ArticlePDF AvailableLiterature Review

Abstract and Figures

Legionella pneumophila is an aquatic organism that interacts with amoebae and ciliated protozoa as the natural hosts, and this interaction plays a central role in bacterial ecology and infectivity. Upon transmission to humans, L. pneumophila infect and replicate within alveolar macrophages causing pneumonia. Intracellular proliferation of L. pneumophila within the two evolutionarily distant hosts is facilitated by bacterial exploitation of evolutionarily conserved host processes that are targeted by bacterial protein effectors injected into the host cell by the Dot/Icm type VIB translocation system. Although cysteine is semi-essential for humans and essential for amoeba, it is a metabolically favorable source of carbon and energy generation by L. pneumophila. To counteract host limitation of cysteine, L. pneumophila utilizes the AnkB Dot/Icm-translocated F-box effector to promote host proteasomal degradation of polyubiquitinated proteins within amoebae and human cells. Evidence indicates ankB and other Dot/Icm-translocated effector genes have been acquired through inter-kingdom horizontal gene transfer.
The environmental life cycle of L. pneumophila within protozoa. (1) Flagellated L. pneumophila infect protozoa in the aquatic environment. (2) The LCV evades the default endosomal-lysosomal degradation pathway and becomes rapidly remodeled by the ER through intercepting ER-toGolgi vesicle traffic and becomes rapidly decorated with polyubiquitinated proteins in an AnkB-dependent manner. (3) Under unfavorable stress conditions, such as nutrient deprivation, amoebae encyst, and bacterial proliferation will not occur due to nutrient limitation. Under growth-permissive conditions for the amoeba, the LCV is decorated with polyubiquitinated proteins, which are targeted for proteasomal degradation leading to elevated cellular levels of amino acids (AA) that power bacterial proliferation of the wild-type strain, while the ankB mutant is defective in this process and is unable to grow despite formation of ER-remodeled replicative LCV. (4) During late stages of infection, the LCV becomes disrupted leading to bacterial egress into the cytosol where the last 1-2 rounds of proliferations are completed. Upon nutrient depletion (see magnified box), RelA and SpoT are triggered leading to increased level of ppGpp, which triggers phenotypic transition into a flagellated virulent phenotype followed by lysis of the amoeba and bacterial escape from the host cell. Excreted vesicles filled with bacteria are also released. The infectious particle is not known but may include excreted Legionella-filled vesicles, intact Legionella-filled amoebae, or free Legionella that have been released from host cell. (5) Transmission to humans occurs via aerosols generated from man-made devices and installations, such as cooling towers, whirlpools, and showerheads.
… 
Content may be subject to copyright.
www.landesbioscience.com Virulence 307
Virulence 4:4, 307–314; May 15, 2013; © 2013 Landes Bioscience
REVIEW
REVIEW
Legionnaire Disease
The first recognized outbreak of Legionnaire disease occurred
in 1976 during the 56th annual American Legion Convention
in Philadelphia.
1
There were 180 individuals who were diag-
nosed with severe pneumonia and of these individuals, 29 died.
Less than one year after the outbreak the causative agent of the
disease was isolated.
2
The bacterium, which was designated
Legionella pneumophila, is a gram-negative facultative intracel-
lular bacterium that proliferates within alveolar macrophages,
causing Legionnaire disease.
3
There are more than 50 species of
Legionellae, but L. pneumophila continues to be responsible for
more than 80% of cases of Legionnaire disease in most of the
world. The exception is Western Australia, where L. longbeachae
is the most predominant species in causing disease, but its patho-
genesis is distinct from L. pneumophila.
4-6
*Correspondence to: Yousef Abu Kwaik; Email: abukwaik@louisville.edu
Submitted: 02/14/13; Revised: 03/12/13; Accepted: 03/13/13
http://dx.doi.org/10.4161/viru.24290
Legionella pneumophila is an aquatic organism that interacts
with amoebae and ciliated protozoa as the natural hosts, and
this interaction plays a central role in bacterial ecology and
infectivity. Upon transmission to humans, L. pneumophila
infect and replicate within alveolar macrophages causing
pneumonia. Intracellular proliferation of L. pneumophila within
the two evolutionarily distant hosts is facilitated by bacterial
exploitation of evolutionarily conserved host processes
that are targeted by bacterial protein eectors injected into
the host cell by the Dot/Icm type VIB translocation system.
Although cysteine is semi-essential for humans and essential
for amoebae, it is a metabolically favorable source of carbon
and energy generation by L. pneumophila. To counteract host
limitation of cysteine, L. pneumophila utilizes the AnkB Dot/
Icm-translocated F-box eector to promote host proteasomal
degradation of polyubiquitinated proteins within amoeba
and human cells. Evidence indicates ankB and other Dot/Icm-
translocated eector genes have been acquired through inter-
kingdom horizontal gene transfer.
Cellular microbiology and molecular ecology
of Legionella–amoeba interaction
Ashley M. Richards,
Juanita E. Von Dwingelo,
Christopher T. Price and Yousef Abu Kwaik*
Department of Microbiology and Immunology; College of Medicine; University of Louisville; Louisville, KY USA
These authors contributed equally to this work.
Keywords: polyubiquitin, farnesylation, prenylation, effectors AnkB, Ankyrin B, proteasomes, cysteine, ppGpp, RelA, SpoT,
Ankyrin, Dot/Icm, pneumophila, Legionnaire
Legionnaire disease and Pontiac fever, which is a flu-like ill-
ness caused by Legionella, have only emerged within the past few
decades due to human alterations to the environment. These
alterations include the use of air conditioning systems, whirl-
pools, and water cooling towers that generate aerosols as a vehicle
to transmit Legionella from aquatic sources (Fig. 1).
3
To date there
has been no documented cases of L. pneumophila transmission
between individuals, and transmission from the environment to
the human host is considered to be the main mode of transmis-
sion of L. pneumophila. Once L. pneumophila infects the human
host, its intracellular lifecycle has a striking similarity to that
within the evolutionarily distant natural host, amoebae (Fig. 1).
7
Legionella–Amoebae Interaction
L. pneumophila is ubiquitous in freshwater environments as well
as many man-made water systems worldwide, often in close asso-
ciation with freshwater protozoa.
8
L. pneumophila replicate at
temperatures of 2542 °C with an optimal growth temperature
of 35 °C. Consistent with what L. pneumophila would encounter
in the environment, motility and adherence to host cells are opti-
mal at temperatures below 37 °C.
9
When the temperature in the
aquatic environment is increased, the balance between bacteria
and amoebae can shift, which results in rapid multiplication of
Legionella.
10
The increase in the number of L. pneumophila in
the water as a result of proliferation within protozoa increases
the chance of transmission and disease manifestation.
3
Upon
decrease of temperature or exposure to environmental stress such
as chlorine, the amoebae differentiate into cyst,
10
and intracellular
L. pneumophila are capable of survival within the cyst
11
(Fig. 1).
Encysted amoeba is a highly resistant developmental stage that
contributes to the resistance of intracellular L. pneumophila to
different chemical and physical agents.
12
Therefore, the relation-
ship between L. pneumophila and amoebae plays an important
role in ecology and pathogenicity of the bacterium.
13
One mechanism by which the bacteria are released from
amoebae is within excreted vesicles similar to exocytosis of food
vacuoles.
14,15
L. pneumophila can still be cultured after 6 mo of
residence within excreted vesicles.
16,17
The number of bacteria
isolated from water sources of transmission to humans during
Legionnaire disease outbreaks is usually low or undetectable.
7,8
It is thought that the enhanced infectivity of L. pneumophila as
308 Virulence Volume 4 Issue 4
Numerous methods have been employed to attempt to eradi-
cate L. pneumophila from aquatic environments, with little
success. These attempts, which include chemical biocides, over-
heating water, and UV irradiation, have been successful for short
periods after which the bacteria can be again detected. It has been
suggested that in order to eradicate L. pneumophila from aquatic
environments continuous treatments effective against both the
bacteria and the protozoan host should be employed.
8,12 ,2 7, 33
Legionella-Like Amoebal Pathogens
There are Legionella-like species that cannot be grown on bac-
teriologic media but must be co-cultured with protozoa and are
referred to as Legionella-like amoebal pathogens (LLAP).
14
The
LLAPs are closely related to Legionella phylogenetically and
acquired their name because of their ability to infect and mul-
tiply within amoebae.
14
It is thought the LLAPs play a role in
community-acquired pneumonia and usually act as a co-patho-
gen but not as the sole pathogen.
34
Little is known about LLAPs
and future studies are needed to gain a better understanding of
the significance of these organisms in pulmonary infections.
Entry and Intracellular Trafcking of L. pneumophila
L. pneumophila infection of human alveolar macrophages is an
accidental infection and is thought to be a diversion from its
natural life cycle within amoebae. In the aquatic environment,
L. pneumophila resides in protozoa or in biofilms.
35,36
Amoebae
play a central role in the life cycle of L. pneumophila, and this
was first described in 1980.
37
Upon initial interaction between
L. pneumophila and amoebae, L. pneumophila is often engulfed
by coiling phagocytosis but other forms of internalization also
occur through a Gal/Gal-NAC specific receptor.
38,39
After inter-
nalization of L. pneumophila into the trophozoite of amoeba,
proliferation occurs within the Legionella-containing vacuole
(LCV) followed by bacterial release from the cell to seek a new
host (Fig. 1).
40-42
Once amoebae or mammalian cells engulf
L. pneumophila, the bacteria evade the default trafficking path-
ways into the lysosomal network (Fig. 1). The LCV recruits host
cell organelles, like mitochondria, ribosomes, and small vesicles
to its surface.
43
This accumulation begins during uptake into the
cell and is completed within a few minutes.
35
The ER-to-Golgi
vesicle traffic is intercepted by the LCV and the LCV membrane
becomes derived from the ER (Fig. 1).
13,43
The LCV becomes rap-
idly decorated with polyubiquitinated proteins within amoebae
and human cells.
44-46
Following maturation of the ER-remodeled
LCV and its decoration with polyubiquitinated proteins, rapid
replication of L. pneumophila commences (Fig. 1). During late
stages of intracellular proliferation, L. pneumophila escape from
the LCV to the cytosol where the bacteria finish the last 1–2
rounds of proliferation along with phenotypic modulations in
response to nutrient depletion in the host (Fig. 1).
