+ All Categories
Home > Documents > Chlamydial biology and its associated virulence blockers

Chlamydial biology and its associated virulence blockers

Date post: 05-Jan-2017
Category:
Upload: daisy
View: 218 times
Download: 1 times
Share this document with a friend
16
2014 http://informahealthcare.com/mby ISSN: 1040-841X (print), 1549-7828 (electronic) Crit Rev Microbiol, 2014; 40(4): 313–328 ! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/1040841X.2012.726210 REVIEW ARTICLE Chlamydial biology and its associated virulence blockers Delphine S. Beeckman, Leentje De Puysseleyr*, Kristien De Puysseleyr*, and Daisy Vanrompay Department of Molecular Biotechnology, Faculty of Bioscience Engineering, Ghent University, Coupure Links 653, B-9000 Ghent, Belgium Abstract Chlamydiales are obligate intracellular parasites of eukaryotic cells. They can be distinguished from other Gram-negative bacteria through their characteristic developmental cycle, in addition to special biochemical and physical adaptations to subvert the eukaryotic host cell. The host spectrum includes humans and other mammals, fish, birds, reptiles, insects and even amoeba, causing a plethora of diseases. The first part of this review focuses on the specific chlamydial infection biology and metabolism. As resistance to classical antibiotics is emerging among Chlamydiae as well, the second part elaborates on specific compounds and tools to block chlamydial virulence traits, such as adhesion and internalization, Type III secretion and modulation of gene expression. Keywords Antibiotic resistance, chlamydia, infection biology, virulence, virulence blockers History Received 16 May 2012 Revised 26 August 2012 Accepted 29 August 2012 Introduction to Chlamydiaceae Micro-organisms in the family of the Chlamydiaceae are obligate intracellular pathogens of both mammals and birds. The different species in this family infect many hosts, with variable tissue tropism causing a multiplicity of acute and chronic diseases, from sexually transmitted infertility, to trachoma and respiratory and cardiovascular diseases. Chlamydia (C.) trachomatis and Chlamydia (C.) pneumo- niae are the most common chlamydial pathogens in humans, whereas the other chlamydial species mainly occur in other animals. In the so-called developed countries, C. trachomatis is the causative agent of the most frequent sexually transmitted disease (STD), leading to pelvic inflammatory disease, infertility and possibly ectopic pregnancy. Moreover, some C. trachomatis serovars are known to induce trachoma, the leading cause of infectious blindness in developing countries. Ten percent of the pneumonia and 5% of bronchitis and sinusitis cases in adults is attributable to a C. pneumoniae infection and chronic infection could contribute to athero- sclerosis (Campbell & Kuo, 2004, Belland et al., 2004). As proven pathogens of vertebrates, Chlamydiaceae cause reproductive, respiratory, cardiovascular, gastrointestinal or systemic disease, as well as conjunctivitis, arthritis and encephalitis in both live stock, companion and wild animals (Everett, 2000). Chlamydia psittaci primarily infects birds, predominantly impacting on commercial turkey and duck farms, but is also of public health importance due to its zoonotic nature (Beeckman et al., 2009). Other highly prevalent zoonotic species in animals include C. abortus (mainly in sheep and goats), C. suis (endemic in pigs) and C. felis (cats). Due to spatial limitations and for reasons of clarity, the description of the chlamydial infection biology and possible ways to combat the infection will focus on aspects which are important for their survival and are likely to be implicated in virulence. Chlamydial biology Developmental forms During the unique biphasic life cycle (detailed description in section 0), at least two morphologically different structures can be observed: the infectious elementary bodies or EBs and the replicating reticulate bodies or RBs. An overview of the most important discriminative characteristics between these two developmental forms can be found in Table 1. In addition, intermediate bodies (IBs) can be observed during the maturation from EBs to RBs (Vanrompay et al., 1996). Elementary bodies are usually small, spherical, electron dense structures (Costerton et al., 1976; Longbottom & Coulter, 2003), characterized by a dense, eccentric core of condensed DNA and chromatin. The EB has a granular appearance due to the presence of 70S ribosomes. A lipid cytoplasmic membrane and a rigid outer membrane (both ~8 nm) with extensive disulphide bridging between cysteine and methionine residues of ‘Major Outer Membrane Protein’ (MOMP) and other sulphur amino acid rich outer membrane proteins surround the cytoplasm (Newhall & Jones, 1983). This high level of cross-linking possibly compensates for the low amounts of the usual cell wall strengthening substance peptidoglycan in the outer membrane. In this way, the spore- like EBs are osmotically more stable and less permeable than RBs, which allows them to survive up to several months outside the host cell (Longbottom & Coulter, 2003). *These authors contributed equally to this work. Address for correspondence: Daisy Vanrompay, Department of Molecular Biotechnology, Ghent University, Coupure Links 653, Ghent, B-9000, Belgium. E-mail: [email protected] Critical Reviews in Microbiology Downloaded from informahealthcare.com by RMIT University on 06/07/14 For personal use only.
Transcript
Page 1: Chlamydial biology and its associated virulence blockers

2014

http://informahealthcare.com/mbyISSN: 1040-841X (print), 1549-7828 (electronic)

Crit Rev Microbiol, 2014; 40(4): 313–328! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/1040841X.2012.726210

REVIEW ARTICLE

Chlamydial biology and its associated virulence blockers

Delphine S. Beeckman, Leentje De Puysseleyr*, Kristien De Puysseleyr*, and Daisy Vanrompay

Department of Molecular Biotechnology, Faculty of Bioscience Engineering, Ghent University, Coupure Links 653, B-9000 Ghent, Belgium

Abstract

Chlamydiales are obligate intracellular parasites of eukaryotic cells. They can be distinguishedfrom other Gram-negative bacteria through their characteristic developmental cycle, in additionto special biochemical and physical adaptations to subvert the eukaryotic host cell. The hostspectrum includes humans and other mammals, fish, birds, reptiles, insects and even amoeba,causing a plethora of diseases. The first part of this review focuses on the specific chlamydialinfection biology and metabolism. As resistance to classical antibiotics is emerging amongChlamydiae as well, the second part elaborates on specific compounds and tools to blockchlamydial virulence traits, such as adhesion and internalization, Type III secretion andmodulation of gene expression.

Keywords

Antibiotic resistance, chlamydia, infectionbiology, virulence, virulence blockers

History

Received 16 May 2012Revised 26 August 2012Accepted 29 August 2012

Introduction to Chlamydiaceae

Micro-organisms in the family of the Chlamydiaceae are

obligate intracellular pathogens of both mammals and birds.

The different species in this family infect many hosts, with

variable tissue tropism causing a multiplicity of acute and

chronic diseases, from sexually transmitted infertility, to

trachoma and respiratory and cardiovascular diseases.

Chlamydia (C.) trachomatis and Chlamydia (C.) pneumo-

niae are the most common chlamydial pathogens in humans,

whereas the other chlamydial species mainly occur in other

animals. In the so-called developed countries, C. trachomatis

is the causative agent of the most frequent sexually

transmitted disease (STD), leading to pelvic inflammatory

disease, infertility and possibly ectopic pregnancy. Moreover,

some C. trachomatis serovars are known to induce trachoma,

the leading cause of infectious blindness in developing

countries. Ten percent of the pneumonia and 5% of bronchitis

and sinusitis cases in adults is attributable to a C. pneumoniae

infection and chronic infection could contribute to athero-

sclerosis (Campbell & Kuo, 2004, Belland et al., 2004).

As proven pathogens of vertebrates, Chlamydiaceae cause

reproductive, respiratory, cardiovascular, gastrointestinal or

systemic disease, as well as conjunctivitis, arthritis and

encephalitis in both live stock, companion and wild animals

(Everett, 2000). Chlamydia psittaci primarily infects birds,

predominantly impacting on commercial turkey and duck

farms, but is also of public health importance due to its

zoonotic nature (Beeckman et al., 2009). Other highly

prevalent zoonotic species in animals include C. abortus

(mainly in sheep and goats), C. suis (endemic in pigs) and

C. felis (cats).

Due to spatial limitations and for reasons of clarity, the

description of the chlamydial infection biology and possible

ways to combat the infection will focus on aspects which are

important for their survival and are likely to be implicated in

virulence.

Chlamydial biology

Developmental forms

During the unique biphasic life cycle (detailed description in

section 0), at least two morphologically different structures

can be observed: the infectious elementary bodies or EBs and

the replicating reticulate bodies or RBs. An overview of the

most important discriminative characteristics between these

two developmental forms can be found in Table 1. In addition,

intermediate bodies (IBs) can be observed during the

maturation from EBs to RBs (Vanrompay et al., 1996).

Elementary bodies are usually small, spherical, electron

dense structures (Costerton et al., 1976; Longbottom &

Coulter, 2003), characterized by a dense, eccentric core of

condensed DNA and chromatin. The EB has a granular

appearance due to the presence of 70S ribosomes. A lipid

cytoplasmic membrane and a rigid outer membrane (both

~8 nm) with extensive disulphide bridging between cysteine

and methionine residues of ‘Major Outer Membrane Protein’

(MOMP) and other sulphur amino acid rich outer membrane

proteins surround the cytoplasm (Newhall & Jones, 1983).

This high level of cross-linking possibly compensates for the

low amounts of the usual cell wall strengthening substance

peptidoglycan in the outer membrane. In this way, the spore-

like EBs are osmotically more stable and less permeable than

RBs, which allows them to survive up to several months

outside the host cell (Longbottom & Coulter, 2003).

*These authors contributed equally to this work.Address for correspondence: Daisy Vanrompay, Department ofMolecular Biotechnology, Ghent University, Coupure Links 653,Ghent, B-9000, Belgium. E-mail: [email protected]

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 2: Chlamydial biology and its associated virulence blockers

With their 1:1 RNA:DNA ratio, EBs are thought to be

metabolically inert until attachment to a susceptible host cell

and subsequent internalization.

Following uptake by the host cell, the disulphide bonds

between the outer membrane proteins are reduced, resulting

in a more permeable outer membrane for the reticulate body

to facilitate nutrient uptake (Newhall & Jones, 1983). The

differentiation of EB to RB also involves an expansion,

resulting in a less electron dense cytoplasm in which the

nucleus can no longer be clearly distinguished. The bacteria

become transcriptionally more active and as a consequence,

the cytoplasm contains more RNA and ribosomes, required

for protein synthesis (Ward, 1988). The metabolically active

reticulate bodies are no longer infectious and replicate

intracellularly by binary fission.

During maturation from RBs back to EBs, morphologic-

ally intermediate bodies can be observed in the host cells. In

these IBs, with a diameter ranging from 0.3 to 1.0 mm, a

central electron dense core can be observed with radially

arranged individual nucleoid fibers surrounding the core. As

for EBs, IBs are capable of infecting host cells, at least in

vitro (Litwin et al., 1961; Costerton et al., 1976; Vanrompay

et al., 1996; Rockey & Matsumoto, 2000).

Both EBs and RBs bear rosette-like structures and

projections, located in separate patches, anchored in the

cytoplasmic membrane and extending through the outer

membrane. On the RBs, the patches delineate a zone of

close contact between the bacterium and the plasma mem-

brane derived inclusion membrane. Bavoil and Hsia (1998)

speculated that the projections are in fact functional Type III

secretion system (T3SSs), injecting chlamydial virulence

proteins into the host cell cytoplasm. Beeckman and

Vanrompay (2010a) provided a detailed description of the

T3SS. However, its role in several aspects of the chlamydial

developmental cycle will be briefly mentioned here as well.

Outer membrane composition

Chlamydiaceae are surrounded by two membranes, as is the

case for all Gram-negative bacteria: a cytoplasmic inner

membrane and an outer membrane, separated by a periplas-

mic space. The outer membrane of EBs predominantly

consists of phospholipids, lipids, lipopolysaccharides and

proteins. In contrast to other Gram-negative bacteria,

Chlamydiaceae do not (Barbour et al., 1982) or hardly

possess any muramic acid (Fox et al., 1990). Similar to other

Gram-negative bacteria, an important part of the chlamydial

cell wall is insoluble in sarcosyl, an ionic detergent.

This fraction normally consists of peptidoglycan, covalently

linked to lipoproteins. In Chlamydiaceae, however, only

negligible amounts of peptidoglycan are present (Moulder,

1993; Hatch, 1996), although part of their cell wall is also

insoluble in sarcosyl. Moreover, genes for peptidoglycan

synthesis are present in the genome. Nevertheless,

Chlamydiaceae are sensitive to penicillin and other antibiotics

that target peptidoglycan synthesis, known as the ‘‘chlamydial

anomaly’’ (Moulder, 1993). However, three penicillin binding

proteins (PBPs) have so far been described and each of them

binds to and is inhibited by b-lactam antibiotics (Barbour

et al., 1982; Moulder, 1993; Gump, 1996) and a peptidogly-

can-associated lipoprotein (Pal) was recently found in the

COMC, normally anchoring the outer membrane to peptido-

glycan (Liu et al., 2010).

The cell wall fraction insoluble in sarcosyl is called the

‘‘Chlamydia Outer Membrane Complex’’ (COMC) or cell

envelope, and predominantly consists of MOMP, the cysteine

rich proteins (CRP) Omp2 and Omp3, as well as the

polymorphic membrane proteins (pmps) (Hatch et al., 1984;

Sardinia et al., 1988; Stephens & Lammel, 2001). The outer

membrane is further composed of lipopolysaccharides, PorB,

Omp85, the heat shock proteins hsp60 and hsp70 and OprB.

Some of these components will now be discussed in further

detail.

