This article was downloaded by: [University of Southern Queensland] On: 13 March 2015, At: 10:01 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Small GTPases Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/ksgt20

Bacterial factors exploit eukaryotic Rho GTPase signaling cascades to promote invasion and proliferation within their host Michel R Popoff

a

a

Unité des Bactéries Anaérobies et Toxines; Institut Pasteur; Paris, France Published online: 31 Oct 2014.

Click for updates To cite this article: Michel R Popoff (2014) Bacterial factors exploit eukaryotic Rho GTPase signaling cascades to promote invasion and proliferation within their host, Small GTPases, 5:3, 1-18, DOI: 10.4161/sgtp.28209 To link to this article: http://dx.doi.org/10.4161/sgtp.28209

PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http:// www.tandfonline.com/page/terms-and-conditions

REVIEW Small GTPases 5:3, 1--18; November 1, 2014; © 2014 Taylor & Francis Group, LLC

Bacterial factors exploit eukaryotic Rho GTPase signaling cascades to promote invasion and proliferation within their host Michel R Popoff* Unite des Bacteries Anaerobies et Toxines; Institut Pasteur; Paris, France

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

Keywords: toxin, virulence factor, RhoGTPase, actin cytoskeleton, phagocytosis, ADP-ribosylation, adenylation, glucosylation, innate immunity

Actin cytoskeleton is a main target of many bacterial pathogens. Among the multiple regulation steps of the actin cytoskeleton, bacterial factors interact preferentially with RhoGTPases. Pathogens secrete either toxins which diffuse in the surrounding environment, or directly inject virulence factors into target cells. Bacterial toxins, which interfere with RhoGTPases, and to some extent with RasGTPases, catalyze a covalent modification (ADPribosylation, glucosylation, deamidation, adenylation, proteolysis) blocking these molecules in their active or inactive state, resulting in alteration of epithelial and/or endothelial barriers, which contributes to dissemination of bacteria in the host. Injected bacterial virulence factors preferentially manipulate the RhoGTPase signaling cascade by mimicry of eukaryotic regulatory proteins leading to local actin cytoskeleton rearrangement, which mediates bacterial entry into host cells or in contrast escape to phagocytosis and immune defenses. Invasive bacteria can also manipulate RhoGTPase signaling through recognition and stimulation of cell surface receptor (s). In addition, changes in RhoGTPase activation state is sensed by the innate immunity pathways and allows the host cell to adapt an appropriate defense response.

body. Pathogens also manipulate the host actin cytoskeleton to enter non-phagocytic target cells and then to escape the immune defenses, or in contrast to avoid phagocytosis by macrophages. Actin cytoskeleton is a highly dynamic structure which is regulated by numerous proteins and factors including GTPases from the Rho family which play a key role. Due to long interactive life of bacteria with higher organisms, pathogens have engineered various strategies to attack or to survive in hostile eukaryotic cell environment. Certain pathogens secrete potent toxins which diffuse in the surrounding environment and induce tissue destruction allowing bacterial growth and dissemination. Other bacteria inject directly into the target cells virulence factors, which hijack the actin cytoskeleton machinery for their own profit, notably to escape the immune defenses. Excellent publications have described the multifaceted interactions between bacterial effectors and RhoGTPases.1-10 The bacterial effectors which manipulate the eukaryotic GTPases are also called modulins.11 This review is focused on the diversity of RhoGTPase manipulations by bacterial pathogens to disseminate and survive in host.

Rho-GTPase Manipulation by Bacterial Toxins and Alteration of Cell Barriers Introduction Actin cytoskeleton is a major cell structure which is involved in multiple functions and is one of the main targets of bacterial pathogens. Indeed, the actin cytoskeleton controls cellular functions critical for eukaryotic life including locomotion, endocytosis, exocytosis, trafficking of intracellular organelles as well as maintenance of intercellular connections. Since intercellular junctions are essential in building cell barriers, like the epithelia and endothelia that protect an organism from the environment as well as delineate internal compartments, pathogen manipulation of host via the actin cytoskeleton provides an effective means to breach host’s containment capabilities and an access to essential nutrients from more appropriate niches throughout the *Correspondence to: Michel R Popoff; Email: [email protected] Submitted: 12/16/2013; Revised: 02/05/2014; Accepted: 02/13/2014 http://dx.doi.org/10.4161/sgtp.28209

