as indicated by our
observation that the tyrosine residue from two
different N-terminal tag sequences is able to
make intimate contacts with the Ceg4 active site
of a neighboring molecule in two different crystal
lattices. The specific position of the tyrosine
residue is the Ceg4 active site is compatible with
the position of trapped phosphate/phosphate
analog substrates in other structurally-
characterized HAD-like phosphatases,
suggesting that this crystallization observation
may indeed be representative of the interaction
between the Ceg4 HAD-like with its
phosphoprotein substrate.
Effectors have been demonstrated to
target all strata of the MAPK controlled signaling
pathways (MAPKKK, MAPKK, MAPK and
MKs) using several differing enzymatic activities
(68,69). Characterization of Ceg4 phosphatase
activity against yeast and human MAP kinases
adds a new member to this growing list of
MAPK-modulating factors along with the
recently-characterised Coxiella HAD-like
effectors active against the yeast CWI MAPK
(43). Human HEK293t cells transfected with
Ceg4 demonstrated clear reduction of the amount
of phosphorylated p38 MAP kinase, and this was
compromised by mutation of the key Ceg4 active
site residue D9. Previous work has shown that
MAPK phosphorylation in human cells is
increased at very early time points of bacterial
challenge by Legionella, and that this activation
is sustained for some time in an effector-
dependent manner (39). Therefore, it is tempting
to speculate that Ceg4 would serve a functional
role only at significantly later points of infection
and in keeping with this model, RNA-Seq data
taken during infection show that Ceg4
transcription levels are low during the post
exponential/infectious stage but increase 10-fold
during exponential growth (70). An additional
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
9
possibility is that given the C-terminal dependent
localization to the endoplasmic retuculum
membranes and the loss of ability of Ceg4 to
modulate Fus3 activation without this domain it
is possible that Ceg4 acts to reduce specific
MAPK activation in a localized manner, while
allowing general activation throughout the rest of
the cell.
Characterization of the structural and
functional features of the Ceg4 effector presented
in this work provide the first insight into function
of this and other HAD-like effectors during
Legionella infection, paving the way to further
studies into this bacteria’s pathogenic strategy.
EXPERIMENTAL PROCEDURES
Protein purification and size exclusion
chromatography-Based on domain and
transmembrane location predictions, genes
corresponding to Ceg4 residues 1-208 and 1-193
were cloned into p15Tv-LIC-TEV by ligation
independent cloning, and transformed into E. coli
BL21 CodonPlus (DE3) RIPL competent cells
using standard procedures. Several mutants of
Ceg4[1-193] in p15Tv-LIC-TEV (D9A, D11N,
D11A, E26A, D157A, D158A and D162A) along
with a variant of Ceg4[1-208] with the TEV
sequence ‘GRQNLYFQG’ mutated to match the
activation loop sequence from S. cerevisiae Hog1
‘PQMTGYVST’ were prepared by site-directed
mutagenesis. Cultures were grown at 37 °C in M9
media with selenomethionine or LB media,
supplemented with kanamycin and at an OD600
of 0.6-0.8 and expression of 6xHis-TEV tagged
Ceg4 was induced by the addition of 0.4 mM
isopropyl-β-D-1-thiogalactopyranoside and the
temperature of the culture was reduced to 16°C
overnight. The following day, cultures were
harvested by centrifugation and pellets lysed by
sonication on ice in 50 mM HEPES (pH 7.5), 150
mM NaCl, and 1 mM PMSF. All further
purification was conducted at 4 °C. Cell lysate
was clarified by ultracentrifugation at 17,000 × g
for 30 min, and 15 ml of Ni-NTA resin (Qiagen)
was added and incubated with gentle rotation for
30 min. Resin was washed with 50 mM HEPES
(pH 7.5), 150 mM NaCl, and 30 mM imidazole,
and protein eluted with 50 mM HEPES (pH 7.5),
150 mM NaCl, and 500 mM imidazole. Proteins
were further concentrated using a centrifugal
concentrator, flash-frozen in liquid N2, and
stored at −80 °C. Oligomerization of Ceg4[1-
208] was tested by size exclusion
chromatography using a Superdex S200 column
with running buffer 50 mM HEPES (pH 7.5) and
150 mM NaCl.
General enzyme activity screening
General screens for enzyme activity were
performed as previously described (58) using
Ceg4[1-193]. Briefly, 20 µL of purified protein
(at 0.5 µg/µL) was added to wells of a 96-well
plate, then 180 µL of protease, phosphatase,
phosphodiesterase, dehydrogenase, oxidase,
NADH/NADPH oxidase and lipase, or 170 µL of
thioesterase mixes (containing buffers, metal
cations and substrates) were added, as well as 10
µL of 5,5'-dithiobis(2-nitrobenzoic acid) to the
thoiesterase mix. Plates were incubated at 37°C
for 1 hour. Phosphatase, phosphodiesterase,
protease, lipase, and thioesterase results were
read at 410 nm, dehydrogenase, oxidase, and
NADH/NADPH oxidase results were monitored
at 340 nm. All results were obtained using a
Spectramax M2 plate reader.
Natural phosphatase substrate
screeningScreens for activity towards naturally
occurring phosphatase substrates were performed
as previously described (58) using Ceg4[1-193].
Briefly, two 96-well plates were prepared, one as
a blank control and a second for the protein. 10
µL of natural phosphatase substrate at 4 mM (see
table S1 for list) was added to each assay well,
followed by 150 µL of reaction mixture or
reaction mixture containing 2 ug protein to give a
final concentration of 50 mM HEPES-K (pH 7.5),
5 mM MgCl2 1 mM MnCl2, and 0.5 mM NiCl2.
Plates were incubated at 37 °C for 30 mins. After
incubation, 40 µl malachite green development
reagent was added to each well prior to reading
the absorbance at 630 nm. (mM phosphate.min-
1.mg-1) was calculated based on a KH2PO4
standard curve.
Ceg4 metal and pH dependence assays
and kinetics using p-nitrophenylphosphate -
Assays for metal requirements were conducted
using 20 mM p-nitrophenylphosphate (pNPP) in
50 mM HEPES-K (pH 7.0), 0.1 µg.mL-1 of
purified Ceg4 [1-193] and the concentration of
metals indicated, in a final volume of 200 µL. For
pH optimizations, reactions were conducted
using 20 mM pNPP, 15 mM MgCl2, and 0.1
µg.mL-1 Ceg4 [1-193], and one of the following
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
10
buffers at 50 mM; MES (pH 6.5)/HEPES (pH
7)/HEPES (pH 7.5)/HEPES (pH 8)/CHES (pH
9)/CHES (pH 9.5)/CHES (pH 10). For kinetics
determinations, pNPP was used at the
concentrations specified. Reactions were started
by the addition of enzyme and monitored using a
Spectramax M2 plate reader at 405 nm for 25
minutes at 25 °C. Data analysis and curve fitting
was performed using Graphpad Prism 6
Phosphopeptide phosphatase assay
Phosphatase activity towards a library of 53
phosphopeptide sequences was tested as
previously described (46). Briefly, 10 µL peptide
solution was mixed with 150 µL of 50 mM
HEPES (pH 7.0), 15 mM MgCl2 and 0.01 µg of
SBP-purified Ceg4[1-193]. The reaction was
incubated for 10 minutes at 25 °C before addition
of 40 µL of malachite green development reagent.
Absorbance at 630 nM was recorded. Enzyme
velocity (in mM phosphate.min-1.mg-1) was
calculated based on a KH2PO4 standard curve.
Subsequent kinetics determinations with peptide
QMTGpYVSTR were performed using the
method above, with the peptide concentrations
specified in the figure. Data analysis and curve
fitting was performed using Graphpad Prism 6.
Crystallization, structure determination
and analysisCrystals of selenomethionine-
substituted Ceg4[1-208]TEV site were grown at 23
°C using the hanging-drop vapor diffusion
method by mixing 2 μL of 102 mg/mL protein
with 2 μL of reservoir solution containing 0.2 M
MgCl2, 0.5 mM MnCl2, 0.1 M Tris pH 7.3, 30%
(w/v) PEG 4K and 2 mM phosphotyrosine.
Crystals of native Ceg4[1-208]HOG1p were grown
at 23 C using hanging-drop vapor diffusion by
mixing 2 μL of 50 mg/mL protein with 2 μL of
reservoir solution containing 0.2 M NaCl, 0.1 M
Bis-Tris pH 6.5 and 25% (w/v) PEG 3350. All
crystals were cryo-protected with reservoir
solution supplemented with paratone oil.
Diffraction data were collected at 100 K
at beamline 19-ID at Structural Biology Center,
Advanced Photon Source at the wavelength of
0.979 Å (selenium peak). All diffraction data
were reduced with HKL-3000 (71). The Ceg4[1-
208]TEV site structure was solved first by SAD
phasing using PHENIX.solve (72) which
identified all five selenomethionine residues in
the asymmetric unit, followed by model building
by PHENIX.autobuild. The structure of Ceg4[1-
208]HOG1p was determined by Molecular
Replacement using the Ceg4[1-208]TEV site as
search model using PHENIX.phaser. All
structures were refined using PHENIX.refine and
Coot (73). The final Ceg4[1-208]TEV site model
includes the sequence GRQNLYFQG from the
TEV site followed by residues 1-204 of Ceg4; the
final Ceg4[1-208]HOG1p model includes the
sequence TGYVST from yeast Hog1 followed by
residues 1-204 of Ceg4, with residues 146 and
147 unmodeled due to poor electron density. B-
factors were refined as isotropic for all structures.