The MOMP protein has a molecular weight of ~40 kDa,

covering 60% of the outer membrane in RBs and almost 100%

in EBs. It is rich in cysteine residues and is always present as

a trimer. After reduction of disulphide bonds, MOMP can

function as a porin, allowing nutrient uptake by the RB. The

MOMP protein contains four variable domains (VD1-VD4),

which comprise family, genus, species, subspecies (or biovar)

and serovar specific epitopes (Caldwell et al., 1981; Yuan

et al., 1989; Everett, 2000; Kim & DeMars, 2001). In

addition, MOMP has been described to function as an

adhesin, mediating nonspecific (electrostatic and hydropho-

bic) interactions with host cells (Su et al., 1990).

The second most important components of the COMC are

two cysteine rich and highly immunogenic proteins, Outer

membrane protein 2 (Omp2) and 3 (Omp3), and are

abundantly present in EBs but hardly in RBs. Transcription

and translation of the CRP genes is most probably develop-

mentally regulated, the proteins being re-synthesized late in

the growth cycle at the time of differentiation of RBs to EBs

(Newhall, 1987; Sardinia et al., 1988). Outer membrane

protein 2 is the larger of the two CRPs. It is highly

immunogenic, Chlamydiaceae specific and can be used as a

marker for chlamydial infections (Caldwell et al., 1981;

Table 1. Characteristics of chlamydial elementary and reticulate bodies.

Characteristic Elementary body Reticulate body References

Morphology Spherical Spherical Costerton et al., 1976; Longbottom and Coulter, 2003Diameter 0.2–0.3 mm 0.5–1.6 mmElectron density High LowInfectivity for the host High NoneRNA/DNA ratio 1:1 3:1 (more ribosomes)Metabolic activity Relatively inactive Active, binary fissionCell wall Rigid, cross-linked Permeable, fragile Newhall and Jones, 1983Projections (T3SSs) 11–20, small patch Up to 83, larger patch Matsumoto et al., 1976; Matsumoto, 1982a, 1982b

314 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 3: Chlamydial biology and its associated virulence blockers

Sardinia et al., 1988; Sanchez-Campillo et al., 1999). The

Omp2 of C. trachomatis LGV1 is surface exposed, containing

a heparin binding peptide and has been shown to function

as an adhesin (Stephens et al., 2001; Fadel & Eley, 2007;

Fadel & Eley, 2008). The smaller CRP is known as Omp3.

This lipoprotein is probably inserted into the outer membrane

by means of a signal peptide. Its gene sequence (omp3) is less

conserved within the Chlamydiaceae as compared to omp2

(Everett & Hatch, 1991).

The polymorphic membrane proteins or pmps were first

discovered by the group of Longbottom et al. (1996) at the

surface of C. abortus S26/3. Genome sequencing revealed

chlamydial pmp gene families with variable numbers in

different chlamydial species, representing 3–5% of the

genome (Stephens et al., 1998; Read et al., 2003; Thomson

et al., 2005; Harley et al., 2007). Homology searches,

structural comparisons and amino acid sequence analysis

strongly suggest that the pmps in fact belong to the family of

autotransporters. Phylogenetic analysis indicates that they fall

into six subtypes, implying at least six different roles, most

probably also in chlamydial virulence (Henderson & Lam,

2001). Their role in the growth and development of

Chlamydiaceae is still unclear.

Overview of the developmental cycle

Chlamydiaceae have a unique biphasic replication and

survival mechanism. As obligate intracellular bacteria, multi-

plication of metabolically active RBs can only take place in

an eukaryotic host cell, while EBs are adapted to survive in

hostile extracellular environments and can infect new host

cells. An overview of the different stages in the chlamydial

life cycle is presented in Figure 1.

The acute infection starts with the attachment of EBs to a

eukaryotic host cell, followed by internalization within tight,

endocytic vesicles, termed inclusions. Bacteria preferentially

attach near microvilli on the apical cell surface of the host

cell. As the membrane regions at the bases of the microvilli

actively transport extracellular materials into the cells,

attachment there might assist in rapid and efficient entry

(Escalante-Ochoa et al., 1998). In addition, attachment of C.

psittaci EBs is sometimes observed in association with

clathrin-coated pits (Vanrompay et al., 1996). The exact

nature of attachment and entry remains elusive as several

conflicting mechanisms, possibly occurring independently

from each other, have been described in the past, elegantly

summarized as ‘‘parasite specific endocytosis’’ (Byrne &

Moulder, 1978). The newly formed inclusions efficiently

avoid fusion with cellular lysosomes and subsequent acidifi-

cation, and for C. trachomatis, but not for C. pneumoniae or

C. psittaci, fusion of different vacuoles into a larger inclusion

can be observed (Ridderhof & Barnes, 1989; Rockey et al.,

1996; Vanrompay et al., 1996; Hackstadt et al., 1999).

Beginning at 2 h post infection, EBs start differentiating into

RBs, then migrate towards the periphery of the inclusion, to

start replication from 8 h post infection on. Simultaneously,

the surface of the inclusion membrane increases through

acquisition of host plasma proteins and lipids and hijacking of

Golgi-derived vesicles with sphingomyelins (Hackstadt et al.,

1996; Scidmore et al., 1996). In addition, the inclusion

membrane is actively modified through insertion of so-called

‘‘inclusion membrane proteins’’ or Incs (Rockey et al., 2002).

Active replication through binary fission continues until late

in the developmental cycle, when RBs, detached from the

inclusion membrane, revert into EBs again, which are then

stored in the lumen of the inclusion. Depending on the

species, EBs and some non-differentiated RBs are released

from the host cell at 24–72 h post infection through lysis or

reverse endocytosis.

Chlamydiaceae can also engage in a long-term relationship

with the host cell (at least in vitro), a phenomenon known as

persistence, in which no visible growth of the chlamydial

organisms can be observed. The normal developmental cycle

can be interrupted by a number of conditions and agents, such

as antibiotics, nutrient deprivation, or immune factors,

interferon-gamma (IFN-g) in particular (Mpiga &

Ravaoarinoro, 2006). This is generally accompanied by the

development of relatively small inclusions, enlarged pleio-

trophic RBs or persistent bodies (PBs) (Hogan et al., 2004).

Persistent bodies accumulate chromosomes, but genes for cell

division are no longer expressed (Byrne et al., 2001; Mathews

et al., 2001; Gerard et al., 2001). Once the stress-inducing

factor is removed, PBs revert to normal RBs, complete

the developmental cycle and generate infectious EBs.

Whether the different described in vitro persistence models

are relevant to the in vivo described chronic infections

remains to be seen.

Attachment: about glycosaminoglycans, adhesins andhost cell receptors

Attachment of EBs to the host cell surface is thought to

consist of at least two individual steps: an initial, reversible

electrostatic interaction with heparan sulphate-like glycosa-

minoglycans (GAGs) (Zhang & Stephens, 1992; Su et al.,

1996; Davis & Wyrick, 1997), followed by an irreversible,

temperature-dependent binding of a chlamydial ligand to an

unknown host cell receptor, inducing internalization (Carabeo

& Hackstadt, 2001; Fudyk et al., 2002). There has been much

debate on whether the involved GAGs are of chlamydial or

host cell origin. However, genome sequencing indicated that

no chlamydial genes coding for GAG biosynthesis are present,

so the GAGs must be of host cell origin. Given the differential

effects of GAG on the attachment and infectivity of different

Chlamydiaceae species (Zhang & Stephens, 1992; Su et al.,

1996; Rasmussen-Lathrop et al., 2000; Fadel & Eley, 2004), it

seems likely there are both GAG-dependent and GAG-

independent mechanisms operating at different stages of

chlamydial attachment.

The specific host cell receptors and chlamydial ligands

involved in the irreversible attachment of EBs to their host

cells are largely undefined. However, possible bacterial

ligands so far described include MOMP (Su et al., 1990),

Hsp70 (Raulston et al., 1993), OmcB (Ting et al., 1995;

Moelleken & Hegemann, 2008) and pmp21 (pmpD), pmp6

and pmp20 (Wehrl et al., 2004; Crane et al., 2006; Moelleken

& Hegemann, 2008). These pmps thought to be implicated in

bacterial adhesion, although they are not classified amongst

the adhesive trimeric autotransporters (Cotter et al., 2005).

Chlamydial T3SS translocon components (CopB, CopD and

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 315

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 4: Chlamydial biology and its associated virulence blockers

LcrV) as well could possibly serve as adhesions (Watarai

et al., 1996; Skoudy et al., 2000).

Internalization

On lipid rafts and clathrin-coated pits

Based on electron microscopic studies, two major possible

mechanisms for entry are described. The first involves

sequential zipper-like microfilament dependent phagocytosis

induced by binding of chlamydial adhesins to host cell

receptors (Byrne & Moulder, 1978), while receptor-mediated

endocytosis into clathrin-coated pits, independent of

microfilaments, has been described as a second entry

mechanism (Hodinka et al., 1988; Vanrompay et al., 1996).

Zipper-like microfilament dependent entry is clathrin-

independent and most probably occurs through cholesterol-

rich lipid raft microdomains, as has been shown for some

chlamydial species and serovars (Stuart et al., 2003). These

microdomains are also known as lipid rafts and are

characterized by a high cholesterol and glycosphingolipid

content. Moreover, these rafts are intimately connected to the

actin cytoskeleton (Lillemeier et al., 2006) and function as

signaling platforms (Simons & Toomre, 2000) to control

endocytosis (Parton & Richards, 2003), intracellular vesicle

Figure 1. Schematic overview of the developmental cycle of Chlamydiaceae. Bacteria attach preferentially at the base of microvilli and then enter thehost cell through parasite specific endocytosis. Within the thus formed inclusion, avoiding fusion with host cell lysosomes, EBs transform into RBs.RBs proliferate at the boundaries of the inclusion by binary fission, until detachment from the inclusion membrane. RBs revert back to EBs and arestored in the lumen of the inclusion until liberation through lysis or reverse endocytosis.

316 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 5: Chlamydial biology and its associated virulence blockers

trafficking (Helms & Zurzolo, 2004) and activation of

immune response and apoptosis (Gombos et al., 2006).

Their unique composition might therefore be a trigger to

release bacterial effectors in close proximity to these signal-

ing hotspots. Disruption of the rafts through extraction of

plasma membrane cholesterol inhibited internalization

(but not attachment) of C. trachomatis L2, however, as

C. trachomatis serovars A, B, C and L2 and C. muridarum

MoPn internalization occurred independently from lipid rafts,

while C. trachomatis serovar K, C. psittaci, C. pneumoniae

and C. caviae did enter the host cell through raft domains

(Stuart et al., 2003). Different methods or adaptation of the

strains used to different laboratory culture protocols could

have contributed to these conflicting results. The fact that

entry via lipid rafts remodels the actin skeleton and raft-

derived endosomes do not enter the lysosomal degradation

pathway (Helms & Zurzolo, 2004), two key elements of the

chlamydial life cycle, strongly suggests that microfilament

dependent entry occurs through these lipid rafts. The exact

mechanism remains elusive, but could be as proposed in

Figure 2.

A second proposed entry mechanism is receptor-mediated

endocytosis by clathrin-coated pits, based on observed

association of chlamydial organisms with clathrin-coated

pits, and uptake into clathrin-coated vesicles for C. tracho-

matis, C. psittaci and C. caviae strains (Reynolds & Pearce,

1990). Hybiske and Stephens (2007) found clathrin and

its coordinate accessory factors required for entry of

C. trachomatis LGV serovar L2. Other studies demonstrate

that clathrin-coated vesicles are not absolutely required for

chlamydial internalization (Dautry-Varsat et al., 2005; Balana

et al., 2005). In addition, conventional clathrin pits, normally

used for the uptake of physiologically essential and large

molecules, measure only 100 nm in diameter, and thus are too

small to accommodate chlamydial EBs (average size 250 nm).

Nevertheless, clathrin seems to be associated with chlamydial

entry in some way, its importance however strongly linked to

the inoculation route (static or centrifuge assisted) and culture

conditions (Wyrick et al., 1989; Prain & Pearce, 1989;

Reynolds & Pearce, 1990).

Actin recruitment

Within minutes upon attachment of EBs to the host cell,

chlamydiae recruit actin to the site of invasion leading to the

formation of an actin-rich pedestal underneath the attachment

site. This recruitment of actin is transient, eventually leading

to the uptake of EBs into membrane-bound vesicles, as has

already been shown for C. trachomatis (Carabeo et al., 2002),

C. pneumoniae (Coombes & Mahony, 2002), C. caviae

(Subtil et al., 2004) and C. psittaci (Beeckman et al., 2007).

A protein involved in this process has recently been identified

and was termed Tarp for Translocated Actin-Recruiting

Phosphoprotein. Tarp is translocated into the cytoplasm of

the host cell in the early stages of invasion, using a T3SS, and

is spatially and temporarily associated with the recruitment

of actin at the site of internalization (Clifton et al., 2004). The

N-terminal region of C. trachomatis Tarp contains a number

of tyrosine-rich tandem repeats (approximately 50 residues in

length) that are phosphorylated inside the host cell, initiating

a signal transduction cascade by interacting with guanosine

nucleotide exchange factors and small GTPases (Lane et al.,

2008). Orthologs of Tarp are also present in all pathogenic

Chlamydiaceae species examined to date. Moreover, recent

Figure 2. Attachment and entry of chlamydial EBs. Chlamydiaceae interact with the host cell through reversible electrostatic interaction with heparinsulphate-like GAGs on the cell surface, followed by an irreversible interaction with unidentified host cell receptors, possibly associated withcholesterol-rich lipid raft microdomains. Next, host specific signal transduction pathways mediate internalization of the bacteria bv actin recruitmentand pedestal formation, possibly after injection ot T3SS effector proteins (e.g. Tarp). Infection leads to rapid phosphorylation of host cell proteins.Image reproduced from (Dautry-Varsat et al., 2005).