8,40,47
The strategy used by L. pneumophila to avoid lysosomes and
to modulate cellular processes is dependent on the Dot/Icm type
IVB secretion system.
43,48,49
This secretion system injects ~300
effector proteins into the host cell, which accounts for ~10%
a result of growth within amoebae could compensate for the low
infectious dose in the water sources.
17-20
The ability of L. pneu-
mophila to parasitize human macrophages and to cause human
disease is thought to be a consequence of its prior adaptation to
intracellular growth within various protozoan hosts.
7,8
This is
most likely due to bacterial acquisition of eukaryotic genes dur-
ing its co-evolution with amoebae and adaptation to the intracel-
lular life within primitive eukaryotic hosts.
7,21-24
L. pneumophila and amoebae have been isolated from the same
source of infection during outbreaks of Legionnaire disease.
8
The
isolated amoebae have also been shown to support intracellular
replication of L. pneumophila.
25
It has been shown that L. pneu-
mophila that cannot be cultured in vitro using classical methods
can be resuscitated and proliferate if they are co-cultured with
amoebae.
8,26,27
Dictyostelium discoideum has been adapted as a
genetically amenable amoeba model system to decipher molec-
ular and cellular bases of L. pneumophila-amoeba interaction.
28
Therefore, amoebae are not only the environmental host for this
human pathogen, but constitute a genetically amenable model
system to study pathogenesis of L. pneumophila. These findings
demonstrate that studies on the L. pneumophila-amoeba interac-
tion will continue to contribute to our knowledge of the central
role of the amoeba host in the pathogenesis of these bacteria.
Amoebae Aid in the Persistence
of Legionella pneumophila in the Environment
Amoebae not only enhance the pathogenicity of L. pneumoph-
ila, but they also enable the bacteria to persist in the environ-
ment. Fourteen species of amoebae, with Hartmannellae and
Acanthamoeba being the most prominent, and two species of cili-
ated protozoa have been shown to support intracellular replica-
tion of L. pneumophila.
8
L. pneumophila infects the trophozoite
form of amoebae and the presence of the bacteria within amoebae
serves to protect the bacteria from harsh environments.
27
L. pneu-
mophila does not proliferate within encysted amoeba but when
conditions become unfavorable, protozoa can differentiate from
their trophozoite form into a cyst form that protects the organ-
isms and ensures their survival (Fig. 1). L. pneumophila released
from free-living amoebae also show an increased resistance
to harsh conditions compared with those grown in vitro.
29-32
When compared with bacteria grown in vitro, bacteria grown
in amoebae have changes in biochemistry, physiology, and viru-
lence potential.
12
These changes include an enhanced resistance
to chemicals and antibiotics, an altered fatty acid profile and pro-
tein profile, shorter size, motility, an increased ability to infect
amoebae and mammalian cells, an increase in environmental fit-
ness, and an increase in uptake.
8,19,20,26,27
In addition, the bacteria
found within vesicles excreted from amoebae are highly resistant
to biocides while the vesicle is resistant to freezing and sonica-
tion.
17
After prolonged starvation of L. pneumophila or treat-
ment with chlorine that renders L. pneumophila non-culturable
in aquatic environments, the bacteria can’t be cultured on rich
media but they can be resuscitated by infection of amoebae,
clearly indicating the remarkable protection of L. pneumophila
through its intracellular niche within amoebae.
26,27
www.landesbioscience.com Virulence 309
exponential phase of intracellular growth within macrophages
and Acanthamoeba,
47,59- 61
but the function of most of these pro-
teins has yet to be determined.
Growth phase-dependent regulation of bacterial virulence.
The intracellular lifecycle of L. pneumophila consists of a repli-
cative phase within the LCV, and a transmissive phase, exhib-
ited upon escape into the cytosol.
40,47,62,63
This biphasic lifestyle
is characterized by dramatic changes in the transcriptome, that
result in phenotypic modulations.
59-61
During the replicative
of the coding capacity of the genome of L. pneumophila.
7,50,51
Although a large number of effectors are injected into the host
cell, but with only few exceptions, deletion of individual effectors
does not result in reduced intracellular proliferation, suggesting
potential functional redundancy.
7,43,51,52
Strikingly, many L. pneu-
mophila effector proteins harbor eukaryotic protein domains,
which include ankyrin repeats, leucine-rich repeats, Sel-1,
U-box, F-box, and a C-terminal CaaX prenylation motif.
23,53-58
Expression of a large number of effectors is induced during the
Figure 1. The environmental life cycle of L. pneumophila within protozoa. (1) Flagellated L. pneumophila infect protozoa in the aquatic environment.
(2) The LCV evades the default endosomal–lysosomal degradation pathway and becomes rapidly remodeled by the ER through intercepting ER-to-
Golgi vesicle trac and becomes rapidly decorated with polyubiquitinated proteins in an AnkB-dependent manner. (3) Under unfavorable stress con-
ditions, such as nutrient deprivation, amoebae encyst, and bacterial proliferation will not occur due to nutrient limitation. Under growth-permissive
conditions for the amoeba, the LCV is decorated with polyubiquitinated proteins, which are targeted for proteasomal degradation leading to elevated
cellular levels of amino acids (AA) that power bacterial proliferation of the wild-type strain, while the ankB mutant is defective in this process and is
unable to grow despite formation of ER-remodeled replicative LCV. (4) During late stages of infection, the LCV becomes disrupted leading to bacterial
egress into the cytosol where the last 1–2 rounds of proliferations are completed. Upon nutrient depletion (see magnied box), RelA and SpoT are trig-
gered leading to increased level of ppGpp, which triggers phenotypic transition into a agellated virulent phenotype followed by lysis of the amoeba
and bacterial escape from the host cell. Excreted vesicles lled with bacteria are also released. The infectious particle is not known but may include ex-
creted Legionella-lled vesicles, intact Legionella-lled amoebae, or free Legionella that have been released from host cell. (5) Transmission to humans
occurs via aerosols generated from man-made devices and installations, such as cooling towers, whirlpools, and showerheads.
310 Virulence Volume 4 Issue 4
phase, the bacterium is undergoing exponential (E) growth; it
is non-motile and represses transmissive traits, such as lysosomal
evasion (Fig. 1).
64
The stringent-like response is triggered upon
transition of L. pneumophila into post-exponential (PE) growth,
when the bacteria become cytotoxic, motile, sodium-sensitive,
osmotic-resistant, and capable of lysosomal evasion.
63,64
These
phenotypic modulations are necessary to escape from the wasted
host and invade a new host to start a second cycle of intracellular
proliferation.
41,42,47,59,60,63-65
The transition between replicative and transmissive pheno-
types is highly orchestrated, and is governed by many factors that
are influenced by intracellular nutrient levels.
7,24,63,66
Upon amino
acid depletion, uncharged bacterial tRNAs activate RelA to syn-
thesize the bacterial alarmone 3',5'-bispyrophosphate (ppGpp),
a master regulator of numerous genes of L. pneumophila, which
triggers phenotypic modulations upon transition into the PE
phase.
63
SpoT, a bifunctional synthetase/hydrolase that responds
to a variety of stimuli, such as fatty acid starvation, also synthe-
sizes ppGpp leading to increased levels of the alarmone (Fig. 1).
67
RpoS and several global response two-component regulators,
such as LetA/S and PmrA/B, function as downstream cascades
of regulatory networks that govern phenotypic modulations at
the PE phase.
61,66,68-70
Small non-coding RNAs, such as RsmY
and RsmZ, are induced at the PE phase by the regulatory cascade
of networks triggered by elevated ppGpp levels.
66,68,71
The RNA
polymerase interacting protein, DskA, also responds to increased
levels of ppGpp and other stress signals to coordinate phenotypic
modulations of L. pneumophila at the PE phase and its transmis-
sion to a new host.
67
In addition to triggering flagellation and various virulence-
related traits, elevated ppGpp levels result in upregulation of the
type IV-secretion components and many of its exported effec-
tors.
59-61
One of the Dot-Icm-translocated effectors important
in the intracellular infection of amoebae and human cells is the
eukaryotic-like AnkB,
33,44,45
which is temporally and differen-
tially regulated at the PE phase.
33,35,61,72
Therefore, complex cas-
cades of regulatory networks govern phenotypic transition at the
PE phase and most or all of these networks are under the direct
or indirect control of ppGpp.
In addition to phenotypic modulations at the PE phase,
L. pneumophila undergoes a differentiation cycle that is dimor-
phic, cycling between a replicating form and a planktonic spore-
like cyst form, designated as mature intracellular form (MIF).
73-75
The MIF is near dormant metabolically, resistant to detergents
and antibiotics, and is more invasive.
73
The MIFs are detectable
in HeLa cells but do not form in macrophages, which is likely
due to early apoptotic lysis within 1–3 d of the infection, while
the MIFs are formed later in HeLa cells.
73
The MIFs of L. pneu-
mophila germinate following entry into a susceptible eukaryotic
host cell or in rich media in vitro. It is possible that the MIFs
contribute to the ecology of L. pneumophila during starvation in
the water system when nutrients are depleted and the amoebal
hosts are encysted and not susceptible to infection.
Exploitation of conserved eukaryotic processes by the
eukaryotic-like AnkB effector of L. pneumophila. L. pneumoph-
ila harbors a plethora of eukaryotic-like effectors that interfere
with host processes by mimicking eukaryotic functions.
21,23,53,54
Many translocated effectors of L. pneumophila are functionally
and structurally similar to eukaryotic proteins and interact with
and disrupt various eukaryotic processes such as signaling, pro-
tein synthesis, apoptosis, posttranslational modification, vesicu-
lar trafficking, ubiquitination, and proteasomal degradation.
43
Among the ~300 effectors of L. pneumophila, AnkB is the only
effector known to be indispensable for the intracellular infec-
tion of both human cells and amoebae, and the biological func-
tion of this effector has been deciphered. It is not surprising that
recent studies on the AnkB effector and its exploitation of mul-
tiple highly conserved eukaryotic processes may just be the tip of
the iceberg of our continued unraveling of L. pneumophila–host
interaction and its evolution from invading amoebae to invading
human cells and causing pneumonia.