All geometries were verified with
PHENIX.refine and the wwPDB Validation
server. Structure coordinates were deposited to
the Protein Databank under accession codes
6AOK and 6AOJ for the Ceg4[1-208]TEV site and
Ceg4[1-208]HOG1p structures, respectively.
Structural orthologs were identified using the
Dali lite server (74). Active site volume was
calculated by the CastP (75)
Yeast transformation and MAPK
pathway activation assayCeg4[1-397] or the
mutants specified were cloned into pYES2 NT/A,
transformed into S. cerevisiae (W303 MATa,
bar1::NatR, far1Δ, mfa2::pFus1-GFP, ura3::Kan-
pSTL1-BFP, his3, trp1, leu2) using the LiAc
procedure (76) and grown on selective media for
two days at 30 °C. For analysis of mating and
high osmolarity glycerol (HOG) pathway
responses by flow cytometry, transformants were
grown either in duplicate or triplicate in selective
medium overnight at 30 °C. Empty plasmid was
grown as a negative control. Overnight cultures
were diluted to OD600 between 0.1-0.2 and
grown to early log phase. For analysis of mating
pathway response, yeast cells were treated with 1
mM α-factor and incubated at 30 °C for two
hours. For analysis of the HOG pathway
response, cells were treated with 2 mM KCl and
incubated at 30 °C for one hour. For both
conditions, cells were then treated with the
protein synthesis inhibitor cycloheximide for 30
minutes. For each sample, 10,000 cells were
measured with a MACSQuant Vyb (Miltenyi
Biotech). Data shown are mean fluorescence
(GFP for mating response and BFP for HOG
response) and standard deviation of duplicates or
triplicates.
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
11
HEK293t and HeLa cell culture and
transfectionFor MAPK activation assays,
HEK293t cells were maintained in Dulbecco’s
Modified Eagle’s Medium (DMEM),
supplemented with 10 % FBS at 37 °C and 5 %
CO2, and grown to a confluence of approximately
70 % at the time of transfection. Cells were
transfected using Lipofectamine 3000 (Life
Technologies) according to the manufacturer’s
instructions, with endotoxin free pcDNA3-N-
Flag-LIC containing Ceg4[1-397] or Ceg4[1-
397]D9A.
For cell localization studies, HeLa cells
were grown on poly-L-lysine treated glass
coverslips, maintained in DMEM, supplemented
with 10 % FBS at 37 °C and 5 % CO2, and grown
to a confluence of approximately 70 % at time of
transfection. Cells were transfected with either
pEGFP-N1 containing Ceg4[1-397], Ceg4[1-
397]D9A, Ceg4[1-207] or Ceg4[208-397] using
Lipofectamine 3000 (Life Technologies)
according to the manufacturer’s instructions
followed by counterstaining with DAPI during
fixation. For co-localization studies, pEGFP-N1
Ceg4[1-397]D9A transfected cells were grown in
chambered coverslips and incubated with ER-
tracker and Lyso-tracker red dyes according to
the manufacturer’s instructions prior to fixation
and counterstaining with DAPI. Microscopy data
were collected using a Nikon TiE inverted
microscope and Nikon C2 confocal system, with
a 60X oil immersion lens.
MAPK activation and Immunoblotting
24 hours post-transfection with Ceg4[1-397] or
Ceg4[1-397]D9A, 5 µg/ml anisomycin or 200 nM
TPA was added to the culture medium. After 30
minutes, cells were gently washed once with PBS
followed by lysis directly into SDS-PAGE
loading buffer. SDS-PAGE was performed with
the addition of 0.5% 2,2,2-trichloroethanol added
to the resolving portion of the gel. After
electrophoresis, gels were exposed to UV light
for 2 minutes and imaged with a GelDoc
(BioRad) to obtain loading controls prior to
western blotting. Proteins were transferred to
nitrocellulose using a Transblot Turbo (BioRad).
Membranes were blocked with blocking buffer
(5% (w/v) BSA in TBS with 0.1% Tween 20) for
1 hour, followed by incubation with either anti-
p38 (Cell Signalling Technology, #8690, 1:1000
dilution in blocking buffer) or anti phospho-p38
(Cell Signaling Technology, #4511, 1:1000
dilution in blocking buffer) overnight. After
washing, membranes were incubated in 5% non-
fat skim milk in TBS with 0.1% Tween 20
containing anti-rabbit HRP (Cell Signalling
Technology, #7074, 1:4000 dilution) for 1 hour.
Blots were developed with BioRad Clarity
Western plus reagent and imaged using a GelDoc
(BioRad) and visualized using ImageLab
(BioRad).
Visualization of total cellular tyrosine
phosphorylation by immunoblot in S. cerevisiae
and HEK293T.. BY4741 S. cerevisiae (MATa
his3Δ1 leu2Δ0 lys2Δ0 ura3Δ0 (77)) were
transformed with the indicated constructs in
pYES2NT/A using the LiAc procedure (76) and
grown on CM agar -uracil selective media for two
days at 30 °C. Colonies were picked and cultured
overnight in selective media supplemented with
2% (w/v) raffinose. In the morning, 3ODs of cells
were harvested by centrifugation, washed and
resuspended in 5 ml selective media
supplemented with 2 % (w/v) galactose and
incubated at 30 °C with shaking for 5 hrs.
Cultures were harvested by centrifugation and
cells were lysed using an alkaline/SDS lysis
procedure (78). HEK293T cells were cultured in
and transfected in 6 well plates as per procedures
described for MAPK activation assays. Cells
were washed with PBS and harvested directly
using 100 µl/well of SDS-PAGE loading buffer,
followed by brief sonication to reduce viscosity
and 5 minutes of heating at 95 °C. 20 µl of each
yeast lysate and 15 µl was loaded and separated
with 12 % SDS-PAGE gels supplemented with
0.5 % (v/v) 2,2,2-trichloro ethanol, followed
visualization of protein loading after exposure of
gels to UV light for 2-3 minutes. Protein was
transferred to nitrocellulose for immunoblotting
using a Transblot Turbo (BioRad). For total
phosphotyrosine detection, yeast and HEK293T
blots were blocked with 5% w/v BSA, 1X TBS,
0.1% Tween 20 at 4°C for 1hr with gentle shaking
followed by overnight incubation with α-P-Tyr-
100 antibody (Cell Signalling Technology,
#9411, 1:2000 in blocking buffer). For Ceg4
detection in yeast, blocking was performed in 5%
non-fat skim milk in TBS with 0.1% Tween 20
followed by incubation with α-Xpress
(Thermofisher Scientific, #R910-25 1:4000)
overnight. For detection of Ceg4 in HEK293T
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
12
cells, blocking was performed in 3% non-fat skim
milk in TBS followed by incubation with α‐
FLAG M2 (1:2000, Cat# F1804, Sigma‐Aldrich)
for 45 minutes at room temperature.
After 3, 5 minute washes with TBS with
0.1% Tween 20, all blots were incubated in 5%
non-fat skim milk in TBS with 0.1% Tween 20
containing anti-mouse HRP (Cell Signalling
Technology, #7076, 1:4000 dilution) for 1 hour.
Blots were developed with BioRad Clarity
Western plus reagent and imaged using a GelDoc
(BioRad) and visualized using ImageLab
(BioRad).
Acknowledgements: Special thanks to Greg Brown, for his invaluable assistance with screening and
activity assays. We would also like to thank the Claycomb lab for training and use of their microscopy
facilities.
Conflict of interest: The authors declare no conflicts of interest.
Author contributions: AQ performed human in vivo work with the help of DV. EE and OE purified and
crystalized Ceg4. OE and AQ performed enzyme activity screening. PK and SP performed yeast MAPK
assays. BN collected crystal data. PS solved the structures and wrote the manuscript. AQ and AS wrote the
manuscript with input from AWE and AS. All work was performed under the supervision of AWE, AFY
and AS.