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 317

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 6: Chlamydial biology and its associated virulence blockers

functional studies concluded the C-terminal region of Tarp to

be involved in the actin recruitment and nucleation of actin

filaments (Jewett et al., 2006). A proline-rich domain (S625-

N650 in C. trachomatis L2) induces homo-oligomerization

of Tarp and, in conjunction with the actin-binding domain

(D726-S825), the ability to nucleate actin. Both domains

are conserved among chlamydial strains sequenced so far,

suggesting that Tarp-mediated actin polymerization is not

merely the result of a stable association between Tarp and

actin but is more complex, involving multiple domains of

Tarp (Jewett et al., 2006).

Another protein involved in actin recruitment upon

attachment, at least in C. trachomatis, is the host protein

ezrin, an ezrin-radixin-moesin family protein, colocalizing

with actin at the tips and crypts of microvilli, the site of

chlamydial attachment and entry, respectively. Initial ezrin

activation through threonine phosphorylation is ubiquitous

among chlamydiae. However, subsequent tyrosine phosphor-

ylation of this protein was only observed for an infection of

cells with C. trachomatis strains. This might relate to an

undefined species-specific mechanism of pathogen entry

that involves chlamydial specific ligand(s) and host cell

co-receptor usage (Swanson et al., 2007). The fact that ezrin

is known to interact with the cytoplasmic domain of several

receptors, including CD44 and members of the integrin

superfamily, strengthens the hypothesis that T3SS translocon

components like CopB, CopD and LcrV could function as

possible chlamydial adhesins (see above).

A third mechanism represents host-receptor mediated

initiation of actin recruitment upon chlamydial attachment

to a Platelet-derived growth factor-like receptor (PDGFRb),

as inhibition of PDGFRb by RNA interference or by PDGFRbneutralizing antibodies significantly reduces bacterial binding

(Elwell et al., 2008).

In conclusion, Chlamydiaceae most probably enter their

host cells by more than one pathway, the different steps of

which may overlap to some extent and a strict equilibrium

between these different pathways may well be controlled by

small GTPases.

Inhibition of the phagolysosomal fusion

In both professional and non-professional phagocytes, phago-

somes containing a pathogen normally fuse with lysosomes,

where after the resulting phagolysosomes produce acidic

hydrolases to eradicate the pathogen. However,

Chlamydiaceae have evolved a mechanism to efficiently

thwart phagolysosomal fusion. This particular inhibition is

restricted to the chlamydial inclusion vacuoles, since in mixed

infections with C. psittaci and Saccharomyces cerevisiae or

E. coli, vacuoles not containing Chlamydiae do fuse with

lysosomes, and neither EBs or RBs can protect the co-

infecting organism from degradation (Eissenberg & Wyrick,

1981). In addition, in vitro analyses show that not all

inclusions escape phagolysosomal fusion depending on the

host cell, chlamydial strain and the mode of chlamydial entry

(Moulder, 1991). The chlamydial inclusion is not merely an

endosomal inclusion, as shown by the absence of markers of

the plasma cell membrane, or markers for either the early or

late endosome or for lysosomes (Heinzen et al., 1996;

Scidmore et al., 1996; Taraska et al., 1996; Al-Younes

et al., 1999). In accordance with the absence of vacuolar H+

ATPase, no acidification of the inclusion lumen can be

observed, although the neutral pH of the vacuole can also

result from the activity of the Na+, K+-ATPase ion pumps

(Heinzen et al., 1996; Schramm et al., 1996; Grieshaber et al.,

2002). Many theories explaining the phagolysosomal inhib-

ition have been proposed and these are elegantly summarized

in a review by Escalante-Ochoa et al. (1998). Even when

chlamydial protein synthesis is blocked through the addition

of chloramphenicol, EB containing vesicles only slowly

acquire lysosomal characteristics. Based on these results, a

two-stage mechanism for chlamydial avoidance of lysosomal

fusion has been proposed: (i) an initial phase of delayed

maturation to lysosomes due to an intrinsic property of EBs

and (ii) an active modification of the inclusion membrane

requiring chlamydial protein synthesis (Scidmore et al.,

2003), including the synthesis of chlamydial inclusion

membrane proteins. This so-called intrinsic property of EBs

could well be the translocation of already produced Type III

secretion (T3S) effector proteins to escape fusion with

lysosomes (Hackstadt et al., 1997; Wyrick, 2000).

Proliferation

Bacterial division through binary fission takes place during

the RB stage. Genome sequencing indicated that

Chlamydiae lack an identifiable ftsZ (filamentation tempera-

ture sensitive mutant Z) ortholog, which encodes a key

protein in bacterial cell division and is highly conserved

among all other sequenced eubacteria. However, RBs

synthesize small amounts of peptidoglycan that may play a

role in bacterial cell division, perhaps by substituting for the

lack of FtsZ in the formation of nascent division septa

(Chopra et al., 1998) and reviewed in McCoy and Maurelli

(2006).

Type III secretion

The role of the Type III secretion system in the chlamydial

life cycle has partly been described above but will be, for

reasons of clarity, summarized here. Firstly, T3SS translocon

components may be involved in the irreversible attachment of

EBs, assuming that CopB functions as an adhesin, binds the

hyaluronan receptor CD44 in lipid rafts, thereby inducing

complete assembly of the T3SS translocon in the eukaryotic

membrane. This could then be followed by an interaction of

the CopB-CopD-LcrV complex with a nearby a5b1 integrin

receptor on the host cell (Watarai et al., 1996; Skoudy et al.,

2000). Chlamydiaceae also employ receptor-mediated uptake

by clathrin-coated pits. Possible ligands could include surface

exposed T3SS components, e.g. the outer membrane secretin

SctC or the needle protein SctF (Beeckman & Vanrompay,

2010b).

Translocation of the T3S effector protein Tarp into the host

cell within minutes following attachment induces actin

recruitment and the formation of pedestal-like structures

beneath the attached EB. In C. trachomatis at least two

accompanying signaling cascades (Arp2/3 dependent or not)

have already been described (Jewett et al., 2006; Lane et al.,

2008), but most likely, additional currently uncharacterized

318 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 7: Chlamydial biology and its associated virulence blockers

T3SEs, are implicated in other signaling pathways mediating

EB internalization.

As the infection progresses, inclusion membrane proteins

or Incs are being inserted, presumably through T3S, into the

chlamydial inclusion membrane. At least some of them

are believed to actively prevent fusion between the inclusion

and lysosomes (Hackstadt et al., 1997; Wyrick, 2000), thus

preventing acidification of the inclusion body and ensuring

chlamydial survival. Other Incs such as CT229, Cpn0585 or

IncA are believed to divert intracellular host cell trafficking

to the nascent inclusion to intercept nutrients and constitu-

ent molecules from normal eukaryotic trafficking

pathways (Rzomp et al., 2006; Cortes et al., 2007; Delevoye

et al., 2008).

The Chlamydia protein associating with death domains

(CADD), another T3S effector protein, may help in the

reprogramming of the host cell or modulate host cell

apoptosis via binding to the death domains of tumor

necrosis factor receptor (Stenner-Liewen et al., 2002;

Schwarzenbacher et al., 2004).

While proliferating, RBs are intimately connected to the

inclusion membrane through a well-delineated patch contain-

ing T3SSs. Because of restrictions in host cell size, some RBs

and therefore also the T3SSs lose contact with the inclusion

membrane, inactivating the T3SS. This induces the asyn-

chronous re-differentiation into EBs (Wilson et al., 2006;

Peters et al., 2007).

Nucleotide acquisition

Multiple findings in literature indicate that the

Chlamydiaceae behave as ‘‘energy parasites.’’ However,

genome sequencing revealed that Chlamydiae possess genes

allowing to produce their own ATP, probably by both the

glycolytic pathway and their truncated tricarboxylic acid

cycle (Iliffe-Lee & McClarty, 1999). Still, Chlamydiaceae

posses nucleotide transport proteins (NTTs), enabling them to

perform ATP-ADP counter-exchange and import of nucleo-

tides. Similarly, the genome of C. trachomatis contains two

genes coding for nucleoside phosphate transporters 1 and 2

(Npt1Ct and Npt2Ct), each performing a different type of

transport (Tjaden et al., 1999). Npt1Ct is an ATP-ADP

exchanger, able to function in ATP acquisition from the host

cytosol. On the contrary, Npt2Ct catalyses transport of H+ and

all four ribonucleoside triphosphates into the cell, and thus

provides for the net uptake of ribonucleoside triphosphates

required for anabolic reactions. The ribonucleoside triphos-

phate/H+ transporters of other Chlamydia spp. may be of

different specificity (Tjaden et al., 1999). Both transporters

are present in C. pneumonia (Kalman et al., 1999), and

also C. psittaci exhibits an ATP/ADP exchange (Hatch

et al., 1982).

Emerging antibiotic resistance in Chlamydiaceae

In most cases, chlamydial infections can easily be resolved by

treatment with antibiotics such as tetracycline derivatives and

macrolides (especially azithromycin). However, reports

regarding treatment failure have been described already

(Johnson and Spencer, 1983; Jones et al., 1990; Lefevre &

Lepargneur, 1998; Misyurina et al., 2004; Di Francesco

et al., 2008; Andersen & Rogers, 1998) and frequently

heterotypic resistance (resistance in which only a small

portion of the population displays the resistant phenotype) is

observed. In these cases, it is not always clear whether

persistence or actual antibiotic resistance is involved. Drug

resistance frequently arises through point mutations, altering

the expression or the functionality of the antibiotic target, or

following the insertion of resistance genes into the bacterial

genome. An excellent overview on mutations described in

Chlamydiae to acquire resistance against antibiotics can be

found in Sandoz and Rockey (2010). Until recently, it was

generally believed that the acquisition of antibiotic resistance

in Chlamydia spp. through lateral gene transfer from other

organisms was limited due to their obligate intracellular life

style. However, Dugan et al. (2004, 2007) were the first to

demonstrate stable tetracycline resistance in C. suis probably

occurs through horizontal gene transfer. This was followed by

studies of DeMars and colleagues (2007, 2008) demonstrating

in vitro lateral gene transfer and homologous recombination

between single antibiotic resistant C. trachomatis strains to

obtain doubly resistant isolates. Moreover, evidence exists

that antibiotic resistance genes can also be transferred

within and among chlamydial species, such as C. suis,

C. trachomatis and C. muridarum, and into clinical isolates

from human patients infected with C. trachomatis (Suchland

et al., 2009).

The supplementation of feeds with antibiotics (especially

tetracyclines) was widespread in the poultry, porcine and live

stock industry in order to promote growth and counter

bacterial infections (Sarmah et al., 2006; Castanon, 2007;

Moulin et al., 2008; BelVet-SAC, 2012). Considering the high

prevalence of tetracycline resistance in e.g. porcine C. suis

isolates (Schautteet et al., 2012), it is not unconceivable that

this also is or will be the case in other meat producing

industries, resulting in treatment difficulties and potentially

severe economic losses due to antibiotics resistance. Perhaps

more importantly, there is a potential risk for public health as

contact between tetracycline resistant and tetracycline sensi-

tive Chlamydia spp., in different settings such as farms,

veterinary clinics and slaughterhouses, may lead to transfer of

both the resistance genes and the resulting phenotype, which

could then be propagated and selected for in patients treated

with tetracycline. This would interfere with the treatment of

chlamydial infections, resulting in more severe complications

and a higher mortality rate. In order to combat pathogenic

bacteria which are untreatable using conventional antibiotics,

new means of therapy should be developed, thereby focusing

on traits which are indispensable for pathogenic characteris-

tics, the so-called virulence factors, rather than merely killing

the bacteria.

Blocking chlamydial virulence

Virulence blockers can be defined as compounds that

specifically target virulence determinants of pathogenic

bacteria, thereby preventing the bacteria to colonize the host

and allowing the host immune system to clear the infection.

As most of these blockers do not directly kill the bacteria –

they disarm rather than destroy – it is presumed that the

evolutionary pressure for the development of resistant strains

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 319

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 8: Chlamydial biology and its associated virulence blockers

is smaller than with classic antibiotics. Popular targets include

biofilm formation, bacterial toxins, specialized secretion

systems, organism-specific virulence gene expression or

cell-to-cell signaling, as Rasko and Sperandio (2010) ele-

gantly reviewed. For the purpose of this review, we will focus

on possible mechanisms and compounds that may efficiently

block different stages in the chlamydial life cycle.

Inhibition of adhesion

The very first interaction between bacteria and their host cell

is the process of adhesion to the cell membrane. Therefore, in

order to effectively prevent bacterial colonization of the host,

one could already prevent attachment of the pathogen to the

host cell membrane. For most bacteria adhesins such as

fimbriae (Type1 and 4 pili) or adhesive autotransporters

(see below) have been described. Assembly of these pili by

the chaperone/usher pathway can be effectively blocked by

treatment with so-called pilicides (Aberg & Almqvist, 2007).

However, the adhesion mechanism in chlamydial species

remains rather elusive, and pili do not seem to be involved.

Research should therefore focus on already characterized

chlamydial adhesins such as MOMP and the pmp-proteins.

Such adhesins could effectively be blocked by specific

antibodies, thereby neutralizing chlamydial infectivity and

reducing colonization by blocking chlamydial attachment to

epithelial cells. In this respect, it has been shown that

monoclonal antibodies against MOMP could neutralize

chlamydial infection in vitro (Peeling et al., 1984, Peterson

et al., 1991) and could provide a modest level of protection

against infection when administered passively to mice (Cotter

et al., 1995). Similarly, antibodies specific to PmpD of

C. trachomatis and C. pneumoniae and Pmp2 and 10 of

C. pneumoniae were shown to be neutralizing, at least in vitro

(Wehrl et al., 2004; Finco et al., 2005; Crane et al., 2006).