www.landesbioscience.com

Toxigenic bacteria release in their environment potent toxins, which interact with the host tissues and are responsible for the lesions and symptoms of the diseases caused by these microorganisms. Various bacterial toxins manipulate the actin cytoskeleton either by interfering directly with actin monomers or by modulating RhoGTPases. Indeed, clostridial toxins from the Iota and C2 toxin families are proteins, which ADP-ribosylate actin monomers at Arg177 and thus impair actin monomer/monomer assembly. This results in loss of cellular actin filaments.12,13 Other toxins target specifically RhoGTPases that they covalently modify. This triggers an amplified effect on the actin cytoskeleton compared with the toxins active on the actin monomers, since RhoGTPases are upstream regulators of actin polymerization. Moreover, these toxins also alter other cellular functions controlled by RhoGTPases. Enzymatic modification of RhoGTPases by bacterial toxins results either in activation or inactivation of these molecules. Toxins, which inactivate RhoGTPases, add a cumbersome

Small GTPases

1

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

moiety (AMP, ADP-ribose, glucose, or N-acetyl-glucosamine) to a residue of switch I or in a close area, whereas activating toxins of RhoGTPases catalyze the ADP-ribosylation or deamidation of a critical glutamine of switch II in glutamic acid. RhoGTPases are active in the GTP-bound form and inactive in the GDPbound form. These proteins adopt a conformational change localized in two regions called switch I and switch II, when they are in GTP- or in GDP-bound form. The transition from GDP(inactive) to GTP-bound form (active) is enhanced by an exchange factor (GEF) in response to an external signal transduced by a membrane receptor. Switch I interacts with the downstream effectors and switch II is involved in the GTPase activity that is stimulated by a GTPase activating protein (GAP). GTPases act as molecular switches between a membrane receptor activated by an external factor and downstream effectors. RhoGTPases bound to GDP are localized in the cytosol in association with a protein called guanine dissociation inhibitor (GDI), and they translocate to the membrane after dissociation of the Rho-GDI complex possibly mediated by ERM (EzrinRadixin-Moesin) proteins. At the membrane, Rho is activated by GEFs and interacts with its effectors14-16 (Fig. 1). Indeed, modification of residues in switch I impair RhoGTPase binding to

their effectors and RhoGTPases are blocked in their inactive state, whereas alteration of functional residue in switch II prevents their GTPase activity yielding permanent active molecules (Fig. 2 and 3).

RhoGTPase inactivating toxins and endothelial barrier permeability C3 exoenzyme is produced by Clostridium botulinum type C and D, and C3 related exoenzymes are also synthesized by Clostridium limosum, Bacillus cereus, Bacillus thuringiensis and Staphylococcus aureus in which it is called epidermal cell differentiation inhibitor (EDIN).1,17 C3 from C. botulinum was the first toxin, which has been found to interact with Rho proteins and was of a great interest to elucidate their function on the control of actin polymerization. All C3 exoenzymes recognize RhoA, B and C, and in addition, EDIN also modifies RhoE.18 C3 exoenzymes are small proteins (about 28 kDa) which only possess a catalytic domain and lack the binding and translocation domains permitting their entry into cells. The crystal structure shows that C3 consists of a core structure of five antiparallel b-strands packed against a three-stranded antiparallel b-sheet, which is flanked by four consecutive a-helices.19,20 Interestingly, the C3 structure is similar to that of the catalytic domain of the actin ADP-ribosylating toxins such as C. perfringens Iota toxin and Bacillus vegetative insecticidal protein (VIP).13,19,21 Although there is no significant overall sequence homology with other ADP-ribosylating toxins, C3 retains the conserved NAD binding site and catalytic pocket which consists of an a-helix (a3 in C3) bent over the two antiparallel b-sheets forming a central cleft. The amino acid (Glu214) that has an essential role in ADP-ribosylation, is conserved in C3 exoenzymes.19,22,23 C3 ADP-ribosylates RhoA at Asn-41 which is localized on an extended stretch close to the switch I. Rho-GDP is a preferential substrate for C3 as Rho-Asn41 is solvent accessible in the GDP structure.24 In contrast, the Asn41 residue of Rho found in a Rho-GDI-complex is hidden and thus resistant to C3-mediated ADP-ribosylation.25 ADP-ribosylation of Rho-Asn41 by C3 does not impair GDP/GTP exchange, does not affect intrinsic and GAP-stimulated Figure 1. RhoGTPase cycle. In the inactive GDP-bound form, RhoGTPases are localized in the cytosol in complex associated with a guanine nucleotide dissociation inhibitor (GDI), which prevents nucleoGTPase activity, and does not impinge tide exchange. Upon a cell signaling, a guanine nucleotide exchange factor (GEF) induces the release upon Rho interaction with its effecof GDP. Since the GTP concentration in the cytosol is 100 fold higher than that of GDP, RhoGTPases tors.26-28 However, C3 prevents GEF load GTP and become active. The active GTP-bound RhoGTPase adopts a conformational change of activation of Rho.29 In addition, ADPtwo surface loops named switch I (I) and switch II (II). Switch I is the main region for interactions with ribosylated Rho reassociates more effieffector molecules and switch II plays an important role in GTP catalysis. The intrinsic low GTPase activity of RhoGTPase is enhanced by GTPase-activating proteins (GAPs), thereby inactivating RhoGTciently with GDI than unmodified Rho, Pases. Guanine-nucleotide-dissociation inhibitors (GDIs) stabilize the GDP, and thereby the inactive thus causing an accumulation of inactive forms of RhoGTPases in the cytosol. Rho in the cytosol and preventing its