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
13
REFERENCES
1. Viboud, G. I., and Bliska, J. B. (2005) Yersinia outer proteins: role in modulation of host
cell signaling responses and pathogenesis. Annu Rev Microbiol 59, 69-89
2. Watarai, M., Tobe, T., Yoshikawa, M., and Sasakawa, C. (1995) Contact of Shigella with
host cells triggers release of Ipa invasins and is an essential function of invasiveness. The
EMBO journal 14, 2461-2470
3. Zierler, M. K., and Galan, J. E. (1995) Contact with cultured epithelial cells stimulates
secretion of Salmonella typhimurium invasion protein InvJ. Infect Immun 63, 4024-4028
4. Zhang, L., Mei, M., Yu, C., Shen, W., Ma, L., He, J., and Yi, L. (2016) The Functions of
Effector Proteins in Yersinia Virulence. Pol J Microbiol 65, 5-12
5. Parsot, C. (2009) Shigella type III secretion effectors: how, where, when, for what
purposes? Curr Opin Microbiol 12, 110-116
6. Huang, L., Boyd, D., Amyot, W. M., Hempstead, A. D., Luo, Z. Q., O'Connor, T. J., Chen,
C., Machner, M., Montminy, T., and Isberg, R. R. (2011) The E Block motif is associated
with Legionella pneumophila translocated substrates. Cell Microbiol 13, 227-245
7. Zhu, W., Banga, S., Tan, Y., Zheng, C., Stephenson, R., Gately, J., and Luo, Z. Q. (2011)
Comprehensive identification of protein substrates of the Dot/Icm type IV transporter of
Legionella pneumophila. PLoS One 6, e17638
8. Burstein, D., Zusman, T., Degtyar, E., Viner, R., Segal, G., and Pupko, T. (2009) Genome-
scale identification of Legionella pneumophila effectors using a machine learning
approach. PLoS Pathog 5, e1000508
9. de Felipe, K. S., Pampou, S., Jovanovic, O. S., Pericone, C. D., Ye, S. F., Kalachikov, S.,
and Shuman, H. A. (2005) Evidence for acquisition of Legionella type IV secretion
substrates via interdomain horizontal gene transfer. J Bacteriol 187, 7716-7726
10. Kubori, T., Hyakutake, A., and Nagai, H. (2008) Legionella translocates an E3 ubiquitin
ligase that has multiple U-boxes with distinct functions. Mol Microbiol 67, 1307-1319
11. Luo, Z. Q., and Isberg, R. R. (2004) Multiple substrates of the Legionella pneumophila
Dot/Icm system identified by interbacterial protein transfer. Proceedings of the National
Academy of Sciences of the United States of America 101, 841-846
12. Ensminger, A. W. (2016) Legionella pneumophila, armed to the hilt: justifying the largest
arsenal of effectors in the bacterial world. Curr Opin Microbiol 29, 74-80
13. Chien, M., Morozova, I., Shi, S., Sheng, H., Chen, J., Gomez, S. M., Asamani, G., Hill, K.,
Nuara, J., Feder, M., Rineer, J., J, G. J., Steshenko, V., Park, S. H., and Geringer-Sameth,
A. (2004) The genomic sequence of the accidental pathogen Legionella pneumophila.
Science 305, 1966-1968
14. Horwitz, M. A. (1983) The Legionnaires' disease bacterium (Legionella pneumophila)
inhibits phagosome-lysosome fusion in human monocytes. J Exp Med 158, 2108-2126
15. Horwitz, M. A. (1983) Formation of a novel phagosome by the Legionnaires' disease
bacterium (Legionella pneumophila) in human monocytes. J Exp Med 158, 1319-1331
16. O'Connor, T. J., Adepoju, Y., Boyd, D., and Isberg, R. R. (2011) Minimization of the
Legionella pneumophila genome reveals chromosomal regions involved in host range
expansion. Proc Natl Acad Sci U S A 108, 14733-14740
17. Gruenheid, S., DeVinney, R., Bladt, F., Goosney, D., Gelkop, S., Gish, G. D., Pawson, T.,
and Finlay, B. B. (2001) Enteropathogenic E. coli Tir binds Nck to initiate actin pedestal
formation in host cells. Nat Cell Biol 3, 856-859
18. Michard, C., Sperandio, D., Bailo, N., Pizarro-Cerda, J., LeClaire, L., Chadeau-Argaud,
E., Pombo-Gregoire, I., Hervet, E., Vianney, A., Gilbert, C., Faure, M., Cossart, P., and
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
14
Doublet, P. (2015) The Legionella Kinase LegK2 Targets the ARP2/3 Complex To Inhibit
Actin Nucleation on Phagosomes and Allow Bacterial Evasion of the Late Endocytic
Pathway. mBio 6, e00354-00315
19. Nagai, H., Kagan, J. C., Zhu, X., Kahn, R. A., and Roy, C. R. (2002) A bacterial guanine
nucleotide exchange factor activates ARF on Legionella phagosomes. Science 295, 679-
682
20. Hernandez, L. D., Hueffer, K., Wenk, M. R., and Galan, J. E. (2004) Salmonella modulates
vesicular traffic by altering phosphoinositide metabolism. Science 304, 1805-1807
21. Mittal, R., Peak-Chew, S. Y., and McMahon, H. T. (2006) Acetylation of MEK2 and I
kappa B kinase (IKK) activation loop residues by YopJ inhibits signaling. Proc Natl Acad
Sci U S A 103, 18574-18579
22. Selyunin, A. S., Sutton, S. E., Weigele, B. A., Reddick, L. E., Orchard, R. C., Bresson, S.
M., Tomchick, D. R., and Alto, N. M. (2011) The assembly of a GTPase-kinase signalling
complex by a bacterial catalytic scaffold. Nature 469, 107-111
23. Cui, J., Yao, Q., Li, S., Ding, X., Lu, Q., Mao, H., Liu, L., Zheng, N., Chen, S., and Shao,
F. (2010) Glutamine deamidation and dysfunction of ubiquitin/NEDD8 induced by a
bacterial effector family. Science 329, 1215-1218
24. Ashida, H., Kim, M., Schmidt-Supprian, M., Ma, A., Ogawa, M., and Sasakawa, C. (2010)
A bacterial E3 ubiquitin ligase IpaH9.8 targets NEMO/IKKgamma to dampen the host NF-
kappaB-mediated inflammatory response. Nat Cell Biol 12, 66-73; sup pp 61-69
25. Ashida, H., and Sasakawa, C. (2017) Bacterial E3 ligase effectors exploit host ubiquitin
systems. Curr Opin Microbiol 35, 16-22
26. Lin, D. Y., Diao, J., and Chen, J. (2012) Crystal structures of two bacterial HECT-like E3
ligases in complex with a human E2 reveal atomic details of pathogen-host interactions.
Proc Natl Acad Sci U S A 109, 1925-1930
27. Wu, B., Skarina, T., Yee, A., Jobin, M. C., Dileo, R., Semesi, A., Fares, C., Lemak, A.,
Coombes, B. K., Arrowsmith, C. H., Singer, A. U., and Savchenko, A. (2010) NleG Type
3 effectors from enterohaemorrhagic Escherichia coli are U-Box E3 ubiquitin ligases. PLoS
Pathog 6, e1000960
28. Singer, A. U., Rohde, J. R., Lam, R., Skarina, T., Kagan, O., Dileo, R., Chirgadze, N. Y.,
Cuff, M. E., Joachimiak, A., Tyers, M., Sansonetti, P. J., Parsot, C., and Savchenko, A.
(2008) Structure of the Shigella T3SS effector IpaH defines a new class of E3 ubiquitin
ligases. Nature structural & molecular biology 15, 1293-1301
29. Quezada, C. M., Hicks, S. W., Galan, J. E., and Stebbins, C. E. (2009) A family of
Salmonella virulence factors functions as a distinct class of autoregulated E3 ubiquitin
ligases. Proc Natl Acad Sci U S A 106, 4864-4869
30. Quaile, A. T., Urbanus, M. L., Stogios, P. J., Nocek, B., Skarina, T., Ensminger, A. W.,
and Savchenko, A. (2015) Molecular Characterization of LubX: Functional Divergence of
the U-Box Fold by Legionella pneumophila. Structure 23, 1459-1469
31. Ensminger, A. W., and Isberg, R. R. (2010) E3 ubiquitin ligase activity and targeting of
BAT3 by multiple Legionella pneumophila translocated substrates. Infect Immun 78, 3905-
3919
32. Khoury, G. A., Baliban, R. C., and Floudas, C. A. (2011) Proteome-wide post-translational
modification statistics: frequency analysis and curation of the swiss-prot database. Sci Rep
1
33. Cohen, P. (2002) The origins of protein phosphorylation. Nat Cell Biol 4, E127-130
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
15
34. Manning, G., Plowman, G. D., Hunter, T., and Sudarsanam, S. (2002) Evolution of protein
kinase signaling from yeast to man. Trends Biochem Sci 27, 514-520
35. Seger, R., and Krebs, E. G. (1995) The MAPK signaling cascade. FASEB J 9, 726-735
36. Cargnello, M., and Roux, P. P. (2011) Activation and function of the MAPKs and their
substrates, the MAPK-activated protein kinases. Microbiol Mol Biol Rev 75, 50-83
37. N'Guessan, P. D., Etouem, M. O., Schmeck, B., Hocke, A. C., Scharf, S., Vardarova, K.,
Opitz, B., Flieger, A., Suttorp, N., and Hippenstiel, S. (2007) Legionella pneumophila-
induced PKCalpha-, MAPK-, and NF-kappaB-dependent COX-2 expression in human
lung epithelium. Am J Physiol Lung Cell Mol Physiol 292, L267-277
38. Shin, S., Case, C. L., Archer, K. A., Nogueira, C. V., Kobayashi, K. S., Flavell, R. A., Roy,
C. R., and Zamboni, D. S. (2008) Type IV secretion-dependent activation of host MAP
kinases induces an increased proinflammatory cytokine response to Legionella
pneumophila. PLoS Pathog 4, e1000220
39. Welsh, C. T., Summersgill, J. T., and Miller, R. D. (2004) Increases in c-Jun N-terminal
kinase/stress-activated protein kinase and p38 activity in monocyte-derived macrophages
following the uptake of Legionella pneumophila. Infect Immun 72, 1512-1518
40. Haenssler, E., and Isberg, R. R. (2011) Control of host cell phosphorylation by legionella
pneumophila. Front Microbiol 2, 64
41. Zhang, Z. Y., Clemens, J. C., Schubert, H. L., Stuckey, J. A., Fischer, M. W., Hume, D.
M., Saper, M. A., and Dixon, J. E. (1992) Expression, purification, and physicochemical
characterization of a recombinant Yersinia protein tyrosine phosphatase. The Journal of
biological chemistry 267, 23759-23766
42. Li, H., Xu, H., Zhou, Y., Zhang, J., Long, C., Li, S., Chen, S., Zhou, J. M., and Shao, F.
(2007) The phosphothreonine lyase activity of a bacterial type III effector family. Science
315, 1000-1003
43. Lifshitz, Z., Burstein, D., Schwartz, K., Shuman, H. A., Pupko, T., and Segal, G. (2014)
Identification of novel Coxiella burnetii Icm/Dot effectors and genetic analysis of their
involvement in modulating a mitogen-activated protein kinase pathway. Infect Immun 82,
3740-3752
44. Chen, R. E., and Thorner, J. (2007) Function and regulation in MAPK signaling pathways:
lessons learned from the yeast Saccharomyces cerevisiae. Biochim Biophys Acta 1773,
1311-1340
45. Koonin, E. V., and Tatusov, R. L. (1994) Computer analysis of bacterial haloacid
dehalogenases defines a large superfamily of hydrolases with diverse specificity.