As described above, heparin sulphate-like glycosaminogly-

cans are also involved in the chlamydial attachment process.

Monoclonal antibodies specifically directed against heparan

sulphate specifically bind glycosaminoglycans localized to

the surface of C. trachomatis and C. pneumoniae and

effectively neutralize their infectivity (Rasmussen-Lathrop

et al., 2000).

However, evidence exists that chlamydial bacteria most

likely use different mechanisms of attachment to the host cell

(see above), rendering the development of a general anti-

adhesion therapy that would completely block chlamydial

attachment unlikely.

Inhibition of internalization

As described above, chlamydiae induce actin recruitment to

the site of infection, followed by a localized and temporary

nucleation, to facilitate uptake into host cell membrane-bound

vesicles. The most straightforward way to block internaliza-

tion would therefore be to interfere with this polymerization

by treatment with molecules such as cytochalasin, latrunculin,

phalloidin, taxol or colchicines (Peterson & Mitchison, 2002).

However, as actin is also implicated in other cellular functions

such as cell shape or cell migration, the side-effects of a

similar treatment would be considerable. One would therefore

have to focus on the process of endocytosis itself to prevent

chlamydiae from invading the host cells. Research could be

directed towards toxins used by pathogenic bacteria such as

Yersinia spp. to prevent phagocytosis. Especially proteins

such as Yersinia YopH and YopE and Pseudomonas

aeruginosa (P. aeruginosa) ExoS and ExoT, interacting with

small GTPases, which are also implicated in chlamydial

internalization, could be of interest. These proteins, which are

Type III secretion substrates, convert Rho family members in

an accelerated manner to their GDP-bound, inactive states and

thus inhibit endocytotic processes (Ernst, 2000).

Inhibition of the Type III secretion system

As described above, T3S is involved in different stages of the

chlamydial life cycle and mediates translocation of virulence

related effector proteins to the host cell cytoplasm (Beeckman

& Vanrompay, 2010b). Consequently, chlamydial disease

might be effectively treated by either blocking T3S or

inhibiting the interaction with the eukaryotic host. In recent

years, several studies describing small molecules specifically

inhibiting T3S have been published (Kauppi et al., 2003;

Keyser et al., 2008). These inhibitors have been identified

through mass screening of chemical libraries using whole-cell

reporter gene assays or ELISA to assess inhibition of T3S and

included salicylideneacylhydrazides, saliylanilides, sulfonyla-

minobenzanilides, salicylideneanilides, phenoxyacetamides,

thiazolidones and N-hydroxybenzimidazoles (Keyser et al.,

2008; Aiello et al., 2010; Escaich, 2010). In the Chlamydia

research community, research has predominantly focused on

the effects of acylated hydrazones of salicylaldehyddes

whereby host cell cytokine expression as well as chlamydial

growth and T3S gene expression, but not entry, were shown to

be affected at non- or low-cytotoxic concentrations (Muschiol

et al., 2006; Wolf et al., 2006; Bailey et al., 2007; Slepenkin

et al., 2007; Prantner & Nagarajan, 2009; Muschiol et al.,

2009; Chiliveru et al., 2010). Wang et al. (2011) identified

putative target proteins of the salicylideneacylhydrazides.

Those proteins are involved in regulation of T3SS gene

expression. In addition, the phenoxyacetamide MBX 1641 is

capable of inhibiting T3S translocation in C. trachomatis

infected Hep-2 cells (Aiello et al., 2010). Although the exact

mode of action has not yet been uncovered, it is very likely

that the conserved (structural) elements of the T3SS or its’

assembly are targeted, especially given the broad spectrum of

bacteria inhibited.

Another strategy is to screen for natural products inhibiting

bacterial T3S. Such components have been described,

including glycolipids (Linington et al., 2002, 2006) and

transferrins (Ochoa et al., 2003; Gomez et al., 2003; Ochoa &

Clearly, 2004; Yekta et al., 2010), of which lactoferrin and

ovotransferrin have proven their potential to inhibit chlamyd-

ial infections in vitro as well (Beeckman et al., 2007).

Moreover, ovotransferrin was shown to efficiently prevent

C. psittaci infection in experimentally infected SPF turkeys

(Van Droogenbroeck et al., 2008) and on a commercial turkey

farm (Van Droogenbroeck et al., 2011). Alternatively, the

chlamydial T3S can also be inhibited in a pure mechanical

manner. Several studies have been published describing

in vitro and in vivo blockage of T3S using antibodies directed

against the translocon adaptor protein LcrV and its analogues

320 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 9: Chlamydial biology and its associated virulence blockers

in other bacteria (Frank et al., 2002; Goure et al., 2005;

Philipovskiy et al., 2005; Gebus et al., 2008; Eisele &

Anderson, 2009; Markham et al., 2010; Van Blarcom et al.,

2010). Whether the LcrV protein is essential in the chlamyd-

ial internalization process as well could be studied while

infecting epithelial cells and/or macrophages in the presence

of anti-LcrV antibodies. If indeed anti-LcrV antibodies could

significantly inhibit C. psittaci internalization and subsequent

replication in vitro, one could test whether active immuniza-

tion with LcrV or passive immunization with anti-LcrV

antibodies could provide protection against C. psittaci

infections in vivo as well, (Mueller et al., 2008). Recently,

a chlamydial T3S effector protein (Tarp) was identified as a

novel immunodominant antigen in human antisera and

immunization with Tarp can induce protective immunity

against chlamydial infection and pathology in mice (Wang

et al., 2009). Information on other T3S effectors in

Chlamydiaceae is scarce, put potential targets for antibody-

mediated inhibition could include the Chlamydia protein

associated with Death Domains CADD, the serine-threonine

kinase Pkn5 or the macrophage infectivity potentiator MIP

(Beeckman & Vanrompay, 2010b).

Blocking bacterial proliferation

Although bacterial proliferation is no virulence determinant

sensu strictu, processes currently not targeted by classical

antibiotics could open possibilities for the generation of novel

antibacterials. Likewise, interest increases in the FtsZ protein

as therapeutic target in the antimicrobial research field. This

protein is essential for bacterial cell division and thus

targeting FtsZ would lead to disruption of cell division and

therefore bacterial infection (Awasthi et al., 2011). However,

as mentioned earlier, Chlamydiaceae do not possess a ftsZ-

ortholog. Nevertheless, some interesting alternative mechan-

ism to inhibit bacterial proliferation exist, such as the

limitation of Fe3+ availability, which is crucial in the bacterial

metabolism and biofilm formation (Raulston, 1997;

Cianciotto, 2007).

The Chlamydiaceae proliferate predominantly in epithelial

cells and macrophages (Vanrompay et al., 1995). The latter

play an important role in the clearance of aged and apoptotic

cells and are therefore continuously exposed to high intracel-

lular iron loads. Though the mode of Fe3+ scavenging from

the environment by Chlamydia and other intracellular bacteria

such as Legionella or Mycobacterium, is undefined, the

cytosolic iron pool is most likely the source.

Accordingly, depletion of cytosolic iron could limit the

growth of intracellular bacteria. This concept is demonstrated

by the incubation of C. psittaci – or L. pneumophila – infected

mouse macrophages with iron chelators deferriprone or

desferasirox which results in a reduced level of bacterial

infections (Paradkar et al., 2008). Both compounds, deferri-

prone and desferasirox, have previously been approved for

human use. They are membrane permeable as they can

remove iron from iron loaded macrophages (Paradkar et al.,

2008). This new generation of chelators has great therapeutic

potential for treatment of persistent bacterial infections.

An alternative strategy to limit intracellular Fe3+-levels is

the use of the ‘‘Trojan horse’’ transition metal gallium (Ga3+),

an ion chemically similar to iron. Unlike Fe3+, it does not

undergo redox reactions and thus cannot execute the cellular

functions of Fe3+ within the bacterial cell (Chitambar &

Narasimhan, 1991). Through competition with Fe3+, gallium

decreases thus bacterial iron uptake. Consequently, the iron

need of the bacteria is not fulfilled and bacterial growth is

inhibited. Furthermore, gallium proved to be effective both

in vitro and in vivo in treatment of Psuedomonas aeroginosa

infections in rabbit and mouse models (Kaneko et al., 2007;

Banin et al., 2008) and is already approved by the FDA for

use in large doses to treat hypercalcemia of malignancy

(Warrell & Bockman, 1989). All together, this hints gallium

as a promising treatment strategy in bacterial infections.

Nucleotide transporters

As mentioned above, Chlamydiaceae scavenge energy mol-

ecules from the host using nucleotide transport proteins.

These proteins not exclusively constitute bacterial mem-

branes, but are similarly essential to plant chloroplasts where

they participate in the import process of cytosolic ATP under

certain conditions (Winkler & Neuhaus, 1999; Linka et al.,

2003). Interestingly, the bacterial and plant transporters do not

exhibit structural similarity with mitochondrial and peroxi-

somal adenylate transporters, belonging to the mitochondrial

carrier (MC) family (Klingenberg, 1989; Saier, 2000;

Ren et al., 2004). Hence, NTTs are absent in mammalian

and human cells and thus represent an attractive target for the

development of highly specific anti-chlamydial drugs.

However, how the highly charged ATP molecules pass the

inclusion membrane to reach the bacteria is unknown so far,

as pores for passive diffusion are absent in the inclusion

membrane (Heinzen & Hackstadt, 1997).

As earlier stated, the genome of C. trachomatis and

C. pneumoniae contains genes that might encode enzymes of

the glycolytic pathway, pentose pathway and tricarboxylic

acid cycle (Kalman et al., 1999; Stephens et al., 1998).

Accordingly, this suggests chlamydial bacteria are able to

generate their own ATP through oxidative phosphorylation

and would not be strict auxotrophic. In the case of

C. trachomatis, however, strong upregulation of the ATP/

ADP anti-porter gene occurs early in the developmental life

cycle. Namely, uptake of ATP from the host cell would be

most relevant early in the infection process when the whole

complement of enzymes for ATP generation via glycolytic

and pentose pathway is not present yet, e.g. during initial

differentiation of elementary bodies to reticulate bodies

shortly after infection of the host cell. Moreover, the substrate

for oxidative phosphorylation is now limited in the host

cytosol. Under these circumstances, an alternative way for

energy generation is an advantage. Later on, only after

inclusion niche establishment, structural proteins and proteins

of intermediary metabolism are expressed. Once cell division

starts, the energy need rises and available energy generation

increases by the glycolytic and pentose pathways (Shaw et al.,

2000). As the transport proteins are not present in human

cells, and energy parasitism in the initial phase of the

infection process is crucial for the survival of Chlamydia,

blocking this transport could be a specific and efficient

strategy in controlling chlamydial infections. To our

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 321

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 10: Chlamydial biology and its associated virulence blockers

knowledge, there are until now no inhibitors of chlamydial

nucleotide transport identified.

Regulation of virulence gene expression

Quorum sensing

Though the concept of quorum sensing is commonly used by

bacteria, the exact molecular mechanism may differ among

species. At least six different QS pathways are identified so

far (Table 2) (Surette et al., 1999; Schauder et al., 2001; Chen

et al., 2002; Sperandio et al., 2003; Henke & Bassler, 2004;

Kendall et al., 2007; Higgins et al., 2007). A common

signaling pathway containing the membrane-bound QseC

histidine sensor kinase (Clarke et al., 2006) or its homo-

logues, is present in more than 25 important pathogens of

humans and plants. QseC perceives the bacterial quorum

sensing signal autoinducer 3 (AI-3) and/or host derived

adrenalin and/or noradrenalin. After binding of the signal,

QseC increases its autophosphorylation. The following phos-

phorylation cascade in the bacterial cell regulates the

expression of virulence genes (Sperandio et al., 2003;

Hughes & Sperandio, 2008).

The genome of C. trachomatis comprises two genes, ctcB

and ctcC, with protein sequence similarity to histidine kinase-

response regulator pairs of two-component systems. The latter

are a type of QS pathway, which play a role in stage-specific

gene expression, this could be e.g. in- and outside the host

cell, two completely different environments in the

Chlamydiaceae biphasic life cycle. Generally, the histidine

sensor kinase component in the bacterial membrane autopho-

sphorylates upon signal perception and subsequently phos-

phorylates and activates a response regulator, usually a

transcription factor, which binds to the promoter of a target

gene and initiates transcription upon activation. The sensor

kinase and response regulator pair form a genetic network

together with a range of downstream molecular factors. This

network controls a specific subset of genes, including

virulence genes (Novick, 2003; Lyon & Novick, 2004).

Two-component systems are a primary mechanism to

adapt to environmental conditions. Though little is known

about how transcriptional regulation in Chlamydiae, gene

expression and development are most likely controlled

through recognition of environmental cues or intracellular

conditions. Although many bacteria possess several systems

to adapt to diverse environmental changes, this is the only

complete two-component system identified in C. trachomatis.

Moreover, this ctcB-ctcC system proved to be functional as

it is capable of autophosphorylation and phosphotransfer

reactions (Koo & Stephens, 2003).

The CtcB and CtcC genes posses a late expression profile

and, accordingly, the corresponding proteins are present in

EBs, but not in RBs (Koo & Stephens, 2003). Most two-

component system components, however, are constitutively

expressed to adapt efficiently to a changing environment. The

late expression profile thus implies involvement in the control

of a subset of late genes participating in RB to EB transition.