The AnkB effector harbors multiple eukaryotic domains that
enable this protein to hijack a number of evolutionarily conserved
eukaryotic processes, and is essential for intracellular prolifera-
tion of L. pneumophila in amoebae and human cells and for viru-
lence in the mouse model.
33,44,45,76
The AnkB effector harbors two
Ankyrin domains (ANK), 33-residue repeats involved in protein-
protein interactions, and is the most common domain in eukary-
otic proteins.
56
AnkB also contains a C-terminal eukaryotic
CaaX motif (C, cysteine; a, aliphatic amino acid; X,I any amino
acid) that allows the protein to be lipidated through farnesylation
by the host farnesyltransferase (FTase), which anchors AnkB
into the LCV membrane (Fig. 2).
58,77-79
Farnesylation is a type
of prenylation that covalently links a 15-carbon lipid moiety to a
conserved cysteine residue within the C-terminus “CaaX” motif,
which confers hydrophobicity enabling the lipidated protein to
be anchored into the lipid bi-layer of eukaryotic membranes.
58
However, there is variation in the C-terminus CaaX motif of
AnkB, as some isolates, such as the Paris strain of L. pneumoph-
ila, have a truncated C-terminus without a CaaX motif.
45
The
biological relevance of this variation is still to be determined.
Ubiquitination of proteins is a highly conserved eukaryotic
post-translation modification that is mediated by three enzymes
(E1–E3); E1 is the activating enzyme which transfers a 76-amino
acid ubiquitin polypeptide to the conjugating enzyme (E2) while
the E3 ubiquitin ligase links ubiquitin to the target protein.
7,76
Polyubiquitin is formed by linking ubiquitin monomers through
one of the 7 lysine (K) residues of ubiquitin. Polyubiquitination
through K
48
linkages targets the modified protein for proteo-
lytic degradation by the proteasomes.
76
The AnkB effector is a
bona fide F-box protein that binds the E3 eukaryotic ubiquitin
ligase and functions as a platform for the assembly of K
48
-linked
polyubiquitinated proteins on the LCV (Fig. 2),
44,45
which occurs
within a few minutes of bacterial entry.
45,80
Proteasomal degrada-
tion of K
48
-linked polyubiquitinated proteins results in increased
cellular levels of amino acids (Fig. 2),
7,80
which are essential for
intracellular proliferation of L. pneumophila that is dependent on
amino acids as the major source of carbon and energy to feed the
tri-carboxylic acid (TCA) cycle.
81
The ankB mutant of L. pneumophila is severely defective in
intracellular proliferation in amoebae and human macrophages
due to the defect in assembly of K
48
-linked polyubiquitinated
www.landesbioscience.com Virulence 311
proteins decorating the LCV (Fig. 1).
7,45,80
Due to lack of pro-
teasomal degradation of K
48
-linked polyubiquitin during infec-
tion by the ankB mutant, cellular levels of amino acid do not
increase. This triggers a bacterial starvation response, mediated
by the induced expression of RelA and SpoT, and results in ele-
vated ppGpp levels.
7,80
Intracellular growth can be restored to the
ankB mutant within amoebae and human cells by supplement-
ing excess amino acids.
7,24,80
Thus, higher levels of cellular amino
acids are needed for intracellular replication of L. pneumophila.
Remarkably, supplementation of infected cells with certain sin-
gle amino acids, such as cysteine, reverses the growth defect of
the ankB mutant in amoebae and human cells. Interestingly, in
human cells cysteine is semi-essential and is the least abundant
amino acid, but in amoebae cysteine is essential.
24,80
Similar
to cysteine, supplementation of infected cells with pyruvate or
citrate to feed the TCA cycle, rescues the ankB mutant for intra-
cellular proliferation.
7,24,80
Interestingly, in vitro growth of L.
pneumophila in rich medium requires supplementation with 3.3
mM cysteine.
80
Therefore, AnkB is a remarkable example of an
effector involved in exploitation of multiple host processes that
are highly conserved in unicellular eukaryotes and mammals.
7
By promoting proteasomal degradation in amoebae and human
cells though the AnkB F-box effector, L. pneumophila generates
a gratuitous supply of cellular amino acids (Fig. 2), particularly
the limiting ones such as Cys, which is a metabolically favorable
source of carbon and energy for L. pneumophila to power intra-
cellular growth within amoebae and human cells.
24
Nutritional adaptation of L. pneumophila to amoebae.
Amino acids are the main sources of carbon and energy to feed
the TCA cycle of L. pneumophila, which has a defect in the gly-
colytic pathway, but carbohydrate metabolism plays a minor role
in contribution to central metabolism.
81
Cysteine and Serine are
particularly important to feed central metabolism of L. pneu-
mophila. Both amino acids are converted into pyruvate that feed
the TCA cycle, which is the primary pathway of carbon and
energy production in L. pneumophila.
24,80,81
L. pneumophila and
its primary host (Acanthamoebae or Dictyostelium discoideum) are
both auxotrophic for cysteine while L. pneumophila is also auxo-
trophic for arginine, isoleucine, leucine, methionine, valine, and
threonine.
7,24
Interestingly, toxicity of proteasomal inhibition in
human cells results from the lack of protein synthesis due to low
levels of limiting amino acids, particularly cysteine.
82
Therefore,
Figure 2. Nutritional and metabolic adaptation of L. pneumophila to the intracellular life within amoebae and human cells is facilitated by the AnkB
eector and its exploitation of multiple highly conserved eukaryotic processes. The AnkB eector is translocated into host cells by the Dot/Icm type
IV secretion system of L. pneumophila, and it is immediately farnesylated by the three host enzymes FTase, RCE1, and ICMT, that are recruited to the
LCV by the Dot/Icm system.
78
Farnesylation of AnkB results in its anchoring into the cytosolic face of the LCV membrane where it interacts with the
eukaryotic SCF1 ubiquitin ligase complex. The AnkB eector functions as a platform for the docking of K
48
-linked polyubiquitinated proteins to the
LCV. Proteasomal degradation of the K
48
-linked polyubiquitinated protein generates 2–24 amino acid (AA) peptides that are rapidly degraded by
oligo- and amino-peptidases. This generates a surplus of cellular amino acids within the cytosol of L. pneumophila-infected amoebae and human cells.
The amino acids are imported into the LCV through various host amino acid transporters present in the LCV membrane, including the neutral amino
acid transporter SLC1A5, which imports Cys, and subsequently into L. pneumophila through numerous ABC transporters and amino acid permeases
such as the threonine transporter PhtA.
94
312 Virulence Volume 4 Issue 4
through high nutritional dependence of L. pneumophila on host
limiting amino acids, such as Cys, and synchronization of amino
acid auxotrophy with its host, L. pneumophila synchronizes its
nutritional needs for growth with the availability of nutrients for
the host.
7,24
This remarkable nutritional adaptation could be what
allows the bacteria to be protected during amoebal encystation
upon nutrient depletion, leading to cessation of intracellular bac-
terial growth (Fig. 1). Encysted amoebae are thought to be pro-
tected against invasion by L. pneumophila.
75
Therefore, potential
growth and release of L. pneumophila from amoeba in an aquatic
environment in which amoebae have encysted in response to
nutrient depletion is unlikely to happen, since it would result in
free-living bacteria without a susceptible host, which may result
is eventual loss of bacterial viability. However, potential differen-
tiation into the MIF under starvation conditions may facilitate
continued presence of L. pneumophila in the aquatic environment
under these conditions.
73,74
Acquisition of eukaryotic genes through inter-kingdom
horizontal gene transfer. The long-term co-evolution of L. pneu-
mophila with various protists and metazoa has influenced the
genomic structure of this organism through inter-kingdom hori-
zontal gene transfer (HGT).
7,23,83,84
This long-term co-evolution
is likely what gave rise to the acquisition of eukaryotic host genes
encoding proteins with eukaryotic-like functions and structures.
Amoeba may act as a gene melting pot, allowing diverse micro-
organisms to evolve by gene acquisition and loss, and then either
adapt to the intra-amoebal lifestyle or evolve into new patho-
gens. Interestingly, mammalian F-box proteins do not have the
ANK domain, while F-box proteins from amoebae do.
7,24,5 6 ,76
Therefore, it is more likely that ankB had been acquired through
inter-kingdom HGT from a primitive eukaryotic host.
7,24,76
L. pneumophila is a naturally competent organism that takes up
DNA and can exchange DNA between bacteria through conju-
gation.
9,48,49
Long-term convergent evolution and modification of
the genes acquired through HGT, splicing of introns, acquisi-
tion of prokaryotic promoters and regulators, and translocation
motifs is likely what allowed eukaryotic-like proteins to become
translocated effectors with functional activities in the host cell.
83
It is to be expected that many of the eukaryotic-like proteins in
L. pneumophila are still undergoing convergent evolution through
modifications that might enable them to become translocated
and functionally active effectors.
7
Long-term co-evolution with its protozoan hosts has likely
contributed to the ability of L. pneumophila to cause dis-
ease in humans, perpetuated by changes in human lifestyle.
Understanding its association with amoebae will give us a bet-
ter understanding of how L. pneumophila causes human dis-
ease though exploitation of evolutionary conserved eukaryotic
processes.
7,24,80
Since L. pneumophila also exploits mammalian-
specific processes such as the inflammasomes and pro- and anti-
apoptosis,
85-92
it is likely that additional virulence properties have
been acquired by L. pneumophila to enhance its capacity to infect
humans. Since many other pathogens are detected within amoe-
bae, this primitive eukaryotic host may represent a reservoir for
many human pathogens.
93
Disclosure of Potential Conflicts of Interest
No potential conicts of interest were disclosed.
References
1. Fraser DW, Tsai TR, Orenstein W, Parkin WE,
Beecham HJ, Sharrar RG, et al. Legionnaires’ disease:
description of an epidemic of pneumonia. N Engl
J Med 1977; 297:1189-97; PMID:335244; http://
dx.doi.org/10.1056/NEJM197712012972201
2. McDade JE, Shepard CC, Fraser DW, Tsai TR, Redus
MA, Dowdle WR. Legionnaires’ disease: isolation of
a bacterium and demonstration of its role in other
respiratory disease. N Engl J Med 1977; 297:1197-
203; PMID:335245; http://dx.doi.org/10.1056/
NEJM197712012972202
3. Fields BS, Benson RF, Besser RE. Legionella and
Legionnaires’ disease: 25 years of investigation. Clin
Microbiol Rev 2002; 15:506-26; PMID:12097254;
http://dx.doi.org/10.1128/CMR.15.3.506-526.2002
4. Asare R, Santic M, Gobin I, Doric M, Suttles J,
Graham JE, et al. Genetic susceptibility and caspase
activation in mouse and human macrophages are dis-
tinct for Legionella longbeachae and L. pneumophila.