Application of an iterative approach to database search. Journal of molecular biology 244,
125-132
46. Kuznetsova, E., Nocek, B., Brown, G., Makarova, K. S., Flick, R., Wolf, Y. I.,
Khusnutdinova, A., Evdokimova, E., Jin, K., Tan, K., Hanson, A. D., Hasnain, G., Zallot,
R., de Crecy-Lagard, V., Babu, M., Savchenko, A., Joachimiak, A., Edwards, A. M.,
Koonin, E. V., and Yakunin, A. F. (2015) Functional Diversity of Haloacid Dehalogenase
Superfamily Phosphatases from Saccharomyces cerevisiae: BIOCHEMICAL,
STRUCTURAL, AND EVOLUTIONARY INSIGHTS. The Journal of biological
chemistry 290, 18678-18698
47. Gohla, A., Birkenfeld, J., and Bokoch, G. M. (2005) Chronophin, a novel HAD-type serine
protein phosphatase, regulates cofilin-dependent actin dynamics. Nat Cell Biol 7, 21-29
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
16
48. Rayapureddi, J. P., Kattamuri, C., Steinmetz, B. D., Frankfort, B. J., Ostrin, E. J., Mardon,
G., and Hegde, R. S. (2003) Eyes absent represents a class of protein tyrosine phosphatases.
Nature 426, 295-298
49. Seifried, A., Knobloch, G., Duraphe, P. S., Segerer, G., Manhard, J., Schindelin, H.,
Schultz, J., and Gohla, A. (2014) Evolutionary and structural analyses of mammalian
haloacid dehalogenase-type phosphatases AUM and chronophin provide insight into the
basis of their different substrate specificities. The Journal of biological chemistry 289,
3416-3431
50. Kim, Y., Gentry, M. S., Harris, T. E., Wiley, S. E., Lawrence, J. C., Jr., and Dixon, J. E.
(2007) A conserved phosphatase cascade that regulates nuclear membrane biogenesis. Proc
Natl Acad Sci U S A 104, 6596-6601
51. Seifried, A., Schultz, J., and Gohla, A. (2013) Human HAD phosphatases: structure,
mechanism, and roles in health and disease. The FEBS journal 280, 549-571
52. Rinaldo-Matthis, A., Rampazzo, C., Reichard, P., Bianchi, V., and Nordlund, P. (2002)
Crystal structure of a human mitochondrial deoxyribonucleotidase. Nature structural
biology 9, 779-787
53. Lahiri, S. D., Zhang, G., Dunaway-Mariano, D., and Allen, K. N. (2006) Diversification of
function in the haloacid dehalogenase enzyme superfamily: The role of the cap domain in
hydrolytic phosphoruscarbon bond cleavage. Bioorg Chem 34, 394-409
54. Kall, L., Krogh, A., and Sonnhammer, E. L. (2004) A combined transmembrane topology
and signal peptide prediction method. Journal of molecular biology 338, 1027-1036
55. Burstein, D., Amaro, F., Zusman, T., Lifshitz, Z., Cohen, O., Gilbert, J. A., Pupko, T.,
Shuman, H. A., and Segal, G. (2016) Genomic analysis of 38 Legionella species identifies
large and diverse effector repertoires. Nat Genet 48, 167-175
56. Finn, R. D., Coggill, P., Eberhardt, R. Y., Eddy, S. R., Mistry, J., Mitchell, A. L., Potter,
S. C., Punta, M., Qureshi, M., Sangrador-Vegas, A., Salazar, G. A., Tate, J., and Bateman,
A. (2016) The Pfam protein families database: towards a more sustainable future. Nucleic
acids research 44, D279-285
57. Zhu, Y., Hu, L., Zhou, Y., Yao, Q., Liu, L., and Shao, F. (2010) Structural mechanism of
host Rab1 activation by the bifunctional Legionella type IV effector SidM/DrrA. Proc Natl
Acad Sci U S A 107, 4699-4704
58. Proudfoot, M., Kuznetsova, E., Sanders, S. A., Gonzalez, C. F., Brown, G., Edwards, A.
M., Arrowsmith, C. H., and Yakunin, A. F. (2008) High throughput screening of purified
proteins for enzymatic activity. Methods Mol Biol 426, 331-341
59. Urbanus, M. L., Quaile, A. T., Stogios, P. J., Morar, M., Rao, C., Di Leo, R., Evdokimova,
E., Lam, M., Oatway, C., Cuff, M. E., Osipiuk, J., Michalska, K., Nocek, B. P., Taipale,
M., Savchenko, A., and Ensminger, A. W. (2016) Diverse mechanisms of metaeffector
activity in an intracellular bacterial pathogen, Legionella pneumophila. Mol Syst Biol 12,
893
60. Wallden, K., Stenmark, P., Nyman, T., Flodin, S., Graslund, S., Loppnau, P., Bianchi, V.,
and Nordlund, P. (2007) Crystal structure of human cytosolic 5'-nucleotidase II: insights
into allosteric regulation and substrate recognition. The Journal of biological chemistry
282, 17828-17836
61. Jung, S. K., Jeong, D. G., Chung, S. J., Kim, J. H., Park, B. C., Tonks, N. K., Ryu, S. E.,
and Kim, S. J. (2010) Crystal structure of ED-Eya2: insight into dual roles as a protein
tyrosine phosphatase and a transcription factor. FASEB J 24, 560-569
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
17
62. Peisach, E., Selengut, J. D., Dunaway-Mariano, D., and Allen, K. N. (2004) X-ray crystal
structure of the hypothetical phosphotyrosine phosphatase MDP-1 of the haloacid
dehalogenase superfamily. Biochemistry 43, 12770-12779
63. Fortpied, J., Maliekal, P., Vertommen, D., and Van Schaftingen, E. (2006) Magnesium-
dependent phosphatase-1 is a protein-fructosamine-6-phosphatase potentially involved in
glycation repair. The Journal of biological chemistry 281, 18378-18385
64. Brewster, J. L., de Valoir, T., Dwyer, N. D., Winter, E., and Gustin, M. C. (1993) An
osmosensing signal transduction pathway in yeast. Science 259, 1760-1763
65. Fontana, M. F., Shin, S., and Vance, R. E. (2012) Activation of host mitogen-activated
protein kinases by secreted Legionella pneumophila effectors that inhibit host protein
translation. Infect Immun 80, 3570-3575
66. Biswas, T., Yi, L., Aggarwal, P., Wu, J., Rubin, J. R., Stuckey, J. A., Woodard, R. W., and
Tsodikov, O. V. (2009) The tail of KdsC: conformational changes control the activity of a
haloacid dehalogenase superfamily phosphatase. The Journal of biological chemistry 284,
30594-30603
67. Burroughs, A. M., Allen, K. N., Dunaway-Mariano, D., and Aravind, L. (2006)
Evolutionary genomics of the HAD superfamily: understanding the structural adaptations
and catalytic diversity in a superfamily of phosphoesterases and allied enzymes. Journal
of molecular biology 361, 1003-1034
68. Tegtmeyer, N., Neddermann, M., Asche, C. I., and Backert, S. (2017) Subversion of host
kinases: a key network in cellular signaling hijacked by Helicobacter pylori CagA. Mol
Microbiol 105, 358-372
69. Krachler, A. M., Woolery, A. R., and Orth, K. (2011) Manipulation of kinase signaling by
bacterial pathogens. J Cell Biol 195, 1083-1092
70. Weissenmayer, B. A., Prendergast, J. G., Lohan, A. J., and Loftus, B. J. (2011) Sequencing
illustrates the transcriptional response of Legionella pneumophila during infection and
identifies seventy novel small non-coding RNAs. PLoS One 6, e17570
71. Minor, W., Cymborowski, M., Otwinowski, Z., and Chruszcz, M. (2006) HKL-3000: the
integration of data reduction and structure solution--from diffraction images to an initial
model in minutes. Acta crystallographica. Section D, Biological crystallography 62, 859-
866
72. Adams, P. D., Afonine, P. V., Bunkoczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd,
J. J., Hung, L. W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W.,
Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C., and Zwart,
P. H. (2010) PHENIX: a comprehensive Python-based system for macromolecular
structure solution. Acta crystallographica. Section D, Biological crystallography 66, 213-
221
73. Emsley, P., and Cowtan, K. (2004) Coot: model-building tools for molecular graphics. Acta
crystallographica. Section D, Biological crystallography 60, 2126-2132
74. Hasegawa, H., and Holm, L. (2009) Advances and pitfalls of protein structural alignment.
Curr Opin Struct Biol 19, 341-348
75. Dundas, J., Ouyang, Z., Tseng, J., Binkowski, A., Turpaz, Y., and Liang, J. (2006) CASTp:
computed atlas of surface topography of proteins with structural and topographical
mapping of functionally annotated residues. Nucleic acids research 34, W116-118
76. Gietz, R. D., and Schiestl, R. H. (2007) High-efficiency yeast transformation using the
LiAc/SS carrier DNA/PEG method. Nature protocols 2, 31-34
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
18
77. Brachmann, C. B., Davies, A., Cost, G. J., Caputo, E., Li, J., Hieter, P., and Boeke, J. D.
(1998) Designer deletion strains derived from Saccharomyces cerevisiae S288C: a useful
set of strains and plasmids for PCR-mediated gene disruption and other applications. Yeast
14, 115-132
78. von der Haar, T. (2007) Optimized protein extraction for quantitative proteomics of yeasts.