Moreover, the sensor kinase CtcB possesses a redox sensing

domain (Koo & Stephens, 2003). This could sense the change

in redox state when EBs enter the host cells and disulfide-

linkage in the outer membrane proteins are reduced, which

results in a higher membrane flexibility and increased nutrient

uptake. Similarly, a decrease in energy sources or reducing

agents results in oxidation of sulfhydrylgroups, hindering RB

development and decelerating metabolic activity (Bavoil

et al., 1984; Hackstadt et al., 1985; Ward, 1988). To conclude,

CtcB and CtcC are developmentally late-expressed proteins

with redox sensing domain. This domain is most likely

involved in late gene activation, including the regulation of

RBs to EBs differentiation.

Quorum sensing inhibitors

Blocking QS is more and more considered as a viable

approach for developing therapeutics in the treatment of

bacterial infections.

The ideal QS inhibitor (QSI) is a chemically stable, low-

molecular mass molecule without toxic side-effect on the

bacterium or host, and chemically stable and resistant to

metabolization and disposal by the host. It should be specific

for the particular regulon and have a significant and similar

reduction in expression on all the QS regulon comprised

genes, however this is not always the case. The strength of an

inhibitor depends on the percentage of QS-controlled genes it

targets (Arevalo-Ferro et al., 2003; Hentzer et al., 2003;

Rasmussen et al., 2005b). QSIs fall roughly into three

categories according to the level of interruption of the

signalization: repressors of signal generation, disruptors of the

signals or signal molecules and inhibitors of the signal

perception. Alternatively, inhibitors are categorized into four

different classes: nonpeptide small molecules, peptides,

enzymes and antibodies (Pan & Ren, 2009).

To our knowledge, no chlamydial QSI compounds are

known yet. As there is evidence that Chlamydiaceae can sense

the redox-state of their environment, blockage or destruction

of receptor proteins could be an interesting strategy for

therapeutic purposes in this context. One method for receptor

blockage is the use of an analogue of the signal molecule.

More knowledge about the signal perception is needed to

Table 2. Quorum sensing pathways in bacteria.

Pathway Signal molecules Bacteria References

AHL (AI-1 pathway) AHLs Gram-negative Salmond et al., 1995; Ravn et al., 2001; Zavil’gel’skii andManukhov, 2001

4Qs pathway PQS and HHL Gram-negative Diggle et al., 2006AI-3 pathway AI-3 Gram-negative Sperandio et al., 2003; Kendall et al., 2007AI-2 pathway Two different forms Gram-negative and Gram-positives Surette et al., 1999; Schauder et al., 2001; Chen et al., 2002AIP pathway Oligopeptides Gram-positive McDowell et al., 2001CAI-1 Hydroyketones Gram-negative Henke and Bassler, 2004; Higgins et al., 2007

322 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 11: Chlamydial biology and its associated virulence blockers

explore this possibility. Generally, a synthetic library of signal

molecule derivatives is used to screen for inhibitors. However,

random compound libraries with natural and synthetic

compounds can evenly be used (Smith et al., 2003a, 2003b;

Suga & Smith; 2003). In both cases, a screening system and

further validation is necessary to be able to identify potential

inhibitors.

A valuable source for QSI compounds are other bacteria,

fungi and plants. These organisms have co-existed for

millions of years and some of them probably produce QSI

compounds, such as Penicillium species for example

(Rasmussen et al., 2005b). Examples of plants producing

QS inhibitors are garlic, carrot, soybean, tomato and many

more (Rasmussen et al., 2005a).

Beside the species specific inhibitors discussed above,

also broad spectrum inhibitors are already described in

literature. Most QS signals only appear in a small number

of species. However, certain signaling pathways are

common for a range of species, while they are not found

in the eukaryotic hosts. A high throughput screen of a

library of 150.000 small organic compounds identified the

lead structure LED209 (N-phenyl-4-[[(phenylamino) thiox-

omethyl]amino]-benzenesulphonamide). This non-toxic

compound has no effect on pathogen growth but blocks

binding of signaling molecules to QseC, thus preventing the

autophosphorylation of QseC and consequent activation of

virulence genes. LED209 was tested for inhibitory effect,

and showed both in vitro and in vivo a virulence decrease in

models of infection for several pathogens. Furthermore,

molecular concentrations showed a 10-fold reduction

compared to previously characterized virulence inhibitory

compounds (Rasko et al., 2008). LED209 can be considered

as the proof of concept that blocking inter-kingdom

chemical signaling is a viable strategy to develop novel

drugs to control bacterial infection. Unlike the LED209

compound mentioned above, most inhibitors show efficacy

in vitro, but have not been tested in vivo in animal models

yet. This partially explains why no QSI is at clinical stage

of drug development. Moreover, because most of these QSIs

are not in the clinical phase, no information on their

efficacy or toxicity in humans is available. Therefore, more

research in the field of quorum sensing and in vivo testing

is required in order to explore the potential, advantages and

limitations of QSIs as therapeutics in the control of

bacterial infections.

Conclusion

Antibiotic resistance has been reported in Chlamydia.

Virulence blockers could fulfill a role in future prevention

and/or treatment of Chlamydia infections, as they do not

directly inhibit the growth of pathogens, but rather target

virulence associated processes. Therefore, they are considered

to exert a lower selective pressure to develop resistance

compared to classic antibiotics. So far, only ovotransferrin,

the avian homologue of mammalian lactoferrin, has been

tested in an animal (turkeys) model and in veterinary clinical

trials. Ovotransferrin efficiently prevented C. psittaci respira-

tory disease in commercially raised broiler turkeys demon-

strating its potential for veterinary use.

To our knowledge, anti-virulence strategies for human

chlamydial infections have not been implemented in animal

models or human clinical trials. Nevertheless, promising

virulence blockers such as deferriprone and desferasirox are

already approved for human use.

Further in vivo testing of innovative candidate virulence

blockers as well as in vivo testing in animal models and

clinical trials is needed. Not only to assess the potential of

Chlamydia virulence blockers but also to study possible

limitations and safety.

Declaration of interest

The Research Foundation Flanders (FWO-Vlaanderen) is

acknowledged for providing a grant to Delphine S.A.

Beeckman. This study was funded by the Federal Public

Service of Health, Food Chain Safety and Environment

(convention RF-10/6234).

References

Aberg V, Almqvist F. (2007). Pilicides-small molecules targetingbacterial virulence. Org Biomol Chem, 5, 1827–1834.

Aiello D, Williams JD, Majgier-Baranowska H, Patel I, Peet NP, HuangJ, Lory S, Bowlin TL, Moir DT. (2010). Discovery and characteriza-tion of inhibitors of Pseudomonas aeruginosa type III secretion.Antimicrob Agents Chemother, 54, 1988–1999.

Al-Younes HM, Rudel T, Meyer TF. (1999). Characterization andintracellular trafficking pattern of vacuoles containing Chlamydiapneumoniae in human epithelial cells. Cell Microbiol, 1, 237–247.

Andersen AA, Rogers DG. (1998). Resistance to tetracycline andsulfadiazine in swine C. trachomatis isolates. In: Stephens RS, ed.Ninth International Symposium on Human Chlamydial Infection,San Fransico; 313–316.

Arevalo-Ferro C, Hentzer M, Reil G, Gorg A, Kjelleberg S, Givskov M,Riedel K, Eberl L. (2003). Identification of quorum-sensing regulatedproteins in the opportunistic pathogen Pseudomonas aeruginosa byproteomics. Environ Microbiol, 5, 1350–1369.

Awasthi D, Kumar K, Ojima I. (2011). Therapeutic potential of FtsZinhibition: a patent perspective. Expert Opin Ther Pat, 21, 657–679.

Bailey L, Gylfe A, Sundin C, Muschiol S, Elofsson M, Nordstrom P,Henriques-Normark B, Lugert R, Waldenstrom A, Wolf-Watz H,Bergstrom S. (2007). Small molecule inhibitors of type III secretion inYersinia block the Chlamydia pneumoniae infection cycle. FEBS Lett,581, 587–595.

Balana ME, Niedergang F, Subtil A, Alcover A, Chavrier P, Dautry-Varsat A. (2005). ARF6 GTPase controls bacterial invasion by actinremodelling. J Cell Sci, 118, 2201–2210.

Banin E, Lozinski A, Brady KM, Berenshtein E, Butterfield PW, MosheM, Chevion M, Greenberg EP, Banin E. (2008). The potential ofdesferrioxamine-gallium as an anti-Pseudomonas therapeutic agent.Proc Natl Acad Sci USA, 105, 16761–16766.

Barbour AG, Amano K, Hackstadt T, Perry L, Caldwell HD. (1982).Chlamydia trachomatis has penicillin-binding proteins but notdetectable muramic acid. J Bacteriol, 151, 420–428.

Bavoil P, Ohlin A, Schachter J. (1984). Role of disulfide bonding in outermembrane structure and permeability in Chlamydia trachomatis.Infect Immun, 44, 479–485.

Bavoil PM, Hsia RC. (1998). Type III secretion in Chlamydia: a case ofdeja vu? Mol Microbiol, 28, 860–862.

Beeckman DS, Van Droogenbroeck CM, De Cock BJ, Van Oostveldt P,Vanrompay DC. (2007). Effect of ovotransferrin and lactoferrins onChlamydophila psittaci adhesion and invasion in HD11 chickenmacrophages. Vet Res, 38, 729–739.

Beeckman DS, Meesen G, Van Oostveldt P, Vanrompay D. (2009).Digital titration: automated image acquisition and analysis of load andgrowth of Chlamydophila psittaci. Microsc Res Tech, 72, 398–402.

Beeckman DS, Vanrompay DC. (2010a). Biology and intracellularpathogenesis of high or low virulent Chlamydophila psittaci strains inchicken macrophages. Vet Microbiol, 141, 342–353.

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 323

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 12: Chlamydial biology and its associated virulence blockers

Beeckman DSA, Vanrompay DCG. (2010b). Bacterial Secretion Systemswith an Emphasis on the Chlamydial Type III Secretion System.Curr Issues Mol Biol, 12, 17–41.

Belland RJ, Ouellette SP, Gieffers J, Byrne GI. (2004). Chlamydiapneumoniae and atherosclerosis. Cell Microbiol, 6, 117–127.

BelVet-SAC. (2012). BelVet-SAC report 2012. Belgian VeterinarySurveillance of Antimicrobial Consumption. Available at:www.belvetsac.ugent.be.

Byrne GI, Moulder JW. (1978). Parasite-specified phagocytosis ofChlamydia psittaci and Chlamydia trachomatis by L and HeLa cells.Infect Immun, 19, 598–606.

Byrne GI, Ouellette SP, Wang Z, Rao JP, Lu L, Beatty WL, Hudson AP.(2001). Chlamydia pneumoniae expresses genes required for DNAreplication but not cytokinesis during persistent infection of HEp-2cells. Infect Immun, 69, 5423–5429.

Caldwell HD, Kromhout J, Schachter J. (1981). Purification and partialcharacterization of the major outer membrane protein of Chlamydiatrachomatis. Infect Immun, 31, 1161–1176.

Campbell LA, Kuo CC. (2004). Chlamydia pneumoniae–an infectiousrisk factor for atherosclerosis? Nat Rev Microbiol, 2, 23–32.

Carabeo RA, Hackstadt T. (2001). Isolation and characterization of amutant Chinese hamster ovary cell line that is resistant to Chlamydiatrachomatis infection at a novel step in the attachment process. InfectImmun, 69, 5899–5904.

Carabeo RA, Grieshaber SS, Fischer E, Hackstadt T. (2002). Chlamydiatrachomatis induces remodeling of the actin cytoskeleton duringattachment and entry into HeLa cells. Infect Immun, 70, 3793–3803.

Castanon JI. (2007). History of the use of antibiotic as growth promotersin European poultry feeds. Poult Sci, 86, 2466–2471.

Chen X, Schauder S, Potier N, Van Dorsselaer A, Pelczer I, Bassler BL,Hughson FM. (2002). Structural identification of a bacterial quorum-sensing signal containing boron. Nature, 415, 545–549.

Chiliveru S, Birkelund S, Paludan SR. (2010). Induction of interferon-stimulated genes by Chlamydia pneumoniae in fibroblasts is mediatedby intracellular nucleotide-sensing receptors. PLoS ONE, 5, e10005.

Chitambar CR, Narasimhan J. (1991). Targeting iron-dependent DNAsynthesis with gallium and transferrin-gallium. Pathobiology, 59,3–10.

Chopra I, Storey C, Falla TJ, Pearce JH. (1998). Antibiotics, peptido-glycan synthesis and genomics: the chlamydial anomaly revisited.Microbiology (Reading, Engl), 144 (Pt 10), 2673–2678.

Cianciotto NP. (2007). Iron acquisition by Legionella pneumophila.Biometals, 20, 323–331.

Clarke MB, Hughes DT, Zhu C, Boedeker EC, Sperandio V. (2006).The QseC sensor kinase: a bacterial adrenergic receptor. Proc NatlAcad Sci USA, 103, 10420–10425.

Clifton DR, Fields KA, Grieshaber SS, Dooley CA, Fischer ER, MeadDJ, Carabeo RA, Hackstadt T. (2004). A chlamydial type IIItranslocated protein is tyrosine-phosphorylated at the site of entryand associated with recruitment of actin. Proc Natl Acad Sci USA,101, 10166–10171.

Coombes BK, Mahony JB. (2002). Identification of MEK- andphosphoinositide 3-kinase-dependent signalling as essential eventsduring Chlamydia pneumoniae invasion of HEp2 cells. CellMicrobiol, 4, 447–460.

Cortes C, Rzomp KA, Tvinnereim A, Scidmore MA, Wizel B. (2007).Chlamydia pneumoniae inclusion membrane protein Cpn0585 inter-acts with multiple Rab GTPases. Infect Immun, 75, 5586–5596.