Infect Immun 2007; 75:1933-45; PMID:17261610;
http://dx.doi.org/10.1128/IAI.00025-07
5. Asare R, Abu Kwaik Y. Early trafficking and intracel-
lular replication of Legionella longbeachaea within
an ER-derived late endosome-like phagosome. Cell
Microbiol 2007; 9:1571-87; PMID:17309675; http://
dx.doi.org/10.1111/j.1462-5822.2007.00894.x
6. Cazalet C, Gomez-Valero L, Rusniok C, Lomma M,
Dervins-Ravault D, Newton HJ, et al. Analysis of the
Legionella longbeachae genome and transcriptome
uncovers unique strategies to cause Legionnaires’ dis-
ease. PLoS Genet 2010; 6:e1000851; PMID:20174605;
http://dx.doi.org/10.1371/journal.pgen.1000851
7. Al-Quadan T, Price CT, Abu Kwaik Y. Exploitation of
evolutionarily conserved amoeba and mammalian pro-
cesses by Legionella. Trends Microbiol 2012; 20:299-
306; PMID:22494803; http://dx.doi.org/10.1016/j.
tim.2012.03.005
8. Molmeret M, Horn M, Wagner M, Santic M, Abu
Kwaik Y. Amoebae as training grounds for intracellular
bacterial pathogens. Appl Environ Microbiol 2005;
71:20-8; PMID:15640165; http://dx.doi.org/10.1128/
AEM.71.1.20-28.2005
9. Stone BJ, Kwaik YA. Natural competence for DNA
transformation by Legionella pneumophila and its
association with expression of type IV pili. J Bacteriol
1999; 181:1395-402; PMID:10049368.
10. Ohno A, Kato N, Sakamoto R, Kimura S, Yamaguchi
K. Temperature-dependent parasitic relationship
between Legionella pneumophila and a free-living
amoeba (Acanthamoeba castellanii). Appl Environ
Microbiol 2008; 74:4585-8; PMID:18502936; http://
dx.doi.org/10.1128/AEM.00083-08
11. Kilvington S, Price J. Survival of Legionella pneu-
mophila within cysts of Acanthamoeba polyph-
aga following chlorine exposure. J Appl Bacteriol
1990; 68:519-25; PMID:2196257; http://dx.doi.
org/10.1111/j.1365-2672.1990.tb02904.x
12. Abu Kwaik Y, Gao LY, Stone BJ, Venkataraman C,
Harb OS. Invasion of protozoa by Legionella pneu-
mophila and its role in bacterial ecology and patho-
genesis. Appl Environ Microbiol 1998; 64:3127-33;
PMID:9726849.
13. Segal G, Shuman HA. Legionella pneumophila utilizes
the same genes to multiply within Acanthamoeba castel-
lanii and human macrophages. Infect Immun 1999;
67:2117-24; PMID:10225863.
14. Adeleke A, Pruckler J, Benson R, Rowbotham T,
Halablab M, Fields BS. Legionella-like amebal
pathogens--phylogenetic status and possible role in
respiratory disease. Emerg Infect Dis 1996; 2:225-
30; PMID:8903235; http://dx.doi.org/10.3201/
eid0203.960311
15. Chen J, de Felipe KS, Clarke M, Lu H, Anderson OR,
Segal G, et al. Legionella effectors that promote non-
lytic release from protozoa. Science 2004; 303:1358-
61; PMID:14988561; http://dx.doi.org/10.1126/sci-
ence.1094226
16. Bouyer S, Imbert C, Rodier MH, Héchard Y. Long-
term survival of Legionella pneumophila associat-
ed with Acanthamoeba castellanii vesicles. Environ
Microbiol 2007; 9:1341-4; PMID:17472646; http://
dx.doi.org/10.1111/j.1462-2920.2006.01229.x
17. Berk SG, Ting RS, Turner GW, Ashburn RJ. Production
of respirable vesicles containing live Legionella pneu-
mophila cells by two Acanthamoeba spp. Appl Environ
Microbiol 1998; 64:279-86; PMID:9435080.
18. O’Brien SJ, Bhopal RS. Legionnaires’ disease: the
infective dose paradox. Lancet 1993; 342:5-6;
PMID:8100317; http://dx.doi.org/10.1016/0140-
6736(93)91877-O
19. Cirillo JD, Falkow S, Tompkins LS. Growth of
Legionella pneumophila in Acanthamoeba castellanii
enhances invasion. Infect Immun 1994; 62:3254-61;
PMID:8039895.
20. Cirillo JD, Cirillo SL, Yan L, Bermudez LE, Falkow S,
Tompkins LS. Intracellular growth in Acanthamoeba
castellanii affects monocyte entry mechanisms and
enhances virulence of Legionella pneumophila. Infect
Immun 1999; 67:4427-34; PMID:10456883.
www.landesbioscience.com Virulence 313
21. Lurie-Weinberger MN, Gomez-Valero L, Merault N,
Glöckner G, Buchrieser C, Gophna U. The origins of
eukaryotic-like proteins in Legionella pneumophila. Int
J Med Microbiol 2010; 300:470-81; PMID:20537944;
http://dx.doi.org/10.1016/j.ijmm.2010.04.016
22. Buchrieser C. Legionella: from protozoa to humans.
Front Microbiol 2011; 2:182; PMID:22016745;
http://dx.doi.org/10.3389/fmicb.2011.00182
23. Gomez-Valero L, Rusniok C, Cazalet C, Buchrieser
C. Comparative and functional genomics of legio-
nella identified eukaryotic like proteins as key players
in host-pathogen interactions. Front Microbiol 2011;
2:208; PMID:22059087; http://dx.doi.org/10.3389/
fmicb.2011.00208
24. Price C, Abu Kwaik Y. Amoebae and Mammals Deliver
Protein-Rich Atkins Diet Meals to Legionella. Microbe
2012; 7:506-13.
25. Fields BS, Nerad TA, Sawyer TK, King CH, Barbaree
JM, Martin WT, et al. Characterization of an axenic
strain of Hartmannella vermiformis obtained from an
investigation of nosocomial legionellosis. J Protozool
1990; 37:581-3; PMID:2086787.
26. Steinert M, Emödy L, Amann R, Hacker J.
Resuscitation of viable but nonculturable Legionella
pneumophila Philadelphia JR32 by Acanthamoeba cas-
tellanii. Appl Environ Microbiol 1997; 63:2047-53;
PMID:9143134.
27. García MT, Jones S, Pelaz C, Millar RD, Abu Kwaik
Y. Acanthamoeba polyphaga resuscitates viable non-
culturable Legionella pneumophila after disinfection.
Environ Microbiol 2007; 9:1267-77; PMID:17472639;
http://dx.doi.org/10.1111/j.1462-2920.2007.01245.x
28. Hägele S, Köhler R, Merkert H, Schleicher M, Hacker
J, Steinert M. Dictyostelium discoideum: a new
host model system for intracellular pathogens of the
genus Legionella. Cell Microbiol 2000; 2:165-71;
PMID:11207573; http://dx.doi.org/10.1046/j.1462-
5822.2000.00044.x
29. Abu Kwaik Y, Gao LY, Harb OS, Stone BJ.
Transcriptional regulation of the macrophage-induced
gene (gspA) of Legionella pneumophila and pheno-
typic characterization of a null mutant. Mol Microbiol
1997; 24:629-42; PMID:9179855; http://dx.doi.
org/10.1046/j.1365-2958.1997.3661739.x
30. Barker J, Brown MRW, Collier PJ, Farrell I, Gilbert
P. Relationship between Legionella pneumophila and
Acanthamoeba polyphaga: physiological status and
susceptibility to chemical inactivation. Appl Environ
Microbiol 1992; 58:2420-5; PMID:1514790.
31. Barker J, Lambert PA, Brown MRW. Influence of intra-
amoebic and other growth conditions on the surface
properties of Legionella pneumophila. Infect Immun
1993; 61:3503-10; PMID:8335382.
32. Barker J, Scaife H, Brown MRW. Intraphagocytic
growth induces an antibiotic-resistant phenotype of
Legionella pneumophila. Antimicrob Agents Chemother
1995; 39:2684-8; PMID:8593002; http://dx.doi.
org/10.1128/AAC.39.12.2684
33. Al-Khodor S, Price CT, Habyarimana F, Kalia A, Abu
Kwaik YA. A Dot/Icm-translocated ankyrin protein of
Legionella pneumophila is required for intracellular
proliferation within human macrophages and protozoa.
Mol Microbiol 2008; 70:908-23; PMID:18811729.