PLoS One 2, e1078
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
19
FOOTNOTES
Results shown in this report are supported in part by a National Institutes of Health grant (GM094585) to
AS through the Midwest Center for Structural Genomics and by the U. S. Department of Energy, Office of
Biological and Environmental Research (contract DEAC0206CH11357). The work performed at Argonne
National Laboratory, Structural Biology Center at the Advanced Photon Source. Argonne is operated by
UChicago Argonne, LLC, for the U.S. Department of Energy, Office of Biological and Environmental
Research under contract DE-AC02-06CH11357
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
20
TABLES
Table 1 X-ray diffraction data collection and refinement statistics
Structure
Ceg4[1-208]TEV site
Se-Met
Ceg4[1-208]HOG1p
Native
PDB Code
6AOK
6AOJ
Data collection
Space group
P212121
C2
Cell dimensions
a, b, c (Å)
β (°)
40.04, 48.03, 100.10
90.0
78.97, 75.92, 47.91
104.2
Resolution, Å
40.0 1.88
50.0 1.90
Rmergea
Rpim
0.071 (0.525)b
0.035 (0.266)
0.052 (0.259)
0.031 (0.149)
I / (I)
21.4 (3.9)
39.3 (4.5)
Completeness, %
100 (99.9)
97.2 (92.5)
Redundancy
5.1 (4.9)
3.9 (3.8)
Refinement
Resolution, Å
37.17 1.88
36.53 1.90
No. of unique reflections:
working, test
16264, 1461
20876, 1044
R-factor/free R-factorc
15.6/19.8 (19.9/27.1)
16.2/19.3 (22.1/28.3)
No. of refined atoms
Protein
Magnesium
Chloride
Water
1811
1
2
269
1738
1
1
252
B-factors
Protein
Magnesium
Chloride
Water
23.1
12.7
23.5
37.8
51.4
32.7
38.3
59.3
r.m.s.d.
Bond lengths, Å
Bond angles,
0.016
1.267
0.009
0.954
aRsym = hi|Ii(h) - I(h)/hiIi(h), where Ii(h) and I(h) are the ith and mean measurement of the intensity
of reflection h.
bFigures in parentheses indicate the values for the outer shells of the data.
cR = |Fpobs Fpcalc|/Fpobs, where Fpobs and Fpcalc are the observed and calculated structure factor amplitudes,
respectively.
* = molecules in the active site cleft.
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
21
FIGURE LEGENDS
FIGURE 1. Domain organization and species Ceg4 orthologues. Phylogenetic analysis of the HAD
domains of Legionella Dot/ICM effectors shows that Ceg4 (Lpg0096) belongs to a group of effectors in
which the majority of members possess a two transmembrane (TM) region. Numerals indicate node
posterior probabilities, scale bar illustrates substitutions per site.
FIGURE 2. Ceg4 is a phosphotyrosine phosphatase. A, General screening for a range of activities
indicated that Ceg4 is a phosphatase. B, Kinetics were obtained for the generic phosphatase substrate pNPP.
Error bars denote SEM values. C, Incubation of Ceg4 with a variety of naturally occurring phosphatase
substrates, and subsequent detection of released phosphates, reveals that Ceg4 preferentially removes the
phosphate from phosphotyrosine in vitro. See also supplemental data, Figure S1E for the full list of tested
substrates. Error bars show 95% confidence intervals from triplicate data. D, Kinetics were obtained for the
generic phosphatase substrate pNPP. Error bars denote SEM values E, Ceg4 dephosphorylates all tested of
phosphotyrosine-containing peptides at levels 4-5 times higher than phosphoserine or phosphothreonine-
containing peptides in vitro. Error bars denote 95% confidence intervals from triplicate data. See also
Supplemental Figure S1E for all of tested peptides.
FIGURE 3. Crystal structure of Ceg4’s HAD-like domain provides molecular insight into phosphatase
activity. A, Schematic of the final structural model of Ceg4. Amino acids highlighted in blue belong to the
core Rossmann-like fold, gray to the cap domain and orange to the expression tag used for purification. B,
Overall structure of the Ceg4 HAD-like domain, highlighting positions and topology of the C1-cap
subdomain and the ‘squiggle motif’.C, Close-up of the Ceg4 HAD-like domain active site, indicating
positions of catalytic site residues (small red spheres denote positions of water). D, Catalytic site residues
of Ceg4 form a ~350 Å pocket, occupied by a magnesium and a chloride ion and partially covered by N20
of the cap subdomain.
FIGURE 4. Ceg4[1-208] crystallizes with the N-terminal fusion tag’s tyrosine bound deeply in the catalytic
pocket of the adjacent molecule. A, Crystal packing of Ceg4[1-208]HOG1p shows the N-terminal peptide
packed against the active site of the adjacent Ceg4[1-208]HOG1p molecule. B, Close-up examination of the
HOG1p tag-Ceg4 interaction shows the tyrosine bound deeply in the catalytic pocket. C, Hog1 and TEV
fusion tag peptides bind the active site of Ceg4[1-208] with strikingly similar conformations.
FIGURE 5. Ceg4 supresses MAPK responses in S. cerevisiae. A, Ceg4[1-397] is able to reduce the Hog1-
mediated high osmolarity response to high salt concentrations and the Fus3 mediated pheromone/mating
response to α-factor. Removal of the C-terminal TM domain reduced Ceg4’s ability to lessen the pheromone
response, but slightly increased its ability to reduce activation of the hyperosmolarity response, suggesting
off-target effects. B, Ceg4[1-397] mutants K17A, N20A, E26A, E108A, E131A and T161A all retained
wildtype levels of activity, whereas mutants D11A, V23D, F24A, Y43A, K104A, K135, and D158 all
exhibited reduced ability to supress the pheromone response. See also Supplementary Figures S5 for
expression testing. C, Location of tested mutants illustrated on the structure of Ceg4. Mutation of residues
highlighted in red reduced ability to supress the pheromone response.
FIGURE 6. Ceg4 localization to the endoplasmic reticulum requires C-terminal TM domain and dampens
activation of human MAPK p38 in vivo. A, Ceg4 localization is dependent on C-terminal TM domains. N-
terminal GFP fusions of Ceg4[1-397], Ceg4[1-397]D9A and Ceg4[208-397] expressed in HeLa cells show
distinct subcellular localisation that is lost in Ceg4[1-207] lacking transmembrane regions. B, GFP-Ceg4[1-
397]D9A colocalizes with ER-tracker red dye. See also Supplemental figure S6. C, Activation and
phosphorylation of p38 MAPK was achieved by incubating HEK293 cells expressing Ceg4[1-397] or
Ceg4[1-397]D9A with TPA or anisomycin. While total levels of p38 remained unchanged, cells expressing
Ceg4[1-397] showed reduced levels of phospho-p38 compared to D9A mutant constructs.
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
22
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
23
Figure 1
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
24
Figure 2
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
25
Figure 3
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
26
Figure 4
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
27
Figure 5
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Legionella phosphotyrosine phosphatase activity towards MAPKs
28
Figure 6
by guest on January 4, 2018http://www.jbc.org/Downloaded from
Alexander W. Ensminger and Alexei Savchenko
Boguslaw Nocek, Purnima S. Kompella, Sergio Peisajovich, Alexander F. Yakunin,
Andrew T. Quaile, Peter J. Stogios, Olga Egorova, Elena Evdokimova, Dylan Valleau,
attenuates activation of eukaryotic MAPK pathways
The Legionella pneumophila effector Ceg4 is a phosphotyrosine phosphatase that
published online January 4, 2018J. Biol. Chem.
10.1074/jbc.M117.812727Access the most updated version of this article at doi:
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
by guest on January 4, 2018http://www.jbc.org/Downloaded from
... Referring to other bacteria, HAD family protein could mediate metabolic and pathogenic interactions with host cells. As in Legionella pneumophila, Lem4 has been reported as a tyrosine phosphatase belonging to the HAD family that shows strong specificity for phenylphosphate, as well as an effector involved in interactions with Legionella-containing vacuoles or plasma membrane phosphorylated host targets (37). In Porphyromonas gingivalis, a secreted phosphoserine phosphatase SerB653 belong ing to the HAD family was found to be involved in bacteria internalization and persistence in contact with gingival epithelial cells (38). ...