Costerton JW, Poffenroth L, Wilt JC, Kordova N. (1976). Ultrastructuralstudies of the nucleoids of the pleomorphic forms of Chlamydiapsittaci 6BC: a comparison with bacteria. Can J Microbiol, 22, 16–28.

Cotter TW, Meng Q, Shen ZL, Zhang YX, Su H, Caldwell HD. (1995).Protective efficacy of major outer membrane protein-specificimmunoglobulin A (IgA) and IgG monoclonal antibodies in amurine model of Chlamydia trachomatis genital tract infection.Infect Immun, 63, 4704–4714.

Cotter SE, Surana NK, St Geme 3rd JW. (2005). Trimeric autotran-sporters: a distinct subfamily of autotransporter proteins. TrendsMicrobiol, 13, 199–205.

Crane DD, Carlson JH, Fischer ER, Bavoil P, Hsia RC, Tan C, Kuo CC,Caldwell HD. (2006). Chlamydia trachomatis polymorphic membraneprotein D is a species-common pan-neutralizing antigen. Proc NatlAcad Sci USA, 103, 1894–1899.

Dautry-Varsat A, Subtil A, Hackstadt T. (2005). Recent insights into themechanisms of Chlamydia entry. Cell Microbiol, 7, 1714–1722.

Davis CH, Wyrick PB. (1997). Differences in the association ofChlamydia trachomatis serovar E and serovar L2 with epithelialcells in vitro may reflect biological differences in vivo. Infect Immun,65, 2914–2924.

Delevoye C, Nilges M, Dehoux P, Paumet F, Perrinet S, Dautry-Varsat A,Subtil A. (2008). SNARE protein mimicry by an intracellularbacterium. PLoS Pathog, 4, e1000022.

DeMars R, Weinfurter J. (2008). Interstrain gene transfer in Chlamydiatrachomatis in vitro: mechanism and significance. J Bacteriol, 190,1605–1614.

Demars R, Weinfurter J, Guex E, Lin J, Potucek Y. (2007). Lateral genetransfer in vitro in the intracellular pathogen Chlamydia trachomatis.J Bacteriol, 189, 991–1003.

Di Francesco A, Donati M, Rossi M, Pignanelli S, Shurdhi A, Baldelli R,Cevenini R. (2008). Tetracycline-resistant Chlamydia suis isolates inItaly. Vet Rec, 163, 251–252.

Diggle SP, Cornelis P, Williams P, Camara M. (2006). 4-quinolonesignalling in Pseudomonas aeruginosa: old molecules, new perspec-tives. Int J Med Microbiol, 296, 83–91.

Dugan J, Rockey DD, Jones L, Andersen AA. (2004). Tetracyclineresistance in Chlamydia suis mediated by genomic islands insertedinto the chlamydial inv-like gene. Antimicrob Agents Chemother, 48,3989–3995.

Dugan J, Andersen AA, Rockey DD. (2007). Functional characterizationof IScs605, an insertion element carried by tetracycline-resistantChlamydia suis. Microbiology (Reading, Engl), 153, 71–79.

Eisele NA, Anderson DM. (2009). Dual-function antibodies to Yersiniapestis LcrV required for pulmonary clearance of plague. Clin VaccineImmunol, 16, 1720–1727.

Eissenberg LG, Wyrick PB. (1981). Inhibition of phagolysosome fusionis localized to Chlamydia psittaci-laden vacuoles. Infect Immun, 32,889–896.

Elwell CA, Ceesay A, Kim JH, Kalman D, Engel JN. (2008). RNAinterference screen identifies Abl kinase and PDGFR signaling inChlamydia trachomatis entry. PLoS Pathog, 4, e1000021.

Ernst JD. (2000). Bacterial inhibition of phagocytosis. Cell Microbiol, 2,379–386.

Escaich S. (2010). Novel agents to inhibit microbial virulence andpathogenicity. Expert Opin Ther Pat, 20, 1401–1418.

Escalante-Ochoa C, Ducatelle R, Haesebrouck F. (1998). The intracel-lular life of Chlamydia psittaci: how do the bacteria interact with thehost cell? FEMS Microbiol Rev, 22, 65–78.

Everett KD, Hatch TP. (1991). Sequence analysis and lipid modificationof the cysteine-rich envelope proteins of Chlamydia psittaci 6BC.J Bacteriol, 173, 3821–3830.

Everett KD. (2000). Chlamydia and Chlamydiales: more than meets theeye. Vet Microbiol, 75, 109–126.

Fadel S, Eley A. (2004). Chlorate: a reversible inhibitor of proteoglycansulphation in Chlamydia trachomatis-infected cells. J Med Microbiol,53, 93–95.

Fadel S, Eley A. (2007). Chlamydia trachomatis OmcB protein is asurface-exposed glycosaminoglycan-dependent adhesin. J MedMicrobiol, 56, 15–22.

Fadel S, Eley A. (2008). Differential glycosaminoglycan binding ofChlamydia trachomatis OmcB protein from serovars E and LGV.J Med Microbiol, 57, 1058–1061.

Finco O, Bonci A, Agnusdei M, Scarselli M, Petracca R, Norais N,Ferrari G, Garaguso I, Donati M, Sambri V, Cevenini R, Ratti G,Grandi G. (2005). Identification of new potential vaccine candidatesagainst Chlamydia pneumoniae by multiple screenings. Vaccine, 23,1178–1188.

Fox A, Rogers JC, Gilbart J, Morgan S, Davis CH, Knight S, Wyrick PB.(1990). Muramic acid is not detectable in Chlamydia psittaci orChlamydia trachomatis by gas chromatography-mass spectrometry.Infect Immun, 58, 835–837.

Frank DW, Vallis A, Wiener-Kronish JP, Roy-Burman A, Spack EG,Mullaney BP, Megdoud M, Marks JD, Fritz R, Sawa T. (2002).Generation and characterization of a protective monoclonalantibody to Pseudomonas aeruginosa PcrV. J Infect Dis, 186,64–73.

Fudyk T, Olinger L, Stephens RS. (2002). Selection of mutant cell linesresistant to infection by Chlamydia spp [corrected]. Infect Immun, 70,6444–6447.

Gebus C, Caroline G, Faudry E, Eric F, Bohn YS, Elsen S, Sylvie E,Attree I. (2008). Oligomerization of PcrV and LcrV, protective

324 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 13: Chlamydial biology and its associated virulence blockers

antigens of Pseudomonas aeruginosa and Yersinia pestis. J Biol Chem,283, 23940–23949.

Gerard HC, Krausse-Opatz B, Wang Z, Rudy D, Rao JP, Zeidler H,Schumacher HR, Whittum-Hudson JA, Kohler L, Hudson AP. (2001).Expression of Chlamydia trachomatis genes encoding productsrequired for DNA synthesis and cell division during active versuspersistent infection. Mol Microbiol, 41, 731–741.

Gombos I, Kiss E, Detre C, Laszlo G, Matko J. (2006). Cholesterol andsphingolipids as lipid organizers of the immune cells’ plasmamembrane: their impact on the functions of MHC molecules, effectorT-lymphocytes and T-cell death. Immunol Lett, 104, 59–69.

Gomez HF, Ochoa TJ, Carlin LG, Cleary TG. (2003). Humanlactoferrin impairs virulence of Shigella flexneri. J Infect Dis, 187,87–95.

Goure J, Broz P, Attree O, Cornelis GR, Attree I. (2005). Protective anti-V antibodies inhibit Pseudomonas and Yersinia translocon assemblywithin host membranes. J Infect Dis, 192, 218–225.

Grieshaber S, Swanson JA, Hackstadt T. (2002). Determination of thephysical environment within the Chlamydia trachomatis inclusionusing ion-selective ratiometric probes. Cell Microbiol, 4, 273–283.

Gump DW. (1996). Antimicrobial susceptibility testing for some atypicalmicroorganisms: chlamydiae, mycoplasmas, rickettsia and spiro-chetes. Baltimore: Williams & Wilkins.

Hackstadt T, Fischer ER, Scidmore MA, Rockey DD, Heinzen RA.(1997). Origins and functions of the chlamydial inclusion. TrendsMicrobiol, 5, 288–293.

Hackstadt T, Todd WJ, Caldwell HD. (1985). Disulfide-mediatedinteractions of the chlamydial major outer membrane protein: rolein the differentiation of chlamydiae? J Bacteriol, 161, 25–31.

Hackstadt T, Rockey DD, Heinzen RA, Scidmore MA. (1996).Chlamydia trachomatis interrupts an exocytic pathway to acquireendogenously synthesized sphingomyelin in transit from the Golgiapparatus to the plasma membrane. EMBO J, 15, 964–977.

Hackstadt T, Scidmore-Carlson MA, Shaw EI, Fischer ER. (1999). TheChlamydia trachomatis IncA protein is required for homotypic vesiclefusion. Cell Microbiol, 1, 119–130.

Harley R, Herring A, Egan K, Howard P, Gruffydd-Jones T, Azuma Y,Shirai M, Helps C. (2007). Molecular characterisation of 12Chlamydophila felis polymorphic membrane protein genes. VetMicrobiol, 124, 230–238.

Hatch TP, Al-Hossainy E, Silverman JA. (1982). Adenine nucleotideand lysine transport in Chlamydia psittaci. J Bacteriol, 150,662–670.

Hatch TP, Allan I, Pearce JH. (1984). Structural and polypeptidedifferences between envelopes of infective and reproductive life cycleforms of Chlamydia spp. J Bacteriol, 157, 13–20.

Hatch TP. (1996). Disulfide cross-linked envelope proteins: the func-tional equivalent of peptidoglycan in chlamydiae? J Bacteriol, 178,1–5.

Heinzen RA, Scidmore MA, Rockey DD, Hackstadt T. (1996).Differential interaction with endocytic and exocytic pathways distin-guish parasitophorous vacuoles of Coxiella burnetii and Chlamydiatrachomatis. Infect Immun, 64, 796–809.

Heinzen RA, Hackstadt T. (1997). The Chlamydia trachomatisparasitophorous vacuolar membrane is not passively permeable tolow-molecular-weight compounds. Infect Immun, 65, 1088–1094.

Helms JB, Zurzolo C. (2004). Lipids as targeting signals: lipid rafts andintracellular trafficking. Traffic, 5, 247–254.

Henderson IR, Lam AC. (2001). Polymorphic proteins of Chlamydiaspp.–autotransporters beyond the Proteobacteria. Trends Microbiol, 9,573–578.

Henke JM, Bassler BL. (2004). Three parallel quorum-sensing systemsregulate gene expression in Vibrio harveyi. J Bacteriol, 186,6902–6914.

Hentzer M, Wu H, Andersen JB, Riedel K, Rasmussen TB, Bagge N,Kumar N, Schembri MA, Song Z, Kristoffersen P, Manefield M,Costerton JW, Molin S, Eberl L, Steinberg P, Kjelleberg S, Høiby N,Givskov M. (2003). Attenuation of Pseudomonas aeruginosa virulenceby quorum sensing inhibitors. EMBO J, 22, 3803–3815.

Higgins DA, Pomianek ME, Kraml CM, Taylor RK, Semmelhack MF,Bassler BL. (2007). The major Vibrio cholerae autoinducer and itsrole in virulence factor production. Nature, 450, 883–886.

Hodinka RL, Davis CH, Choong J, Wyrick PB. (1988). Ultrastructuralstudy of endocytosis of Chlamydia trachomatis by McCoy cells. InfectImmun, 56, 1456–1463.

Hogan RJ, Mathews SA, Mukhopadhyay S, Summersgill JT, Timms P.(2004). Chlamydial persistence: beyond the biphasic paradigm. InfectImmun, 72, 1843–1855.

Hughes DT, Sperandio V. (2008). Inter-kingdom signalling: communi-cation between bacteria and their hosts. Nat Rev Microbiol, 6,111–120.

Hybiske K, Stephens RS. (2007). Mechanisms of Chlamydia trachomatisentry into nonphagocytic cells. Infect Immun, 75, 3925–3934.

Iliffe-Lee ER, McClarty G. (1999). Glucose metabolism in Chlamydiatrachomatis: the ‘energy parasite’ hypothesis revisited. Mol Microbiol,33, 177–187.

Jewett TJ, Fischer ER, Mead DJ, Hackstadt T. (2006). Chlamydial TARPis a bacterial nucleator of actin. Proc Natl Acad Sci USA, 103,15599–15604.

Johnson FW, Spencer WN. (1983). Multiantibiotic resistance inChlamydia psittaci from ducks. Vet Rec, 112, 208.

Jones RB, Van der Pol B, Martin DH, Shepard MK. (1990). Partialcharacterization of Chlamydia trachomatis isolates resistant to mul-tiple antibiotics. J Infect Dis, 162, 1309–1315.

Kalman S, Mitchell W, Marathe R, Lammel C, Fan J, Hyman RW,Olinger L, Grimwood J, Davis RW, Stephens RS. (1999). Comparativegenomes of Chlamydia pneumoniae and C. trachomatis. Nat Genet,21, 385–389.

Kaneko Y, Thoendel M, Olakanmi O, Britigan BE, Singh PK. (2007).The transition metal gallium disrupts Pseudomonas aeruginosa ironmetabolism and has antimicrobial and antibiofilm activity. J ClinInvest, 117, 877–888.

Kauppi AM, Nordfelth R, Uvell H, Wolf-Watz H, Elofsson M. (2003).Targeting bacterial virulence: inhibitors of type III secretion inYersinia. Chem Biol, 10, 241–249.

Kendall MM, Rasko DA, Sperandio V. (2007). Global effects of the cell-to-cell signaling molecules autoinducer-2, autoinducer-3, and epi-nephrine in a luxS mutant of enterohemorrhagic Escherichia coli.Infect Immun, 75, 4875–4884.