34. Marrie TJ, Raoult D, La Scola B, Birtles RJ, de Carolis
E; Canadian Community-Acquired Pneumonia Study
Group. Legionella-like and other amoebal pathogens
as agents of community-acquired pneumonia. Emerg
Infect Dis 2001; 7:1026-9; PMID:11747734; http://
dx.doi.org/10.3201/eid0706.010619
35. Bitar DM, Molmeret M, Abu Kwaik Y. Molecular
and cell biology of Legionella pneumophila. Int J
Med Microbiol 2004; 293:519-27; PMID:15149027;
http://dx.doi.org/10.1078/1438-4221-00286
36. Molmeret M, Bitar DM, Han L, Kwaik YA. Cell biology
of the intracellular infection by Legionella pneumophi-
la. Microbes Infect 2004; 6:129-39; PMID:14738901;
http://dx.doi.org/10.1016/j.micinf.2003.11.004
37. Rowbotham TJ. Preliminary report on the patho-
genicity of Legionella pneumophila for freshwater
and soil amoebae. J Clin Pathol 1980; 33:1179-
83; PMID:7451664; http://dx.doi.org/10.1136/
jcp.33.12.1179
38. Venkataraman C, Haack BJ, Bondada S, Abu Kwaik
Y. Identification of a Gal/GalNAc lectin in the proto-
zoan Hartmannella vermiformis as a potential receptor
for attachment and invasion by the Legionnaires
disease bacterium. J Exp Med 1997; 186:537-47;
PMID:9254652; http://dx.doi.org/10.1084/
jem.186.4.537
39. Venkataraman C, Gao LY, Bondada S, Kwaik YA.
Identification of putative cytoskeletal protein homo-
logues in the protozoan host Hartmannella vermi-
formis as substrates for induced tyrosine phosphatase
activity upon attachment to the Legionnaires’ dis-
ease bacterium, Legionella pneumophila. J Exp Med
1998; 188:505-14; PMID:9687528; http://dx.doi.
org/10.1084/jem.188.3.505
40. Molmeret M, Bitar DM, Han L, Kwaik YA.
Disruption of the phagosomal membrane and egress
of Legionella pneumophila into the cytoplasm during
the last stages of intracellular infection of macro-
phages and Acanthamoeba polyphaga. Infect Immun
2004; 72:4040-51; PMID:15213149; http://dx.doi.
org/10.1128/IAI.72.7.4040-4051.2004
41. Alli OAT, Gao LY, Pedersen LL, Zink S, Radulic M,
Doric M, et al. Temporal pore formation-mediated
egress from macrophages and alveolar epithelial cells by
Legionella pneumophila. Infect Immun 2000; 68:6431-
40; PMID:11035756; http://dx.doi.org/10.1128/
IAI.68.11.6431-6440.2000
42. Gao LY, Kwaik YA. The mechanism of killing and
exiting the protozoan host Acanthamoeba polyphaga
by Legionella pneumophila. Environ Microbiol 2000;
2:79-90; PMID:11243265; http://dx.doi.org/10.1046/
j.1462-2920.2000.00076.x
43. Isberg RR, O’Connor TJ, Heidtman M. The Legionella
pneumophila replication vacuole: making a cosy niche
inside host cells. Nat Rev Microbiol 2009; 7:13-24;
PMID:19011659; http://dx.doi.org/10.1038/nrmi-
cro1967
44. Price CT, Al-Khodor S, Al-Quadan T, Santic M,
Habyarimana F, Kalia A, et al. Molecular mimicry by
an F-box effector of Legionella pneumophila hijacks a
conserved polyubiquitination machinery within macro-
phages and protozoa. PLoS Pathog 2009; 5:e1000704;
PMID:20041211; http://dx.doi.org/10.1371/journal.
ppat.1000704
45. Lomma M, Dervins-Ravault D, Rolando M, Nora
T, Newton HJ, Sansom FM, et al. The Legionella
pneumophila F-box protein Lpp2082 (AnkB) modu-
lates ubiquitination of the host protein parvin B
and promotes intracellular replication. Cell Microbiol
2010; 12:1272-91; PMID:20345489; http://dx.doi.
org/10.1111/j.1462-5822.2010.01467.x
46. Dorer MS, Kirton D, Bader JS, Isberg RR. RNA
interference analysis of Legionella in Drosophila cells:
exploitation of early secretory apparatus dynamics.
PLoS Pathog 2006; 2:e34; PMID:16652170; http://
dx.doi.org/10.1371/journal.ppat.0020034
47. Molmeret M, Jones S, Santic M, Habyarimana F,
Esteban MT, Kwaik YA. Temporal and spatial trig-
ger of post-exponential virulence-associated regula-
tory cascades by Legionella pneumophila after bacterial
escape into the host cell cytosol. Environ Microbiol
2010; 12:704-15; PMID:19958381; http://dx.doi.
org/10.1111/j.1462-2920.2009.02114.x
48. Vogel JP, Andrews HL, Wong SK, Isberg RR.
Conjugative transfer by the virulence system of
Legionella pneumophila. Science 1998; 279:873-6;
PMID:9452389; http://dx.doi.org/10.1126/sci-
ence.279.5352.873
49. Segal G, Purcell M, Shuman HA. Host cell killing and
bacterial conjugation require overlapping sets of genes
within a 22-kb region of the Legionella pneumophila
genome. Proc Natl Acad Sci U S A 1998; 95:1669-
74; PMID:9465074; http://dx.doi.org/10.1073/
pnas.95.4.1669
50. Zhu W, Banga S, Tan Y, Zheng C, Stephenson R,
Gately J, et al. Comprehensive identification of pro-
tein substrates of the Dot/Icm type IV transporter of
Legionella pneumophila. PLoS One 2011; 6:e17638;
PMID:21408005; http://dx.doi.org/10.1371/journal.
pone.0017638
51. Luo ZQ. Targeting One of its Own: Expanding Roles
of Substrates of the Legionella Pneumophila Dot/Icm
Type IV Secretion System. Front Microbiol 2011;
2:31; PMID:21687422; http://dx.doi.org/10.3389/
fmicb.2011.00031
52. Luo ZQ. Striking a balance: modulation of host cell
death pathways by legionella pneumophila. Front
Microbiol 2011; 2:36; PMID:21687427; http://dx.doi.
org/10.3389/fmicb.2011.00036
53. Cazalet C, Rusniok C, Brüggemann H, Zidane N,
Magnier A, Ma L, et al. Evidence in the Legionella
pneumophila genome for exploitation of host cell
functions and high genome plasticity. Nat Genet
2004; 36:1165-73; PMID:15467720; http://dx.doi.
org/10.1038/ng1447
54. Chien M, Morozova I, Shi S, Sheng H, Chen J,
Gomez SM, et al. The genomic sequence of the
accidental pathogen Legionella pneumophila. Science
2004; 305:1966-8; PMID:15448271; http://dx.doi.
org/10.1126/science.1099776
55. Habyarimana F, Al-Khodor S, Kalia A, Graham JE,
Price CT, Garcia MT, et al. Role for the Ankyrin
eukaryotic-like genes of Legionella pneumophila in
parasitism of protozoan hosts and human macro-
phages. Environ Microbiol 2008; 10:1460-74;
PMID:18279343; http://dx.doi.org/10.1111/j.1462-
2920.2007.01560.x
56. Al-Khodor S, Price CT, Kalia A, Abu Kwaik Y.
Functional diversity of ankyrin repeats in micro-
bial proteins. Trends Microbiol 2010; 18:132-9;
PMID:19962898; http://dx.doi.org/10.1016/j.
tim.2009.11.004
57. Price CTD, Jones SC, Amundson KE, Kwaik YA.
Host-mediated post-translational prenylation of novel
dot/icm-translocated effectors of legionella pneumoph-
ila. Front Microbiol 2010; 1:131; PMID:21687755;
http://dx.doi.org/10.3389/fmicb.2010.00131
58. Al-Quadan T, Price CT, London N, Schueler-Furman
O, AbuKwaik Y. Anchoring of bacterial effectors to
host membranes through host-mediated lipidation by
prenylation: a common paradigm. Trends Microbiol
2011; 19:573-9; PMID:21983544; http://dx.doi.
org/10.1016/j.tim.2011.08.003
59. Faucher SP, Mueller CA, Shuman HA. Legionella
Pneumophila Transcriptome during Intracellular
Multiplication in Human Macrophages. Front
Microbiol 2011; 2:60; PMID:21747786; http://dx.doi.
org/10.3389/fmicb.2011.00060
60. Brüggemann H, Hagman A, Jules M, Sismeiro O,
Dillies MA, Gouyette C, et al. Virulence strategies for
infecting phagocytes deduced from the in vivo tran-
scriptional program of Legionella pneumophila. Cell
Microbiol 2006; 8:1228-40; PMID:16882028; http://
dx.doi.org/10.1111/j.1462-5822.2006.00703.x
61. Al-Khodor S, Kalachikov S, Morozova I, Price CT,
Abu Kwaik Y. The PmrA/PmrB two-component sys-
tem of Legionella pneumophila is a global regulator
required for intracellular replication within macro-
phages and protozoa. Infect Immun 2009; 77:374-
86; PMID:18936184; http://dx.doi.org/10.1128/
IAI.01081-08
62. Amer AO, Swanson MS. A phagosome of one’s own: a
microbial guide to life in the macrophage. Curr Opin
Microbiol 2002; 5:56-61; PMID:11834370; http://
dx.doi.org/10.1016/S1369-5274(02)00286-2
314 Virulence Volume 4 Issue 4
63. Molofsky AB, Swanson MS. Differentiate to thrive:
lessons from the Legionella pneumophila life cycle. Mol
Microbiol 2004; 53:29-40; PMID:15225301; http://
dx.doi.org/10.1111/j.1365-2958.2004.04129.x
64. Joshi AD, Swanson MS. Comparative analysis of
Legionella pneumophila and Legionella micdadei
virulence traits. Infect Immun 1999; 67:4134-42;
PMID:10417184.
65. Swanson MS, Hammer BK. Legionella pneumophila
pathogesesis: a fateful journey from amoebae to mac-
rophages. Annu Rev Microbiol 2000; 54:567-613;
PMID:11018138; http://dx.doi.org/10.1146/annurev.
micro.54.1.567
66. Rasis M, Segal G. The LetA-RsmYZ-CsrA regula-
tory cascade, together with RpoS and PmrA, post-
transcriptionally regulates stationary phase activation
of Legionella pneumophila Icm/Dot effectors. Mol
Microbiol 2009; 72:995-1010; PMID:19400807;
http://dx.doi.org/10.1111/j.1365-2958.2009.06705.x
67. Dalebroux ZD, Yagi BF, Sahr T, Buchrieser C,
Swanson MS. Distinct roles of ppGpp and DksA in
Legionella pneumophila differentiation. Mol Microbiol
2010; 76:200-19; PMID:20199605; http://dx.doi.
org/10.1111/j.1365-2958.2010.07094.x
68. Molofsky AB, Swanson MS. Legionella pneumophila
CsrA is a pivotal repressor of transmission traits and
activator of replication. Mol Microbiol 2003; 50:445-
61; PMID:14617170; http://dx.doi.org/10.1046/
j.1365-2958.2003.03706.x
69. Bachman MA, Swanson MS. RpoS co-operates
with other factors to induce Legionella pneumoph-
ila virulence in the stationary phase. Mol Microbiol
2001; 40:1201-14; PMID:11401723; http://dx.doi.
org/10.1046/j.1365-2958.2001.02465.x
70. Abu-Zant A, Asare R, Graham JE, Abu Kwaik Y. Role
for RpoS but not RelA of Legionella pneumophila in
modulation of phagosome biogenesis and adaptation
to the phagosomal microenvironment. Infect Immun
2006; 74:3021-6; PMID:16622243; http://dx.doi.
org/10.1128/IAI.74.5.3021-3026.2006
71. Sahr T, Brüggemann H, Jules M, Lomma M, Albert-
Weissenberger C, Cazalet C, et al. Two small ncRNAs
jointly govern virulence and transmission in Legionella
pneumophila. Mol Microbiol 2009; 72:741-62;
PMID:19400772; http://dx.doi.org/10.1111/j.1365-
2958.2009.06677.x
72. Al-Khodor S, Al-Quadan T, Abu Kwaik Y. Temporal
and differential regulation of expression of the eukary-
otic-like ankyrin effectors of Legionella pneumophila.