... However, to the best of our knowledge, the mechanism between HAD-like proteins and their interactive proteins is unclear. Only one previous paper showed that the HAD-like protein Ceg4 from Legionella pneumophila can inhibit MAP kinases (Fus3p MAPK) in S. cerevisiae through its C-terminal region (37). This finding provides additional support to our results that C-terminal D15A is associated with MbovP0725 phosphorylation function. ...
... The activity of rMbovP0725 enzymes was measured by monitoring the hydrolysis of pNPP (New England BioLabs) to p-nitrophenyl as described previously (37). Briefly, reactions consisted of 1.77 mM pNPP, 50 mM Tris-HCl (pH 8.0), 10 mM MgCl 2 , and 30 µg rMbovP0725 at 37°C for 0, 0.25, 0.5, 1, 1.5, 2, 4, and 12 h. ...
Article
Full-text available
Mycoplasma species are able to produce and release secreted proteins, such as toxins, adhesins, and virulence-related enzymes, involved in bacteria adhesion, invasion, and immune evasion between the pathogen and host. Here, we investigated a novel secreted protein, MbovP0725, from Mycoplasma bovis encoding a putative haloacid dehalogenase (HAD) hydrolase function of a key serine/threonine phosphatase depending on Mg²⁺ for the dephosphorylation of its substrate pNPP, and it was most active at pH 8 to 9 and temperatures around 40°C. A transposon insertion mutant strain of M. bovis HB0801 that lacked the protein MbovP0725 induced a stronger inflammatory response but with a partial reduction of adhesion ability. Using transcriptome sequencing and quantitative reverse transcription polymerase chain reaction analysis, we found that the mutant was upregulated by the mRNA expression of genes from the glycolysis pathway, while downregulated by the genes enriched in ABC transporters and acetate kinase–phosphate acetyltransferase pathway. Untargeted metabolomics showed that the disruption of the Mbov_0725 gene caused the accumulation of 9-hydroxyoctadecadienoic acids and the consumption of cytidine 5′-monophosphate, uridine monophosphate, and adenosine monophosphate. Both the exogenous and endogenous MbvoP0725 protein created by purification and transfection inhibited lipopolysaccharide (LPS)-induced IL-1β, IL-6, and TNF-α mRNA production and could also attenuate the activation of MAPK-associated pathways after LPS treatment. A pull-down assay identified MAPK p38 and ERK as potential substrates for MbovP0725. These findings define metabolism- and virulence-related roles for a HAD family phosphatase and reveal its ability to inhibit the host pro-inflammatory response. IMPORTANCE Mycoplasma bovis (M. bovis) infection is characterized by chronic pneumonia, otitis, arthritis, and mastitis, among others, and tends to involve the suppression of the immune response via multiple strategies to avoid host cell immune clearance. This study found that MbovP0725, a haloacid dehalogenase (HAD) family phosphatase secreted by M. bovis, had the ability to inhibit the host pro-inflammatory response induced by lipopolysaccharide. Transcriptomic and metabolomic analyses were used to identify MbovP0725 as an important phosphatase involved in glycolysis and nucleotide metabolism. The M. bovis transposon mutant strain T8.66 lacking MbovP0725 induced a higher inflammatory response and exhibited weaker adhesion to host cells. Additionally, T8.66 attenuated the phosphorylation of MAPK P38 and ERK and interacted with the two targets. These results suggested that MbovP0725 had the virulence- and metabolism-related role of a HAD family phosphatase, performing an anti-inflammatory response during M. bovis infection.
... In addition, yeast strains deleted for non-essential MAPK components of different signalling pathways were applied to determine whether yeast MAPK pathway components modulate the growth inhibition effect of bacterial VFs (Furukawa and Hohmann 2013). Stressors and yeast deletion strains have been widely involved in a variety of assays to monitor the modulation of bacterial VFs in MAPK signalling pathways, including agar plate growth phenotypic screening Yang et al. 2019), yeast β-galactosidase assay using the lacZ reporter or fluorescent reporter fused with the MAPK responsive gene (Salomon et al. 2012;Quaile et al. 2018), and immunoblotting with a specific anti-phospho antibody (Kramer et al. 2007;Rohde et al. 2007;Yang et al. 2019). ...
... Ceg4 from L. pneumophila is a phosphotyrosine phosphatase that attenuates the activation of eukaryotic MAPK pathways. Ceg4 has been shown to attenuate the phosphorylation of Hog1 and Fus3 MAP kinases in yeast while attenuating MAPK p38 activation in mammalian cells (Quaile et al. 2018). ...
Article
Full-text available
Pathogenic bacteria employ virulence factors (VF) to establish infection and cause disease in their host. Yeasts, Saccharomyces cerevisiae and Saccharomyces pombe, are useful model organisms to study the functions of bacterial VFs and their interaction with targeted cellular processes because yeast processes and organelle structures are highly conserved and similar to higher eukaryotes. In this review, we describe the principles and applications of the yeast model for the identification and functional characterisation of bacterial VFs to investigate bacterial pathogenesis. The growth inhibition phenotype caused by the heterologous expression of bacterial VFs in yeast is commonly used to identify candidate VFs. Then, subcellular localisation patterns of bacterial VFs can provide further clues about their target molecules and functions during infection. Yeast knockout and overexpression libraries are also used to investigate VF interactions with conserved eukaryotic cell structures (e.g., cytoskeleton and plasma membrane), and cellular processes (e.g., vesicle trafficking, signalling pathways, and programmed cell death). In addition, the yeast growth inhibition phenotype is also useful for screening new drug leads that target and inhibit bacterial VFs. This review provides an updated overview of new tools, principles and applications to study bacterial VFs in yeast.
... Domena Qc może działać jako sensor, który umożliwia kompleksowi IcmR-IcmQ interakcję z krytycznym składnikiem systemu sekrecji, podczas gdy wewnątrzkomórkowy poziom NAD + jest wystarczający do zajścia reakcji wiązania. Niski wewnątrzkomórkowy poziom NAD + prowadzi w tym przypadku do zwiększonego tempa zużycia białka IcmQ, co może nastąpić w czasie wzrostu bakterii, w którym poziom innych białek związanych z wirulencją jest również obniżony [44]. Kompleks IcmR-IcmQ wiąże się z błonami zawierającymi kwaśne fosfolipidy [8]. ...
... Pięć efektorów L. pneumophila wydzielanych za pomocą IV systemu sekrecji hamuje translację białek gospodarza przez aktywację szlaku kinazy białkowej aktywowanej mitogenami (MAPK), modyfikując w ten sposób odpowiedź transkrypcyjną komórki gospodarza [76]. Fosfataza fosfotyrozynowa Ceg4 zawierająca domenę o aktywności dehalogenazy-hydrolazy halo-kwasowej defosforyluje wiele peptydów zawierających fosfotyrozynę i w ten sposób może osłabić aktywację szlaków kontrolowanych przez MAPK zarówno w komórkach drożdży, jak i ludzkich [44]. ...
Article
Full-text available
Abstrakt Bakterie Legionella pneumophila w środowisku naturalnym pasożytują wewnątrz komórek wybranych gatunków pierwotniaków, a po przedostaniu się do sztucznych systemów dystrybucji wody stają się ważnym czynnikiem etiologicznym zapalenia płuc u ludzi. Główną cechą determinującą patogenność tych bakterii jest zdolność do życia i replikacji w makrofagach płucnych, czyli w komórkach wyspecjalizowanych do fagocytozy, zabijania i trawienia mikroorganizmów. Warunkiem wstępnym rozwoju infekcji jest przełamanie mechanizmów bójczych makrofagów i utworzenie wakuoli replikacyjnej LCV ( Legionella containing vacuole). Biogeneza wakuoli LCV jest możliwa dzięki sprawnemu funkcjonowaniu IV systemu sekrecji Dot/Icm, który jest wielobiałkowym, złożonym kompleksem umiejscowionym w wewnętrznej i zewnętrznej membranie osłony komórkowej bakterii. System Dot/Icm liczy 27 elementów, na które składają się m.in. kompleks rdzeniowo-transmembranowy, tworzący strukturalny szkielet całego systemu oraz kompleks białek sprzęgających. Geny kodujące komponenty systemu Dot/Icm są zorganizowane na dwóch regionach chromosomu bak-teryjnego. System sekrecji Dot/Icm umożliwia L. pneumophila wprowadzenie do cytozolu komórki gospodarza ponad 300 białek efektorowych, których skoordynowane działanie powoduje utrzymanie integralności błony wakuoli replikacyjnej oraz pozwala na manipulowanie różnymi procesami komórki. Ważnym elementem strategii wewnątrzkomórkowego namnażania się L. pneumophila jest modulowanie transportu pęcherzykowego, interakcja z retikulum endoplazmatycznym oraz zakłócenie biosyntezy białek, procesów autofagii i apoptozy komórki gospodarza. Poznanie złożonych mechanizmów regulacji i funkcji białek efektorowych systemu Dot/Icm ma decydujące znaczenie w zapobieganiu i leczeniu choroby legionistów.
... localizes to the host ER through its C-terminal transmembrane (TM) domain and exhibits specific phosphatase activity on phosphotyrosine(Quaile et al., 2018). Ceg4 contains a halo acid dehalogenase (HAD)-like domain that is responsible for a variety of activities, including dehalogenase, phosphomutase, and phosphatase activities. ...