Keyser P, Elofsson M, Rosell S, Wolf-Watz H. (2008). Virulenceblockers as alternatives to antibiotics: type III secretion inhibitorsagainst Gram-negative bacteria. J Intern Med, 264, 17–29.

Kim SK, DeMars R. (2001). Epitope clusters in the major outermembrane protein of Chlamydia trachomatis. Curr Opin Immunol, 13,429–436.

Klingenberg M. (1989). Molecular aspects of the adenine nucleotidecarrier from mitochondria. Arch Biochem Biophys, 270, 1–14.

Koo IC, Stephens RS. (2003). A developmentally regulated two-component signal transduction system in Chlamydia. J Biol Chem,278, 17314–17319.

Lane BJ, Mutchler C, Al Khodor S, Grieshaber SS, Carabeo RA. (2008).Chlamydial entry involves TARP binding of guanine nucleotideexchange factors. PLoS Pathog, 4, e1000014.

Lefevre JC, Lepargneur JP. (1998). Comparative in vitro susceptibility ofa tetracycline-resistant Chlamydia trachomatis strain isolated inToulouse (France). Sex Transm Dis, 25, 350–352.

Lillemeier BF, Pfeiffer JR, Surviladze Z, Wilson BS, Davis MM. (2006).Plasma membrane-associated proteins are clustered into islandsattached to the cytoskeleton. Proc Natl Acad Sci USA, 103,18992–18997.

Linington RG, Robertson M, Gauthier A, Finlay BB, van Soest R,Andersen RJ. (2002). Caminoside A, an antimicrobial glycolipidisolated from the marine sponge Caminus sphaeroconia. Org Lett, 4,4089–4092.

Linington RG, Robertson M, Gauthier A, Finlay BB, MacMillan JB,Molinski TF, van Soest R, Andersen RJ. (2006). Caminosides B-D,antimicrobial glycolipids isolated from the marine sponge Caminussphaeroconia. J Nat Prod, 69, 173–177.

Linka N, Hurka H, Lang BF, Burger G, Winkler HH, Stamme C, UrbanyC, Seil I, Kusch J, Neuhaus HE. (2003). Phylogenetic relationships ofnon-mitochondrial nucleotide transport proteins in bacteria andeukaryotes. Gene, 306, 27–35.

Litwin J, Officer JE, Brown A, Moulder JW. (1961). A comparativestudy of the growth cycles of different members of the psittacosisgroup in different host cells. J Infect Dis, 109, 251–279.

Liu X, Afrane M, Clemmer DE, Zhong G, Nelson DE. (2010).Identification of Chlamydia trachomatis outer membrane complexproteins by differential proteomics. J Bacteriol, 192, 2852–2860.

Longbottom D, Russell M, Jones GE, Lainson FA, Herring AJ. (1996).Identification of a multigene family coding for the 90 kDa proteins of

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 325

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 14: Chlamydial biology and its associated virulence blockers

the ovine abortion subtype of Chlamydia psittaci. FEMS MicrobiolLett, 142, 277–281.

Longbottom D, Coulter LJ. (2003). Animal chlamydioses and zoonoticimplications. J Comp Pathol, 128, 217–244.

Lyon GJ, Novick RP. (2004). Peptide signaling in Staphylococcus aureusand other Gram-positive bacteria. Peptides, 25, 1389–1403.

Markham AP, Barrett BS, Esfandiary R, Picking WL, Picking WD, JoshiSB, Middaugh CR. (2010). Formulation and immunogenicity of apotential multivalent type III secretion system-based protein vaccine.J Pharm Sci, 99, 4497–4509.

Mathews S, George C, Flegg C, Stenzel D, Timms P. (2001). Differentialexpression of ompA, ompB, pyk, nlpD and Cpn0585 genes betweennormal and interferon-gamma treated cultures of Chlamydia pneu-moniae. Microb Pathog, 30, 337–345.

Matsumoto A, Fujiwara E, Higashi N. (1976). Observations of thesurface projections of infectious small cell of Chlamydia psittaci inthin sections. J Electron Microsc (Tokyo), 25, 169–170.

Matsumoto A. (1982a). Electron microscopic observations of surfaceprojections on Chlamydia psittaci reticulate bodies. J Bacteriol, 150,358–364.

Matsumoto A. (1982b). Surface projections of Chlamydia psittacielementary bodies as revealed by freeze-deep-etching. J Bacteriol,151, 1040–1042.

McCoy AJ, Maurelli AT. (2006). Building the invisible wall: updatingthe chlamydial peptidoglycan anomaly. Trends Microbiol, 14, 70–77.

McDowell P, Affas Z, Reynolds C, Holden MT, Wood SJ, Saint S,Cockayne A, Hill PJ, Dodd CE, Bycroft BW, Chan WC, Williams P.(2001). Structure, activity and evolution of the group I thiolactonepeptide quorum-sensing system of Staphylococcus aureus. MolMicrobiol, 41, 503–512.

Misyurina OY, Chipitsyna EV, Finashutina YP, Lazarev VN, AkopianTA, Savicheva AM, Govorun VM. (2004). Mutations in a 23S rRNAgene of Chlamydia trachomatis associated with resistance tomacrolides. Antimicrob Agents Chemother, 48, 1347–1349.

Moelleken K, Hegemann JH. (2008). The Chlamydia outer membraneprotein OmcB is required for adhesion and exhibits biovar-specificdifferences in glycosaminoglycan binding. Mol Microbiol, 67,403–419.

Moulder JW. (1991). Interaction of chlamydiae and host cells in vitro.Microbiol Rev, 55, 143–190.

Moulder JW. (1993). Why is Chlamydia sensitive to penicillin in theabsence of peptidoglycan? Infect Agents Dis, 2, 87–99.

Moulin G, Cavalie P, Pellanne I, Chevance A, Laval A, Millemann Y,Colin P, Chauvin C; Antimicrobial Resistance ad hoc Group of theFrench Food Safety Agency. (2008). A comparison of antimicrobialusage in human and veterinary medicine in France from 1999 to 2005.J Antimicrob Chemother, 62, 617–625.

Mpiga P, Ravaoarinoro M. (2006). Chlamydia trachomatis persistence:an update. Microbiol Res, 161, 9–19.

Mueller CA, Broz P, Cornelis GR. (2008). The type III secretion systemtip complex and translocon. Mol Microbiol, 68, 1085–1095.

Muschiol S, Bailey L, Gylfe A, Sundin C, Hultenby K, Bergstrom S,Elofsson M, Wolf-Watz H, Normark S, Henriques-Normark B. (2006).A small-molecule inhibitor of type III secretion inhibits differentstages of the infectious cycle of Chlamydia trachomatis. Proc NatlAcad Sci USA, 103, 14566–14571.

Muschiol S, Normark S, Henriques-Normark B, Subtil A. (2009). Smallmolecule inhibitors of the Yersinia type III secretion system impairthe development of Chlamydia after entry into host cells. BMCMicrobiol, 9, 75.

Newhall WJ, Jones RB. (1983). Disulfide-linked oligomers of the majorouter membrane protein of chlamydiae. J Bacteriol, 154, 998–1001.

Newhall 5th WJ. (1987). Biosynthesis and disulfide cross-linking ofouter membrane components during the growth cycle of Chlamydiatrachomatis. Infect Immun, 55, 162–168.

Novick RP. (2003). Autoinduction and signal transduction in theregulation of staphylococcal virulence. Mol Microbiol, 48,1429–1449.

Ochoa TJ, Noguera-Obenza M, Ebel F, Guzman CA, Gomez HF, ClearyTG. (2003). Lactoferrin impairs type III secretory system function inenteropathogenic Escherichia coli. Infect Immun, 71, 5149–5155.

Ochoa TJ, Clearly TG. (2004). Lactoferrin disruption of bacterial type IIIsecretion systems. Biometals, 17, 257–260.

Pan J, Ren D. (2009). Quorum sensing inhibitors: a patent overview.Expert Opin Ther Pat, 19, 1581–1601.

Paradkar PN, De Domenico I, Durchfort N, Zohn I, Kaplan J, Ward DM.(2008). Iron depletion limits intracellular bacterial growth in macro-phages. Blood, 112, 866–874.

Parton RG, Richards AA. (2003). Lipid rafts and caveolae as portals forendocytosis: new insights and common mechanisms. Traffic, 4,724–738.

Peeling R, Maclean IW, Brunham RC. (1984). In vitro neutralization ofChlamydia trachomatis with monoclonal antibody to an epitope on themajor outer membrane protein. Infect Immun, 46, 484–488.

Peters J, Wilson DP, Myers G, Timms P, Bavoil PM. (2007). Type IIIsecretion a la Chlamydia. Trends Microbiol, 15, 241–251.

Peterson EM, Cheng X, Markoff BA, Fielder TJ, de la Maza LM. (1991).Functional and structural mapping of Chlamydia trachomatis species-specific major outer membrane protein epitopes by use of neutralizingmonoclonal antibodies. Infect Immun, 59, 4147–4153.

Peterson JR, Mitchison TJ. (2002). Small molecules, big impact: ahistory of chemical inhibitors and the cytoskeleton. Chem Biol, 9,1275–1285.

Philipovskiy AV, Cowan C, Wulff-Strobel CR, Burnett SH, Kerschen EJ,Cohen DA, Kaplan AM, Straley SC. (2005). Antibody against Vantigen prevents Yop-dependent growth of Yersinia pestis. InfectImmun, 73, 1532–1542.

Prain CJ, Pearce JH. (1989). Ultrastructural studies on the intracellularfate of Chlamydia psittaci (strain guinea pig inclusion conjunctivitis)and Chlamydia trachomatis (strain lymphogranuloma venereum 434):modulation of intracellular events and relationship with endocyticmechanism. J Gen Microbiol, 135, 2107–2123.

Prantner D, Nagarajan UM. (2009). Role for the chlamydial type IIIsecretion apparatus in host cytokine expression. Infect Immun, 77,76–84.

Rasko DA, Moreira CG, Li de R, Reading NC, Ritchie JM, Waldor MK,Williams N, Taussig R, Wei S, Roth M, Hughes DT, Huntley JF, FinaMW, Falck JR, Sperandio V. (2008). Targeting QseC signaling andvirulence for antibiotic development. Science, 321, 1078–1080.

Rasko DA, Sperandio V. (2010). Anti-virulence strategies to combatbacteria-mediated disease. Nat Rev Drug Discov, 9, 117–128.

Rasmussen-Lathrop SJ, Koshiyama K, Phillips N, Stephens RS. (2000).Chlamydia-dependent biosynthesis of a heparan sulphate-like com-pound in eukaryotic cells. Cell Microbiol, 2, 137–144.

Rasmussen TB, Bjarnsholt T, Skindersoe ME, Hentzer M, KristoffersenP, Kote M, Nielsen J, Eberl L, Givskov M. (2005a). Screening forquorum-sensing inhibitors (QSI) by use of a novel genetic system, theQSI selector. J Bacteriol, 187, 1799–1814.

Rasmussen TB, Skindersoe ME, Bjarnsholt T, Phipps RK, ChristensenKB, Jensen PO, Andersen JB, Koch B, Larsen TO, Hentzer M, EberlL, Hoiby N, Givskov M. (2005b). Identity and effects of quorum-sensing inhibitors produced by Penicillium species. Microbiology(Reading, Engl), 151, 1325–1340.

Raulston JE, Davis CH, Schmiel DH, Morgan MW, Wyrick PB. (1993).Molecular characterization and outer membrane association of aChlamydia trachomatis protein related to the hsp70 family of proteins.J Biol Chem, 268, 23139–23147.

Raulston JE. (1997). Response of Chlamydia trachomatis serovar E toiron restriction in vitro and evidence for iron-regulated chlamydialproteins. Infect Immun, 65, 4539–4547.

Ravn L, Christensen AB, Molin S, Givskov M, Gram L. (2001). Methodsfor detecting acylated homoserine lactones produced by Gram-negative bacteria and their application in studies of AHL-productionkinetics. J Microbiol Methods, 44, 239–251.

Read TD, Myers GS, Brunham RC, Nelson WC, Paulsen IT, HeidelbergJ, Holtzapple E, Khouri H, Federova NB, Carty HA, Umayam LA,Haft DH, Peterson J, Beanan MJ, White O, Salzberg SL, Hsia RC,McClarty G, Rank RG, Bavoil PM, Fraser CM. (2003). Genomesequence of Chlamydophila caviae (Chlamydia psittaci GPIC):examining the role of niche-specific genes in the evolution of theChlamydiaceae. Nucleic Acids Res, 31, 2134–2147.

Ren Q, Kang KH, Paulsen IT. (2004). TransportDB: a relational databaseof cellular membrane transport systems. Nucleic Acids Res, 32,D284–D288.

Reynolds DJ, Pearce JH. (1990). Characterization of the cytochalasinD-resistant (pinocytic) mechanisms of endocytosis utilized bychlamydiae. Infect Immun, 58, 3208–3216.

Ridderhof JC, Barnes RC. (1989). Fusion of inclusions followingsuperinfection of HeLa cells by two serovars of Chlamydiatrachomatis. Infect Immun, 57, 3189–3193.

326 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 15: Chlamydial biology and its associated virulence blockers

Rockey DD, Fischer ER, Hackstadt T. (1996). Temporal analysis of thedeveloping Chlamydia psittaci inclusion by use of fluorescence andelectron microscopy. Infect Immun, 64, 4269–4278.

Rockey DD, Matsumoto A. (2000). The chlamydial developmental cycle.Washington DC: ASM Press.

Rockey DD, Scidmore MA, Bannantine JP, Brown WJ. (2002).Proteins in the chlamydial inclusion membrane. Microbes Infect, 4,333–340.