Environ Microbiol Rep 2010; 2:677-84; http://dx.doi.
org/10.1111/j.1758-2229.2010.00159.x
73. Garduño RA, Garduño E, Hiltz M, Hoffman PS.
Intracellular growth of Legionella pneumophila gives
rise to a differentiated form dissimilar to station-
ary-phase forms. Infect Immun 2002; 70:6273-
83; PMID:12379706; http://dx.doi.org/10.1128/
IAI.70.11.6273-6283.2002
74. Faulkner G, Garduño RA. Ultrastructural analysis of
differentiation in Legionella pneumophila. J Bacteriol
2002; 184:7025-41; PMID:12446652; http://dx.doi.
org/10.1128/JB.184.24.7025-7041.2002
75. Greub G, Raoult D. Morphology of Legionella pneu-
mophila according to their location within Hartmanella
vermiformis. Res Microbiol 2003; 154:619-21;
PMID:14596898; http://dx.doi.org/10.1016/j.res-
mic.2003.08.003
76. Price CT, Kwaik YA. Exploitation of Host
Polyubiquitination Machinery through Molecular
Mimicry by Eukaryotic-Like Bacterial F-Box Effectors.
Front Microbiol 2010; 1:122; PMID:21687758;
http://dx.doi.org/10.3389/fmicb.2010.00122
77. Al-Quadan T, Kwaik YA. Molecular Characterization of
Exploitation of the Polyubiquitination and Farnesylation
Machineries of Dictyostelium Discoideum by the
AnkB F-Box Effector of Legionella Pneumophila. Front
Microbiol 2011; 2:23; PMID:21687415; http://dx.doi.
org/10.3389/fmicb.2011.00023
78. Price CT, Al-Quadan T, Santic M, Jones SC, Abu
Kwaik Y. Exploitation of conserved eukaryotic host
cell farnesylation machinery by an F-box effector of
Legionella pneumophila. J Exp Med 2010; 207:1713-
26; PMID:20660614; http://dx.doi.org/10.1084/
jem.20100771
79. Price CT, Al-Khodor S, Al-Quadan T, Abu Kwaik
Y. Indispensable role for the eukaryotic-like ankyrin
domains of the ankyrin B effector of Legionella pneu-
mophila within macrophages and amoebae. Infect
Immun 2010; 78:2079-88; PMID:20194593; http://
dx.doi.org/10.1128/IAI.01450-09
80. Price CT, Al-Quadan T, Santic M, Rosenshine I, Abu
Kwaik Y. Host proteasomal degradation generates
amino acids essential for intracellular bacterial growth.
Science 2011; 334:1553-7; PMID:22096100; http://
dx.doi.org/10.1126/science.1212868
81. Eylert E, Herrmann V, Jules M, Gillmaier N, Lautner
M, Buchrieser C, et al. Isotopologue profiling of
Legionella pneumophila: role of serine and glucose
as carbon substrates. J Biol Chem 2010; 285:22232-
43; PMID:20442401; http://dx.doi.org/10.1074/jbc.
M110.128678
82. Suraweera A, Münch C, Hanssum A, Bertolotti A.
Failure of amino acid homeostasis causes cell death fol-
lowing proteasome inhibition. Mol Cell 2012; 48:242-
53; PMID:22959274; http://dx.doi.org/10.1016/j.
molcel.2012.08.003
83. de Felipe KS, Pampou S, Jovanovic OS, Pericone CD,
Ye SF, Kalachikov S, et al. Evidence for acquisition of
Legionella type IV secretion substrates via interdomain
horizontal gene transfer. J Bacteriol 2005; 187:7716-
26; PMID:16267296; http://dx.doi.org/10.1128/
JB.187.22.7716-7726.2005
84. Franco IS, Shuman HA, Charpentier X. The perplexing
functions and surprising origins of Legionella pneu-
mophila type IV secretion effectors. Cell Microbiol
2009; 11:1435-43; PMID:19563462; http://dx.doi.
org/10.1111/j.1462-5822.2009.01351.x
85. Amer AO. Modulation of caspases and their non-
apoptotic functions by Legionella pneumophila. Cell
Microbiol 2010; 12:140-7; PMID:19863553; http://
dx.doi.org/10.1111/j.1462-5822.2009.01401.x
86. Molmeret M, Zink SD, Han L, Abu-Zant A, Asari R,
Bitar DM, et al. Activation of caspase-3 by the Dot/
Icm virulence system is essential for arrested biogen-
esis of the Legionella-containing phagosome. Cell
Microbiol 2004; 6:33-48; PMID:14678329; http://
dx.doi.org/10.1046/j.1462-5822.2003.00335.x
87. Abu-Zant A, Santic M, Molmeret M, Jones S, Helbig
J, Abu Kwaik Y. Incomplete activation of macro-
phage apoptosis during intracellular replication of
Legionella pneumophila. Infect Immun 2005; 73:5339-
49; PMID:16113249; http://dx.doi.org/10.1128/
IAI.73.9.5339-5349.2005
88. Abu-Zant A, Jones S, Asare R, Suttles J, Price C,
Graham J, et al. Anti-apoptotic signalling by the
Dot/Icm secretion system of L. pneumophila. Cell
Microbiol 2007; 9:246-64; PMID:16911566; http://
dx.doi.org/10.1111/j.1462-5822.2006.00785.x
89. Gao LY, Abu Kwaik Y. Activation of caspase-3 in
Legionella pneumophila-induced apoptosis. Infect
Immun 1999; 67:4886-94; PMID:10456945
90. Abdelaziz DH, Gavrilin MA, Akhter A, Caution K,
Kotrange S, Khweek AA, et al. Apoptosis-associated
speck-like protein (ASC) controls Legionella pneu-
mophila infection in human monocytes. J Biol Chem
2011; 286:3203-8; PMID:21097506; http://dx.doi.
org/10.1074/jbc.M110.197681
91. Abdelaziz DH, Gavrilin MA, Akhter A, Caution K,
Kotrange S, Khweek AA, et al. Asc-dependent and
independent mechanisms contribute to restriction of
legionella pneumophila infection in murine macro-
phages. Front Microbiol 2011; 2:18; PMID:21713115;
http://dx.doi.org/10.3389/fmicb.2011.00018
92. Akhter A, Caution K, Abu Khweek A, Tazi M,
Abdulrahman BA, Abdelaziz DH, et al. Caspase-11 pro-
motes the fusion of phagosomes harboring pathogenic
bacteria with lysosomes by modulating actin polymer-
ization. Immunity 2012; 37:35-47; PMID:22658523;
http://dx.doi.org/10.1016/j.immuni.2012.05.001
93. Pagnier I, Raoult D, La Scola B. Isolation and iden-
tification of amoeba-resisting bacteria from water
in human environment by using an Acanthamoeba
polyphaga co-culture procedure. Environ Microbiol
2008; 10:1135-44; PMID:18279351; http://dx.doi.
org/10.1111/j.1462-2920.2007.01530.x
94. Sauer JD, Bachman MA, Swanson MS. The phago-
somal transporter A couples threonine acquisition to
differentiation and replication of Legionella pneu-
mophila in macrophages. Proc Natl Acad Sci U S A
2005; 102:9924-9; PMID:15998735; http://dx.doi.
org/10.1073/pnas.0502767102
... Some bacteria have been found to establish intracellular niches within both human cells and unicellular protists, which are evolutionarily distant hosts. Notable examples include Legionella pneumophila [8,9], Pseudomonas aeruginosa [10,11], Francisella novicida [12], Coxiella burnetii [13], and Mycobacterium avium [14]. Ciliates constitute a highly diverse group of protists, with over 8000 documented free-living species [15], and their evolutionary history can be traced back to approximately 1.1 billion years [16]. ...
Article
Full-text available
Background The eukaryotic-bacterial symbiotic system plays an important role in various physiological, developmental, and evolutionary processes. However, our current understanding is largely limited to multicellular eukaryotes without adequate consideration of diverse unicellular protists, including ciliates. Results To investigate the bacterial profiles associated with unicellular organisms, we collected 246 ciliate samples spanning the entire Ciliophora phylum and conducted single-cell based metagenome sequencing. This effort has yielded the most extensive collection of bacteria linked to unicellular protists to date. From this dataset, we identified 883 bacterial species capable of cohabiting with ciliates, unveiling the genomes of 116 novel bacterial cohabitants along with 7 novel archaeal cohabitants. Highlighting the intimate relationship between ciliates and their cohabitants, our study unveiled that over 90% of ciliates coexist with bacteria, with individual hosts fostering symbiotic relationships with multiple bacteria concurrently, resulting in the observation of seven distinct symbiotic patterns among bacteria. Our exploration of symbiotic mechanisms revealed the impact of host digestion on the intracellular diversity of cohabitants. Additionally, we identified the presence of eukaryotic-like proteins in bacteria as a potential contributing factor to their resistance against host digestion, thereby expanding their potential host range. Conclusions As the first large-scale analysis of prokaryotic associations with ciliate protists, this study provides a valuable resource for future research on eukaryotic-bacterial symbioses. 89id7Mc_Y_Zgt6kKNc3JJ4Video Abstract
... For the intracellular replicating bacterium L. pneumophila, all the mentioned pathways play a role in egress. L. pneumophila is a gram-negative bacterium that is naturally parasitic to environmental amoebae (Richards et al., 2013). However, through the inhalation of droplets derived from contaminated water systems, L. pneumophila can be transmitted to humans and cause a severe pneumonic infection, called Legionnaire's disease (Cunha et al., 2016). ...