Article
Full-text available
Legionella pneumophila is a Gram-negative bacterium ubiquitously present in freshwater environments and causes a serious type of pneumonia called Legionnaires’ disease. During infections, L. pneumophila releases >300 effector proteins into host cells through an Icm/Dot type IV secretion system to manipulate the host defense system for survival within the hosts. Notably, certain effector proteins mediate posttranslational modifications (PTMs), serving as useful approaches exploited by L. pneumophila to modify host proteins. Some effectors catalyze the addition of host protein PTMs, while others mediate the removal of PTMs of host proteins. In this review, we summarize L. pneumophila effector-mediated PTMs of host proteins, including phosphorylation, ubiquitination, glycosylation, AMPylation, phosphocholination, methylation, ADP-ribosylation, as well as dephosphorylation, deubiquitination, deAMPylation, deADP-ribosylation, dephosphocholination, and delipidation. We describe their molecular mechanisms and biological functions in the regulation of bacterial growth and Legionella-containing vacuole biosynthesis and in the disruption of host immune and defense machinery.
... Furthermore, SptP targets a tyrosine kinase Syk and a vesicle fusion protein, N-ethylmalemidesensitive factor to suppress immune responses in mast cells through its phosphatase activity (Choi et al., 2013). Additional phosphatase effectors include WipA, WipB, Ceg4, and Lem4 from L. pneumophila (Beyrakhova et al., 2018;Prevost et al., 2017;Quaile et al., 2018) and Coxiella burnetii effector CinF that dephosphorylates IκBα to inhibit host NF-κB signaling (Zhang et al., 2022). ...
Article
Full-text available
Protein post‐translational modifications (PTMs), such as ADP‐ribosylation and phosphorylation, regulate multiple fundamental biological processes in cells. During bacterial infection, effector proteins are delivered into host cells through dedicated bacterial secretion systems and can modulate important cellular pathways by covalently modifying their host targets. These strategies enable intruding bacteria to subvert various host processes, thereby promoting their own survival and proliferation. Despite rapid expansion of our understanding of effector‐mediated PTMs in host cells, analytical measurements of these molecular events still pose significant challenges in the study of host–pathogen interactions. Nevertheless, with major technical breakthroughs in the last two decades, mass spectrometry (MS) has evolved to be a valuable tool for detecting protein PTMs and mapping modification sites. Additionally, large‐scale PTM profiling, facilitated by different enrichment strategies prior to MS analysis, allows high‐throughput screening of host enzymatic substrates of bacterial effectors. In this review, we summarize the advances in the studies of two representative PTMs (i.e., ADP‐ribosylation and phosphorylation) catalyzed by bacterial effectors during infection. Importantly, we will discuss the ever‐increasing role of MS in understanding these molecular events and how the latest MS‐based tools can aid in future studies of this booming area of pathogenic bacteria–host interactions.
... Furthermore, SptP targets a tyrosine kinase Syk and a vesicle fusion protein, Nethylmalemide-sensitive factor to suppress immune responses in mast cells through its phosphatase activity (Choi et al., 2013). Additional phosphatase effectors include WipA, WipB, Ceg4 and Lem4 from L. pneumophila Prevost et al., 2017;Quaile et al., 2018;Beyrakhova et al., 2018) and Coxiella burnetii effector CinF that dephosphorylates IκBα to inhibit host NF-κB signaling (Zhang et al., 2022). ...
Preprint
Protein post-translational modifications (PTMs), such as ADP-ribosylation and phosphorylation, regulate multiple fundamental biological processes in cells. During bacterial infection, effector proteins are delivered into host cells through dedicated bacterial secretion systems and can modulate important cellular pathways by covalently modifying their host targets. These strategies enable intruding bacteria to subvert various host processes, thereby promoting their own survival and proliferation. Despite rapid expansion of our understanding of effector-mediated PTMs in host cells, analytical measurements of these molecular events still pose significant challenges in the study of host-pathogen interactions. Nevertheless, with major technical breakthroughs in the last two decades, mass spectrometry (MS) has evolved to be a valuable tool for detecting protein PTMs and mapping modification sites. Additionally, large-scale PTM profiling, facilitated by different enrichment strategies prior to MS analysis, allows high-throughput screening of host enzymatic substrates of bacterial effectors. In this review, we summarize the advances in the studies of two representative PTMs (i.e., ADP-ribosylation and phosphorylation) catalyzed by bacterial effectors during infection. Importantly, we will discuss the ever increasing role of MS in understanding these molecular events and how the latest MS-based tools can aid in future studies of this booming area of pathogenic bacteria-host interactions.
... The recently discovered Ceg4, with its haloacid dehalogenase (HAD)-like domain, is a phosphotyrosine phosphatase. In vitro, Ceg4 attenuates the activation of mitogen-activated protein kinases (MAPK) (Quaile et al., 2018). ...
Thesis
Legionella pneumophila is an intracellular bacterium that secretes over 300 proteins in the hostcell through a specialized type 4 secretion system. One of these secreted L. pneumophila effectors, RomA, was shown to directly modify the host chromatin by methylating lysine 14 of Histone H3 (H3K14), a usually acetylated residue. This led to the question how deacetylation of this mark might happen during infection. An in-depth bioinformatics search led to the identification of a protein predicted to code for a histone deacetylase (HDAC), named LphD. During my PhD, I showed that LphD is secreted into the host cell during infection and specifically targets the host cell nucleus, where it exhibits deacetylase activity with high efficiency for H3K14. Indeed, I showed that LphD deacetylates the H3K14 residue also during infection, and that the activity of LphD directly influences the levels of H3K14 methylation in infected cells, highlighting the synergy between LphD and RomA. I also could show that LphD and RomA target an endogenous chromatin binding complex, named HBO1, that contains the histone acetyltransferase KAT7, controlling the acetylation status of H3K14. RNAseq of cells infected with either wild type bacteria or the LphD and RomA knockout assessed the influence of these bacterial effectors on the host’s transcriptional landscape, in particular on genes related to immune response. The model I propose is that the two secreted effectors, LphD and RomA, work together to hijack the host’s epigenetic machinery in order to facilitate the subversion of the host immune response and promotes the intracellular replication of L. pneumophila.
Article
Full-text available
Pathogenic species of Legionella can infect human alveolar macrophages through Legionella-containing aerosols to cause a disease called Legionellosis, which has two forms: a flu-like Pontiac fever and severe pneumonia named Legionnaires’ disease (LD). Legionella is an opportunistic pathogen that frequently presents in aquatic environments as a biofilm or protozoa parasite. Long-term interaction and extensive co-evolution with various genera of amoebae render Legionellae pathogenic to infect humans and also generate virulence differentiation and heterogeneity. Conventionally, the proteins involved in initiating replication processes and human macrophage infections have been regarded as virulence factors and linked to pathogenicity. However, because some of the virulence factors are associated with the infection of protozoa and macrophages, it would be more accurate to classify them as survival factors rather than virulence factors. Given that the molecular basis of virulence variations among non-pathogenic, pathogenic, and highly pathogenic Legionella has not yet been elaborated from the perspective of virulence factors, a comprehensive explanation of how Legionella infects its natural hosts, protozoans, and accidental hosts, humans is essential to show a novel concept regarding the virulence factor of Legionella. In this review, we overviewed the pathogenic development of Legionella from protozoa, the function of conventional virulence factors in the infections of protozoa and macrophages, the host’s innate immune system, and factors involved in regulating the host immune response, before discussing a probably new definition for the virulence factors of Legionella.
Article
Full-text available
To prevail in the interaction with eukaryotic hosts, many bacterial pathogens use protein secretion systems to release virulence factors at the host–pathogen interface and/or deliver them directly into host cells. An outstanding example of the complexity and sophistication of secretion systems and the diversity of their protein substrates, effectors, is the Defective in organelle trafficking/Intracellular multiplication (Dot/Icm) Type IVB secretion system (T4BSS) of Legionella pneumophila and related species. Legionella species are facultative intracellular pathogens of environmental protozoa and opportunistic human respiratory pathogens. The Dot/Icm T4BSS translocates an exceptionally large number of effectors, more than 300 per L. pneumophila strain, and is essential for evasion of phagolysosomal degradation and exploitation of protozoa and human macrophages as replicative niches. Recent technological advancements in the imaging of large protein complexes have provided new insight into the architecture of the T4BSS and allowed us to propose models for the transport mechanism. At the same time, significant progress has been made in assigning functions to about a third of L. pneumophila effectors, discovering unprecedented new enzymatic activities and concepts of host subversion. In this review, we describe the current knowledge of the workings of the Dot/Icm T4BSS machinery and provide an overview of the activities and functions of the to-date characterized effectors in the interaction of L. pneumophila with host cells.
Article
Intracellular bacteria have developed a multitude of mechanisms to influence the post-translational modifications (PTMs) of host proteins to pathogen advantages. The recent explosion of insights into the diversity and sophistication of host PTMs and their manipulation by infectious agents challenges us to formulate a comprehensive vision of this complex and dynamic facet of the host-pathogen interaction landscape. As new discoveries continue to shed light on the central roles of PTMs in infectious diseases, technological advances foster our capacity to detect old and new PTMs and investigate their control and impact during pathogenesis, opening new possibilities for chemical intervention and infection treatment. Here, we present a comprehensive overview of these pathogenic mechanisms and offer perspectives on how these insights may contribute to the development of a new class of therapeutics that are urgently needed to face rising antibiotic resistances.