Rzomp KA, Moorhead AR, Scidmore MA. (2006). The GTPase Rab4interacts with Chlamydia trachomatis inclusion membrane proteinCT229. Infect Immun, 74, 5362–5373.

Saier Jr MH. (2000). A functional-phylogenetic classification system fortransmembrane solute transporters. Microbiol Mol Biol Rev, 64,354–411.

Salmond GP, Bycroft BW, Stewart GS, Williams P. (1995). The bacterial‘enigma’: cracking the code of cell-cell communication. MolMicrobiol, 16, 615–624.

Sanchez-Campillo M, Bini L, Comanducci M, Raggiaschi R, MarzocchiB, Pallini V, Ratti G. (1999). Identification of immunoreactiveproteins of Chlamydia trachomatis by Western blot analysis of a two-dimensional electrophoresis map with patient sera. Electrophoresis,20, 2269–2279.

Sandoz KM, Rockey DD. (2010). Antibiotic resistance in Chlamydiae.Future Microbiol, 5, 1427–1442.

Sardinia LM, Segal E, Ganem D. (1988). Developmental regulation ofthe cysteine-rich outer-membrane proteins of murine Chlamydiatrachomatis. J Gen Microbiol, 134, 997–1004.

Sarmah AK, Meyer MT, Boxall AB. (2006). A global perspective on theuse, sales, exposure pathways, occurrence, fate and effects ofveterinary antibiotics (VAs) in the environment. Chemosphere, 65,725–759.

Schauder S, Shokat K, Surette MG, Bassler BL. (2001). The LuxS familyof bacterial autoinducers: biosynthesis of a novel quorum-sensingsignal molecule. Mol Microbiol, 41, 463–476.

Schautteet K, De Clercq E, Miry C, Van Groenweghe F, DelavaP, Kalmar I, Vanrompay D. (2012). Chlamydia suis and reproductivefailure in Belgian, Cypriote, German and Israeli pigs. Journal ofMedical Microbiology (in press).

Schramm N, Bagnell CR, Wyrick PB. (1996). Vesicles containingChlamydia trachomatis serovar L2 remain above pH 6 within HEC-1Bcells. Infect Immun, 64, 1208–1214.

Schwarzenbacher R, Stenner-Liewen F, Liewen H, Robinson H,Yuan H, Bossy-Wetzel E, Reed JC, Liddington RC. (2004).Structure of the Chlamydia protein CADD reveals a redoxenzyme that modulates host cell apoptosis. J Biol Chem, 279,29320–29324.

Scidmore MA, Fischer ER, Hackstadt T. (1996). Sphingolipids andglycoproteins are differentially trafficked to the Chlamydia tracho-matis inclusion. J Cell Biol, 134, 363–374.

Scidmore MA, Fischer ER, Hackstadt T. (2003). Restricted fusion ofChlamydia trachomatis vesicles with endocytic compartments duringthe initial stages of infection. Infect Immun, 71, 973–984.

Shaw EI, Dooley CA, Fischer ER, Scidmore MA, Fields KA, HackstadtT. (2000). Three temporal classes of gene expression during theChlamydia trachomatis developmental cycle. Mol Microbiol, 37,913–925.

Simons K, Toomre D. (2000). Lipid rafts and signal transduction. NatRev Mol Cell Biol, 1, 31–39.

Skoudy A, Mounier J, Aruffo A, Ohayon H, Gounon P, Sansonetti P,Tran Van Nhieu G. (2000). CD44 binds to the Shigella IpaB proteinand participates in bacterial invasion of epithelial cells. CellMicrobiol, 2, 19–33.

Slepenkin A, Enquist PA, Hagglund U, de la Maza LM, Elofsson M,Peterson EM. (2007). Reversal of the antichlamydial activity ofputative type III secretion inhibitors by iron. Infect Immun, 75,3478–3489.

Smith KM, Bu Y, Suga H. (2003a). Induction and inhibition ofPseudomonas aeruginosa quorum sensing by synthetic autoinduceranalogs. Chem Biol, 10, 81–89.

Smith KM, Bu Y, Suga H. (2003b). Library screening for syntheticagonists and antagonists of a Pseudomonas aeruginosa autoinducer.Chem Biol, 10, 563–571.

Sperandio V, Torres AG, Jarvis B, Nataro JP, Kaper JB. (2003). Bacteria-host communication: the language of hormones. Proc Natl Acad SciUSA, 100, 8951–8956.

Stenner-Liewen F, Liewen H, Zapata JM, Pawlowski K, Godzik A, ReedJC. (2002). CADD, a Chlamydia protein that interacts with deathreceptors. J Biol Chem, 277, 9633–9636.

Stephens RS, Kalman S, Lammel C, Fan J, Marathe R, Aravind L,Mitchell W, Olinger L, Tatusov RL, Zhao Q, Koonin EV, Davis RW.(1998). Genome sequence of an obligate intracellular pathogen ofhumans: Chlamydia trachomatis. Science, 282, 754–759.

Stephens RS, Koshiyama K, Lewis E, Kubo A. (2001). Heparin-bindingouter membrane protein of chlamydiae. Mol Microbiol, 40, 691–699.

Stephens RS, Lammel CJ. (2001). Chlamydia outer membrane proteindiscovery using genomics. Curr Opin Microbiol, 4, 16–20.

Stuart ES, Webley WC, Norkin LC. (2003). Lipid rafts, caveolae,caveolin-1, and entry by Chlamydiae into host cells. Exp Cell Res,287, 67–78.

Su H, Watkins NG, Zhang YX, Caldwell HD. (1990). Chlamydiatrachomatis-host cell interactions: role of the chlamydial major outermembrane protein as an adhesin. Infect Immun, 58, 1017–1025.

Su H, Raymond L, Rockey DD, Fischer E, Hackstadt T, Caldwell HD.(1996). A recombinant Chlamydia trachomatis major outer membraneprotein binds to heparan sulfate receptors on epithelial cells. Proc NatlAcad Sci USA, 93, 11143–11148.

Subtil A, Wyplosz B, Balana ME, Dautry-Varsat A. (2004). Analysis ofChlamydia caviae entry sites and involvement of Cdc42 and Racactivity. J Cell Sci, 117, 3923–3933.

Suchland RJ, Sandoz KM, Jeffrey BM, Stamm WE, Rockey DD. (2009).Horizontal transfer of tetracycline resistance among Chlamydia spp.in vitro. Antimicrob Agents Chemother, 53, 4604–4611.

Suga H, Smith KM. (2003). Molecular mechanisms of bacterial quorumsensing as a new drug target. Curr Opin Chem Biol, 7, 586–591.

Surette MG, Miller MB, Bassler BL. (1999). Quorum sensing inEscherichia coli, Salmonella typhimurium, and Vibrio harveyi: a newfamily of genes responsible for autoinducer production. Proc NatlAcad Sci USA, 96, 1639–1644.

Swanson KA, Crane DD, Caldwell HD. (2007). Chlamydia trachomatisspecies-specific induction of ezrin tyrosine phosphorylation functionsin pathogen entry. Infect Immun, 75, 5669–5677.

Taraska T, Ward DM, Ajioka RS, Wyrick PB, Davis-Kaplan SR, DavisCH, Kaplan J. (1996). The late chlamydial inclusion membrane is notderived from the endocytic pathway and is relatively deficient in hostproteins. Infect Immun, 64, 3713–3727.

Thomson NR, Yeats C, Bell K, Holden MT, Bentley SD, Livingstone M,Cerdeno-Tarraga AM, Harris B, Doggett J, Ormond D, Mungall K,Clarke K, Feltwell T, Hance Z, Sanders M, Quail MA, Price C, BarrellBG, Parkhill J, Longbottom D. (2005). The Chlamydophila abortusgenome sequence reveals an array of variable proteins that contributeto interspecies variation. Genome Res, 15, 629–640.

Ting LM, Hsia RC, Haidaris CG, Bavoil PM. (1995). Interaction of outerenvelope proteins of Chlamydia psittaci GPIC with the HeLa cellsurface. Infect Immun, 63, 3600–3608.

Tjaden J, Winkler HH, Schwoppe C, Van Der Laan M, Mohlmann T,Neuhaus HE. (1999). Two nucleotide transport proteins in Chlamydiatrachomatis, one for net nucleoside triphosphate uptake and the otherfor transport of energy. J Bacteriol, 181, 1196–1202.

Van Blarcom TJ, Sofer-Podesta C, Ang J, Boyer JL, Crystal RG,Georgiou G. (2010). Affinity maturation of an anti-V antigen IgGexpressed in situ through adenovirus gene delivery confers enhancedprotection against Yersinia pestis challenge. Gene Ther, 17, 913–921.

Van Droogenbroeck C, Beeckman DS, Harkinezhad T, Cox E,Vanrompay D. (2008). Evaluation of the prophylactic use ofovotransferrin against chlamydiosis in SPF turkeys. Vet Microbiol,132, 372–378.

Van Droogenbroeck C, Dossche L, Wauman T, Van Lent S, Phan TT,Beeckman DS, Vanrompay D. (2011). Use of ovotransferrin as anantimicrobial in turkeys naturally infected with Chlamydia psittaci,avian metapneumovirus and Ornithobacterium rhinotracheale.Vet Microbiol, 153, 257–263.

Vanrompay D, Charlier G, Ducatelle R, Haesebrouck F. (1996).Ultrastructural changes in avian Chlamydia psittaci serovar A-, B-,and D-infected Buffalo Green Monkey cells. Infect Immun, 64,1265–1271.

Vanrompay D, Mast J, Ducatelle R, Haesebrouck F, Goddeeris B. (1995).Chlamydia psittaci in turkeys: pathogenesis of infections in avianserovars A, B and D. Vet Microbiol, 47, 245–256.

Wang J, Chen L, Chen F, Zhang X, Zhang Y, Baseman J, Perdue S, YehIT, Shain R, Holland M, Bailey R, Mabey D, Yu P, Zhong G. (2009).

DOI: 10.3109/1040841X.2012.726210 Chlamydial biology and virulence blockers 327

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.

Page 16: Chlamydial biology and its associated virulence blockers

A chlamydial type III-secreted effector protein (Tarp) is predomin-antly recognized by antibodies from humans infected with Chlamydiatrachomatis and induces protective immunity against upper genitaltract pathologies in mice. Vaccine, 27, 2967–2980.

Wang D, Zetterstrom CE, Gabrielsen M, Beckham KS, Tree JJ,Macdonald SE, Byron O, Mitchell TJ, Gally DL, Herzyk P,Mahajan A, Uvell H, Burchmore R, Smith BO, Elofsson M, RoeAJ. (2011). Identification of bacterial target proteins for thesalicylidene acylhydrazide class of virulence-blocking compounds.J Biol Chem, 286, 29922–29931.

Ward ME. (1988). The Chlamydial developmental cycle. In: Barron AL,ed. Microbiology of Chlamydia. Boca Raton: CRC Press, Inc.

Warrell Jr RP, Bockman RS. (1989). Gallium in the treatment ofhypercalcemia and bone metastasis. Important Adv Oncol, 1989,205–220.

Watarai M, Funato S, Sasakawa C. (1996). Interaction of Ipa proteins ofShigella flexneri with a5b1 integrin promotes entry of the bacteriainto mammalian cells. J Exp Med, 183, 991–999.

Wehrl W, Brinkmann V, Jungblut PR, Meyer TF, Szczepek AJ. (2004).From the inside out–processing of the Chlamydial autotransporterPmpD and its role in bacterial adhesion and activation of human hostcells. Mol Microbiol, 51, 319–334.

Wilson DP, Timms P, McElwain DL, Bavoil PM. (2006). Type IIIsecretion, contact-dependent model for the intracellular developmentof chlamydia. Bull Math Biol, 68, 161–178.

Winkler HH, Neuhaus HE. (1999). Non-mitochondrial ATP transport.Trends Biochem Sci, 24, 64–68.

Wolf K, Betts HJ, Chellas-Gery B, Hower S, Linton CN, Fields KA.(2006). Treatment of Chlamydia trachomatis with a small moleculeinhibitor of the Yersinia type III secretion system disrupts progressionof the chlamydial developmental cycle. Mol Microbiol, 61,1543–1555.

Wyrick PB, Choong J, Davis CH, Knight ST, Royal MO, MaslowAS, Bagnell CR. (1989). Entry of genital Chlamydia trachomatisinto polarized human epithelial cells. Infect Immun, 57,2378–2389.

Wyrick PB. (2000). Intracellular survival by Chlamydia. Cell Microbiol,2, 275–282.

Yekta MA, Verdonck F, Van Den Broeck W, Goddeeris BM, Cox E,Vanrompay D. (2010). Lactoferrin inhibits E. coli O157:H7 growthand attachment to intestinal epithelial cells. Vet Med, 55, 359–368.

Yuan Y, Zhang YX, Watkins NG, Caldwell HD. (1989). Nucleotide anddeduced amino acid sequences for the four variable domains of themajor outer membrane proteins of the 15 Chlamydia trachomatisserovars. Infect Immun, 57, 1040–1049.

Zavil’gel’skii GB, Manukhov IV. (2001). [‘‘Quorum sensing’’, orhow bacteria ‘‘talk’’ to each other]. Mol Biol (Mosk), 35,268–277.

Zhang JP, Stephens RS. (1992). Mechanism of C. trachomatis attachmentto eukaryotic host cells. Cell, 69, 861–869.

328 D. Beeckman et al. Crit Rev Microbiol, 2014; 40(4): 313–328

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y R

MIT

Uni

vers

ity o

n 06

/07/

14Fo

r pe

rson

al u

se o

nly.


Recommended