Article
Full-text available
The phenomenon of host cell escape exhibited by intracellular pathogens is a remarkably versatile occurrence, capable of unfolding through lytic or non‐lytic pathways. Among these pathogens, the bacterium Legionella pneumophila stands out, having adopted a diverse spectrum of strategies to disengage from their host cells. A pivotal juncture that predates most of these host cell escape modalities is the initial escape from the intracellular compartment. This critical step is increasingly supported by evidence suggesting the involvement of several secreted pathogen effectors, including lytic proteins. In this intricate landscape, L. pneumophila emerges as a focal point for research, particularly concerning secreted phospholipases. While nestled within its replicative vacuole, the bacterium deftly employs both its type II (Lsp) and type IVB (Dot/Icm) secretion systems to convey phospholipases into either the phagosomal lumen or the host cell cytoplasm. Its repertoire encompasses numerous phospholipases A (PLA), including three enzymes—PlaA, PlaC, and PlaD—bearing the GDSL motif. Additionally, there are 11 patatin‐like phospholipases A as well as PlaB. Furthermore, the bacterium harbors three extracellular phospholipases C (PLCs) and one phospholipase D. Within this comprehensive review, we undertake an exploration of the pivotal role played by phospholipases in the broader context of phagosomal and host cell egress. Moreover, we embark on a detailed journey to unravel the established and potential functions of the secreted phospholipases of L. pneumophila in orchestrating this indispensable process.
... Many bacterial secretion systems have profound importance in the formation of biofilms and are also involved in predation resistance. Survival of L. pneumophilia inside amoeba and macr opha ge depends on the Dot/Icm type IV secretion system that injects a large number effector proteins into the host cell (Richards et al. 2013 ). Pseudomonas aeruginosa type III secretion systems are involved in killing of biofilm-associated amoeba (Matz et al. 2008 ). ...
Article
Protozoa are eukaryotic organisms that play a crucial role in nutrient cycling and maintaining balance in the food web. Predation, symbiosis and parasitism are three types of interactions between protozoa and bacteria. However, not all bacterial species are equally susceptible to protozoan predation as many are capable of defending against predation in numerous ways and may even establish either a symbiotic or parasitic life-style. Biofilm formation is one such mechanism by which bacteria can survive predation. Structural and chemical components of biofilms enhance resistance to predation compared to their planktonic counterparts. Predation on biofilms gives rise to phenotypic and genetic heterogeneity in prey that leads to trade-offs in virulence in other eukaryotes. Recent advances, using molecular and genomics techniques, allow us to generate new information about the interactions of protozoa and biofilms of prey bacteria. This review presents the current state of the field on impacts of protozoan predation on biofilms. We provide an overview of newly gathered insights into (i) molecular mechanisms of predation resistance in biofilms, (ii) phenotypic and genetic diversification of prey bacteria, and (iii) evolution of virulence as a consequence of protozoan predation on biofilms.
... In addition, a specific interaction is known between Legionella pneumoniae and amoebae. The bacteria can survive in the amoeba vacuole and transfer genes between hosts in the vacuole; as an example, Dot/Icm-translocated effector genes are acquired through inter-kingdom HGT between Legionella pneumophila and host amoebae and ciliates [53]. The soil amoeba Dictyostelium is used as a model host for L. pneumoniae [54]. ...
Article
Full-text available
There is much evidence that the environment is a reservoir of antibiotic resistance genes (ARGs). For ARG latency and stability, environmental factors should play a role since ARGs are transferred among bacterial species, which results in their evolution and dissemination. Recent findings have expanded the novel mechanisms of horizontal gene transfer (HGT) beyond classically accepted HGT mechanisms such as conjugation, transformation, and transduction. In these HGT processes, environmental factors directly or indirectly affect the transfer rate and mechanism. Here, we focus on the effect of protists that affect HGT, because HGT is regulated among the microbial community, in which protist grazing is one factor that enhances HGT. Protist grazing eliminates planktonic bacteria. However, it is reported that environmental DNA release and HGT in protist vacuoles are increased by the grazing, although the effect is not uniform and depends on the environmental conditions. Biofilms protect bacteria and accelerate HGT. In these processes, quorum sensing and organic matter contribute to HGT. Although HGT occurs between bacterial cells, other microorganisms such as protists should be recognized as factors relating to HGT. We should pay attention to microbial ecosystem when consider ARG risk from water environment in terms of “one health” aspect.
... Presumably, ProA anchors itself via a putative farnesylation domain at the Cterminus and cleaves host factors to influence downstream signaling pathways. It is also conceivable that ProA contributes to the acquisition of nutrients for the proliferation of L. pneumophila, thus, supporting known mechanisms of the virulence factor AnkB, which exploits the ubiquitination machinery of the host cell (Richards et al., 2013;Truchan et al., 2017;White et al., 2018). ...
Article
Full-text available
The pathogenicity of L. pneumophila, the causative agent of Legionnaires’ disease, depends on an arsenal of interacting proteins. Here we describe how surface-associated and secreted virulence factors of this pathogen interact with each other or target extra- and intracellular host proteins resulting in host cell manipulation and tissue colonization. Since progress of computational methods like AlphaFold, molecular dynamics simulation, and docking allows to predict, analyze and evaluate experimental proteomic and interactomic data, we describe how the combination of these approaches generated new insights into the multifaceted “protein sociology” of the zinc metalloprotease ProA and the peptidyl-prolyl cis/trans isomerase Mip (macrophage infectivity potentiator). Both virulence factors of L. pneumophila interact with numerous proteins including bacterial flagellin (FlaA) and host collagen, and play important roles in virulence regulation, host tissue degradation and immune evasion. The recent progress in protein-ligand analyses of virulence factors suggests that machine learning will also have a beneficial impact in early stages of drug discovery.
... In general, protists harbour abundant and diverse bacterial endosymbionts (i.e., intracellular bacteria), including several representatives of lineages that also encompass human pathogens, such as Rickettsiales, Legionellales, and Chlamydiae [5][6][7]. Some protists are even able to host, at least temporarily, human pathogenic bacteria such as Legionella [8,9], thus potentially representing natural reservoirs. Those features have led several authors to infer a role of protists as melting pots for the evolution of potentially pathogenic bacteria, able to ...
Article
Full-text available
Symbiotic associations between bacteria and ciliate protists are rather common. In particular, several cases were reported involving bacteria of the alphaproteobacterial lineage Rickettsiales, but the diversity, features, and interactions in these associations are still poorly understood. In this work, we characterized a novel ciliate protist strain originating from Brazil and its associated Rickettsiales endosymbiont by means of live and ultrastructural observations, as well as molecular phylogeny. Though with few morphological peculiarities, the ciliate was found to be phylogenetically affiliated with Pseudokeronopsis erythrina, a euryhaline species, which is consistent with its origin from a lagoon with significant spatial and seasonal salinity variations. The bacterial symbiont was assigned to “Candidatus Trichorickettsia mobilis subsp. hyperinfectiva”, being the first documented case of a Rickettsiales associated with urostylid ciliates. It resided in the host cytoplasm and bore flagella, similarly to many, but not all, conspecifics in other host species. These findings highlight the ability of “Candidatus Trichorickettsia” to infect multiple distinct host species and underline the importance of further studies on this system, in particular on flagella and their regulation, from a functional and also an evolutionary perspective, considering the phylogenetic proximity with the well-studied and non-flagellated Rickettsia.
Article
Full-text available
There has been considerable discussion regarding the environmental life cycle of Legionella pneumophila and its virulence potential in natural and man-made water systems. On the other hand, the bacterium's morphogenetic mechanisms within host cells (amoeba and macrophages) have been well documented and are linked to its ability to transition from a non-virulent, replicative state to an infectious, transmissive state. Although the morphogenetic mechanisms associated with the formation and detachment of the L. pneumophila biofilm have also been described, the capacity of the bacteria to multiply extracellularly is not generally accepted. However, several studies have shown genetic pathways within the biofilm that resemble intracellular mechanisms. Understanding the functionality of L. pneumophila cells within a biofilm is fundamental for assessing the ecology and evaluating how the biofilm architecture influences L. pneumophila survival and persistence in water systems. This manuscript provides an overview of the biphasic cycle of L. pneumophila and its implications in associated intracellular mechanisms in amoeba. It also examines the molecular pathways and gene regulation involved in L. pneumophila biofilm formation and dissemination. A holistic analysis of the transcriptional activities in L. pneumophila biofilms is provided, combining the information of intracellular mechanisms in a comprehensive outline. Furthermore, this review discusses the techniques that can be used to study the morphogenetic states of the bacteria within biofilms, at the single cell and population levels.
Article
Full-text available
Legionella pneumophila is a gram‐negative bacteria found in natural and anthropogenic aquatic environments such as evaporative cooling towers, where it reproduces as an intracellular parasite of cohabiting protozoa. If L. pneumophila is aerosolized and inhaled by a susceptible person, bacteria may colonize their alveolar macrophages causing the opportunistic pneumonia Legionnaires' disease. L. pneumophila utilizes an elaborate regulatory network to control virulence processes such as the Dot/Icm Type IV secretion system and effector repertoire, responding to changing nutritional cues as their host becomes depleted. The bacteria subsequently differentiate to a transmissive state that can survive in the environment until a replacement host is encountered and colonized. In this review, we discuss the lifecycle of L. pneumophila and the molecular regulatory network that senses nutritional depletion via the stringent response, a link to stationary phase‐like metabolic changes via alternative sigma factors, and two‐component systems that are homologous to stress sensors in other pathogens, to regulate differentiation between the intracellular replicative phase and more transmissible states. Together, we highlight how this prototypic intracellular pathogen offers enormous potential in understanding how molecular mechanisms enable intracellular parasitism and pathogenicity.