Article
Full-text available
Pathogens deliver complex arsenals of translocated effector proteins to host cells during infection, but the extent to which these proteins are regulated once inside the eukaryotic cell remains poorly defined. Among all bacterial pathogens, Legionella pneumophila maintains the largest known set of translocated substrates, delivering over 300 proteins to the host cell via its Type IVB, Icm/Dot translocation system. Backed by a few notable examples of effector-effector regulation in L. pneumophila, we sought to define the extent of this phenomenon through a systematic analysis of effector-effector functional interaction. We used Saccharomyces cerevisiae, an established proxy for the eukaryotic host, to query > 108,000 pairwise genetic interactions between two compatible expression libraries of ~330 L. pneumophila-translocated substrates. While capturing all known examples of effector-effector suppression, we identify fourteen novel translocated substrates that suppress the activity of other bacterial effectors and one pair with synergistic activities. In at least nine instances, this regulation is direct-a hallmark of an emerging class of proteins called metaeffectors, or "effectors of effectors". Through detailed structural and functional analysis, we show that metaeffector activity derives from a diverse range of mechanisms, shapes evolution, and can be used to reveal important aspects of each cognate effector's function. Metaeffectors, along with other, indirect, forms of effector-effector modulation, may be a common feature of many intracellular pathogens-with unrealized potential to inform our understanding of how pathogens regulate their interactions with the host cell.
Article
Full-text available
Yersinia species are bacterial pathogens that can cause plague and intestinal diseases after invading into human cells through the Three Secretion System (TTSS). The effect of pathogenesis is mediated by Yersinia outer proteins (Yop) and manifested as down-regulation of the cytokine genes expression by inhibiting nuclear factor-kappa-gene binding (NF-kappa B) and mitogen-activated protein kinase (MAPK) pathways. In addition, its pathogenesis can also manipulate the disorder of host innate immune system and cell death such as apoptosis, pyroptosis, and autophagy. Among the Yersinia effector proteins, YopB and YopD assist the injection of other virulence effectors into the host cytoplasm, while YopE, YopH, YopJ, YopO, and YopT target on disrupting host cell signaling pathways in the host cytosols. Many efforts have been applied to reveal that intracellular proteins such as Rho-GTPase, and transmembrane receptors such as Toll-like receptors (TLRs) both play critical roles in Yersinia pathogenesis, establishing a connection between the pathogenic process and the signaling response. This review will mainly focus on how the effector proteins of Yersinia modulate the intrinsic signals in host cells and disturb the innate immunity of hosts through TTSS.
Article
Full-text available
Infection by the human pathogen Legionella pneumophila relies on the translocation of ∼300 virulence proteins, termed effectors, which manipulate host cell processes. However, almost no information exists regarding effectors in other Legionella pathogens. Here we sequenced, assembled and characterized the genomes of 38 Legionella species and predicted their effector repertoires using a previously validated machine learning approach. This analysis identified 5,885 predicted effectors. The effector repertoires of different Legionella species were found to be largely non-overlapping, and only seven core effectors were shared by all species studied. Species-specific effectors had atypically low GC content, suggesting exogenous acquisition, possibly from the natural protozoan hosts of these species. Furthermore, we detected numerous new conserved effector domains and discovered new domain combinations, which allowed the inference of as yet undescribed effector functions. The effector collection and network of domain architectures described here can serve as a roadmap for future studies of effector function and evolution.
Article
Full-text available
In the last two years the Pfam database (http://pfam.xfam.org) has undergone a substantial reorganisation to reduce the effort involved in making a release, thereby permitting more frequent releases. Arguably the most significant of these changes is that Pfam is now primarily based on the UniProtKB reference proteomes, with the counts of matched sequences and species reported on the website restricted to this smaller set. Building families on reference proteomes sequences brings greater stability, which decreases the amount of manual curation required to maintain them. It also reduces the number of sequences displayed on the website, whilst still providing access to many important model organisms. Matches to the full UniProtKB database are, however, still available and Pfam annotations for individual UniProtKB sequences can still be retrieved. Some Pfam entries (1.6%) which have no matches to reference proteomes remain; we are working with UniProt to see if sequences from them can be incorporated into reference proteomes. Pfam-B, the automatically-generated supplement to Pfam, has been removed. The current release (Pfam 29.0) includes 16 295 entries and 559 clans. The facility to view the relationship between families within a clan has been improved by the introduction of a new tool.
Article
The mitogen-activated protein kinases (MAPKs) regulate diverse cellular programs by relaying extracellular signals to intracellular responses. In mammals, there are more than a dozen MAPK enzymes that coordinately regulate cell proliferation, differentiation, motility, and survival. The best known are the conventional MAPKs, which include the extracellular signal-regulated kinases 1 and 2 (ERK1/2), c-Jun amino-terminal kinases 1 to 3 (JNK1 to -3), p38 (α, β, γ, and δ), and ERK5 families. There are additional, atypical MAPK enzymes, including ERK3/4, ERK7/8, and Nemo-like kinase (NLK), which have distinct regulation and functions. Together, the MAPKs regulate a large number of substrates, including members of a family of protein Ser/Thr kinases termed MAPK-activated protein kinases (MAPKAPKs). The MAPKAPKs are related enzymes that respond to extracellular stimulation through direct MAPK-dependent activation loop phosphorylation and kinase activation. There are five MAPKAPK subfamilies: the p90 ribosomal S6 kinase (RSK), the mitogen- and stress-activated kinase (MSK), the MAPK-interacting kinase (MNK), the MAPK-activated protein kinase 2/3 (MK2/3), and MK5 (also known as p38-regulated/activated protein kinase [PRAK]). These enzymes have diverse biological functions, including regulation of nucleosome and gene expression, mRNA stability and translation, and cell proliferation and survival. Here we review the mechanisms of MAPKAPK activation by the different MAPKs and discuss their physiological roles based on established substrates and recent discoveries.
Article
Helicobacter pylori is a paradigm of persistent pathogens and major risk factor for developing severe diseases including adenocarcinoma in the human stomach. An important bacterial factor linked to gastric disease progression is the cag pathogenicity island-encoded type-IV secretion system (T4SS) effector protein CagA. Translocated CagA undergoes tyrosine phosphorylation at EPIYA-motifs and then activates or inactivates multiple host signaling proteins in a phosphorylation-dependent and phosphorylation-independent fashion. In this way, intracellular CagA acts as a "masterkey" or "picklock", which evolved during evolution to hijack key host cell signal transduction functions. Crucial targets of CagA represent a variety of serine/threonine and tyrosine kinases, which control major checkpoints of eukaryotic signaling. Here we review the signal transmission by translocated CagA on multiple receptor kinases (c-Met and EGFR) and non-receptor kinases (Src, Abl, Csk, aPKC, Par1, PI3K, Akt, FAK, GSK-3, JAK, PAK1, PAK2 and MAP kinases), manipulating a selection of fundamental processes in the human gastric epithelium such as cell adhesion, polarity, proliferation, motility, receptor endocytosis, cytoskeletal rearrangements, apoptosis, inflammation and cell cycle progression. This enormous complexity generates a highly remarkable and puzzling scenario during H. pylori infection. The contribution of these signaling pathways to bacterial survival, persistence and gastric pathogenesis is discussed. This article is protected by copyright. All rights reserved.
Article
Ubiquitination is a crucial post-translational protein modification involved in regulation of various cellular processes in eukaryotes. In particular, ubiquitination is involved in multiple aspects of bacterial infection and host defense mechanisms. In parallel with the identification of ubiquitination as a component of host defense systems, recently accumulated evidence shows that many bacterial pathogens exploit host ubiquitin systems to achieve successful infection. Here, we highlight the strategies by which bacteria subvert host ubiquitin systems by mimicking E3 ubiquitin ligase activity.
Article
Many bacterial pathogens use dedicated translocation systems to deliver arsenals of effector proteins to their hosts. Once inside the host cytosol, these effectors modulate eukaryotic cell biology to acquire nutrients, block microbial degradation, subvert host defenses, and enable pathogen transmission to other hosts. Among all bacterial pathogens studied to date, the gram-negative pathogen, Legionella pneumophila, maintains the largest arsenal of effectors, with over 330 effector proteins translocated by the Dot/Icm type IVB translocation system. In this review, I will discuss some of the recent work on understanding the consequences of this large arsenal. I will also present several models that seek to explain how L. pneumophila has acquired and subsequently maintained so many more effectors than its peers.
Article
LubX is part of the large arsenal of effectors in Legionella pneumophila that are translocated into the host cytosol during infection. Despite such unique features as the presence of two U-box motifs and its targeting of another effector SidH, the molecular basis of LubX activity remains poorly understood. Here we show that the N terminus of LubX is able to activate an extended number of ubiquitin-conjugating (E2) enzymes including UBE2W, UBEL6, and all tested members of UBE2D and UBE2E families. Crystal structures of LubX alone and in complex with UBE2D2 revealed drastic molecular diversification between the two U-box domains, with only the N-terminal U-box retaining E2 recognition features typical for its eukaryotic counterparts. Extensive mutagenesis followed by functional screening in a yeast model system captured functionally important LubX residues including Arg121, critical for interactions with SidH. Combined, these data provide a new molecular insight into the function of this unique pathogenic factor. Copyright © 2015 Elsevier Ltd. All rights reserved.