Article
Full-text available
Anthropogenic eutrophication of lakes threatens their homeostasis and carries an increased risk of development of potentially pathogenic microorganisms. In this paper we show how eutrophication affects seasonal changes in the taxonomic structure of bacterioplankton and whether these changes are associated with the relative abundance of pathogenic bacteria of the genera Legionella and Aeromonas. The subject of the study was a unique system of interconnected lakes in northern Poland (Great Masurian Lakes system), characterized by the presence of eutrophic gradient. We found that the taxonomic structure of the bacterial community in eutrophic lakes was significantly season dependent. No such significant seasonal changes were observed in meso-eutrophic lakes. We found that there is a specific taxonomic composition of bacteria associated with the occurrence of Legionella spp. The highest positive significant correlations were found for families Pirellulaceae, Mycobacteriaceae and Gemmataceae. The highest negative correlations were found for the families Sporichthyaceae, Flavobacteriaceae, the uncultured families of class Verrucomicrobia and Chitinophagaceae. We used also an Automatic Neural Network model to estimate the relative abundance of Legionella spp. based on the relative abundance of dominant bacterial families. In the case of Aeromonas spp. we did not find a clear relationship with bacterial communities inhabiting lakes of different trophic state. Our research has shown that anthropogenic eutrophication causes significant changes in the taxonomic composition of lake bacteria and contributes to an increase in the proportion of potentially pathogenic Legionella spp.
Article
Full-text available
When Legionella pneumophila, which causes Legionnaires' disease, infects humans or amoeba, the bacteria promote host proteasomal degradation to obtain amino acids for growth. The Dot/Icm type IV secretion apparatus of L. pneumophila injects host cells with about 300 different effector proteins, close to 10% of its genomic coding capacity. When Legionella enters an intracellular vacuole, called the LCV, it is soon decorated with polyubiquitinated proteins, a process that depends on the AnkB Dot/Icm-translocated effector being anchored into the LCV membrane. Proteasome-mediated degradation of such polyubiquitinated proteins generates amino acids needed to power replication of L. pneumophila. By synchronizing amino acid auxotrophy with its amoeba hosts, L. pneumophila may be protected within starved amoebic cysts until nutrients are again abundant.
Article
Full-text available
The ubiquitin-proteasome system targets many cellular proteins for degradation and thereby controls most cellular processes. Although it is well established that proteasome inhibition is lethal, the underlying mechanism is unknown. Here, we show that proteasome inhibition results in a lethal amino acid shortage. In yeast, mammalian cells, and flies, the deleterious consequences of proteasome inhibition are rescued by amino acid supplementation. In all three systems, this rescuing effect occurs without noticeable changes in the levels of proteasome substrates. In mammalian cells, the amino acid scarcity resulting from proteasome inhibition is the signal that causes induction of both the integrated stress response and autophagy, in an unsuccessful attempt to replenish the pool of intracellular amino acids. These results reveal that cells can tolerate protein waste, but not the amino acid scarcity resulting from proteasome inhibition.
Article
Since Legionella pneumophila is an intracellular pathogen, entry into and replication within host cells are thought to be critical to its ability to cause disease. L, pneumophila grown in one of its environmental hosts, Acanthamoeba castellanii, is phenotypically different from L. pneumophila grown on standard laboratory medium (BCYE agar). Although amoeba-grown L. pneumophila displays enhanced entry into monocytes compared to BCYE-grown bacteria, the mechanisms of entry used and the effects on virulence have not been examined. To explore whether amoeba-grown L. pneumophila differs from BCYE-grown L. pneumophila in these characteristics, we examined entry into monocytes, replication in activated macrophages, and virulence in mice. Entry of amoeba-grown L. pneumophila into monocytes occurred more frequently by coiling phagocytosis, was less affected by complement opsonization, and was Less sensitive to microtubule and microfilament inhibitors than was entry of BCYE-grown bacteria. In addition, amoeba-grown L. pneumophila displays increased replication in monocytes and is more virulent in A/J, C57BL/6 Beige, and C57BL/6 mice. These data demonstrate for the first time that the intra-amoebal growth environment affects the entry mechanisms and virulence of L. pneumophila.
Article
Legionella pneumophila, the causative agent of Legionnaires' pneumonia, replicates within alveolar macrophages by preventing phagosome-lysosome fusion. Here, a large number of mutants called dot (defective for organelle trafficking) that were unable to replicate intracellularly because of an inability of the bacteria to alter the endocytic pathway of macrophages were isolated. The dot virulence genes encoded a large putative membrane complex that functioned as a secretion system that was able to transfer plasmid DNA from one cell to another.
Article
Upon transition from the exponential (E) to the post-exponential phase (PE) of growth, Legionella pneumophila undergoes a phenotypic modulation from a replicative to a highly infectious form. This transition requires a delicate regulatory cascade that is triggered to induce expression of various virulence-related genes. We have recently characterized eleven L. pneumophila eukaryotic-like ankyrin effectors (Ank) shared between the four sequenced genomes of L. pneumophila. The AnkB effector recruits polyubiquitinated proteins to the Legionella-containing vacuole (LCV). It is not known whether expression of the ank genes is regulated by various regulators triggered at the PE phase and whether this regulation is essential for function. Here we show that temporal and differential regulation of the ank genes is mediated by RelA, the enhancer protein LetE, and the two component systems LetA/S and PmrA/B. Consistent with the expression of ankB at the PE phase, we show that bacteria grown to the PE but not the E phase recruit polyubiquitinated proteins to the LCV within Acanthamoeba in an AnkB-dependant mechanism. We conclude that the genes encoding the eukaryotic-like Ank effectors of L. pneumophila are temporally and spatially regulated at the PE phase.
Article
The soil amoeba Dictyostelium discoideum is a haploid eukaryote that, upon starvation, aggregates and enters a developmental cycle to produce fruiting bodies. In this study, we infected single-cell stages of D. discoideum with different Legionella species. Intracellular growth of Legionella in this new host system was compared with their growth in the natural host Acanthamoeba castellanii. Transmission electron microscopy of infected D. discoideum cells revealed that legionellae reside within the phagosome. Using confocal microscopy, it was observed that replicating, intracellular, green fluorescent protein (GFP)-tagged legionellae rarely co-localized with fluorescent antibodies directed against the lysosomal protein DdLIMP of D. discoideum. This indicates that the bacteria inhibit the fusion of phagosomes and lysosomes in this particular host system. In addition, Legionella infection of D. discoideum inhibited the differentiation of the host into the multicellular fruiting stage. Co-culture studies with profilin-minus D. discoideum mutants and Legionella resulted in higher rates of infection when compared with infections of wild-type amoebae. Because the amoebae are amenable to genetic manipulation as a result of their haploid genome and because a number of cellular markers are available, we show for the first time that D. discoideum is a valuable model system for studying intracellular pathogenesis of microbial pathogens.
Article
The ability of Legionella pneumophila to cause legionnaires' disease is dependent on its capacity to replicate within cells in the alveolar spaces. The bacteria kill mammalian cells in two phases: induction of apoptosis during the early stages of infection, followed by an independent and rapid necrosis during later stages of the infection, mediated by a pore-forming activity. In the environment, L. pneumophila is a parasite of protozoa. The molecular mechanisms by which L. pneumophila kills the protozoan cells, after their exploitation for intracellular proliferation, are not known. In an effort to decipher these mechanisms, we have examined induction of both apoptosis and necrosis in the protozoan Acanthamoeba polyphaga upon infection by L. pneumophila. Our data show that, although A. polyphaga undergoes apoptosis following treatment with actinomycin D, L. pneumophila does not induce apoptosis in these cells. Instead, intracellular L. pneumophila induces necrotic death in A. polyphaga, which is mediated by the pore-forming activity. Mutants of L. pneumophila defective in expression of the pore-forming activity are indistinguishable from the parental strain in intracellular replication within A. polyphaga. The parental strain bacteria cause necrosis-mediated lysis of all the A. polyphaga cells within 48 h after infection, and all the intracellular bacteria are released into the tissue culture medium. In contrast, all cells infected by the mutants remain intact, and the intracellular bacteria are ‘trapped’ within A. polyphaga after the termination of intracellular replication. Failure to exit the host cell after termination of intracellular replication results in a gradual decline in the viability of the mutant strain bacteria within A. polyphaga starting 48 h after infection. Our data show that the pore-forming activity of L. pneumophila is not required for intracellular bacterial replication within A. polyphaga but is required for killing and exiting the protozoan host upon termination of intracellular replication.
Article
A free-living amoeba identified as Hartmannella vermiformis was isolated from a water sample obtained during an investigation of nosocomial legionellosis. Hartmannella vermiformis is known to support the intracellular multiplication of Legionella pneumophila. This strain of H. vermiformis, designated CDC-19, was cloned and established in axenic culture to develop a model for the study of the pathogenicity of legionellae. Isoenzyme patterns of axenically-cultivated strain CDC-19 were compared with two strains of H. vermiformis derived from the type strain, one axenic (ATCC 50236) and the other grown in the presence of bacteria (ATCC 30966). Enzyme patterns suggested that all three strains are assignable to the species H. vermiformis. Axenic H. vermiformis strain CDC-19 has been deposited with the American Type Culture Collection (ATCC 50237) and should prove useful in the study of protozoan-bacterial interaction.
Article
Legionella pneumophila replicates within amoebae and macrophages and causes the severe pneumonia Legionnaires' disease. When broth cultures enter the post-exponential growth (PE) phase or experience amino acid limitation, L. pneumophila accumulates the stringent response signal (p)ppGpp and expresses traits likely to promote transmission to a new phagocyte. The hypothesis that a stringent response mechanism regulates L. pneumophila virulence was bolstered by our finding that the avirulent mutant Lp120 contains an internal deletion in the gene encoding the stationary phase sigma factor RpoS. To test directly whether RpoS co-ordinates virulence with stationary phase, isogenic wild-type, rpoS-120 and rpoS null mutant strains were constructed and analysed. PE phase L. pneumophila became cytotoxic by an RpoS-independent pathway, but their sodium sensitivity and maximal expression of flagellin required RpoS. Likewise, full induction of sodium sensitivity by experimentally induced (p)ppGpp synthesis required RpoS. To replicate efficiently in macrophages, L. pneumophila used both RpoS-dependent and -independent pathways. Like those containing the dotA type IV secretory apparatus mutant, phagosomes harbouring either rpoS or dotA rpoS mutants rapidly acquired the late endosomal protein LAMP-1, but not the lysosomal marker Texas red–ovalbumin. Together, the data support a model in which RpoS co-operates with other regulators to induce L. pneumophila virulence in the PE phase.