2

Small GTPases

Volume 5 Issue 3

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

translocation to the membrane and subsequent activation by GEFs as well as interaction with its effectors.30,31 Thereby, ADP-ribosylated Rho is trapped in a permanent inactive form in the cytosol, and subsequently degraded by the proteasome complex.29 C3 ADP-ribosylates the three isoforms RhoA, B and C. Most of the cellular effects described with this enzyme are related to RhoA. The first evidence that Rho is involved in the actin cytoskeleton organization comes from the initial study of C3 on Vero cells in which the effects are characterized by a cell rounding up and destruction of actin filaments.32 Since then, the effects of C3 on the actin cytoskeleton and related cellular functions are well documented. C3 induces a disorganization of the actin stress fibers, cell morphology change, alteration of epithelial and endothelial barrier function (mainly by perturbing tight junctions), impairment of endocytosis, exocytosis, phagocytosis, cytokinesis, neuronal plasticity, inhibition of cell cycle progression and migration of immune cells, as well as induction of apoptosis (rev in33-35). However, the role of C3 in natural disease such as botulism, is not known. C. botulinum can grow and produce toxins in the environment including contaminated food or in the intestinal lumen, and the passage of botulinum neurotoxin through the intestinal barrier and trafficking to the target motorneurons are responsible for the neurological symptoms of paralysis. C3 does not enter cells actively, since receptor binding and translocation domains are lacking. But, C3 enzymes are selectively internalized into macrophages and monocytes via acidic endosomes.36 Since C3 can inhibit Rho-mediated phagocytosis in macrophages,37 it may play an immunosuppressive role. In addition to its ADP-ribosylation activity, C3 exerts ADP-ribosylation-independent effects. C3 binds to RalA a GTPase from the Ras family, via a site adjacent but distinct from the catalytic site. C3 binding results in a stabilization of RalA in its GDP-bound and thus inactive conformation

www.landesbioscience.com

Figure 2. Inactivation of RhoGTPase signaling by bacterial toxins and virulence factors. Toxins inactivating RhoGTPases modify a residue of switch I by addition of a cumbersome moiety (ADPR by C3, glucose by large clostridial glucosylating toxins (LCGT) or P. asymbiotica toxin (PaTox), AMP by VOPs or IbpA) which impairs the interaction with downstream effector, or proteolytically cleaves RhoGTPase C-terminal part preventing its docking to the membrane and subsequent interaction with downstream effector. Virulence factors modify the RhoGTPase cycle signaling by mimicry eukaryotic regulatory proteins (GAP, GDI, GEF inhibitor).

Figure 3. Activation of RhoGTPase signaling by bacterial toxins and virulence factors. Toxins activating RhoGTPases modify a residue of switch II (deamidation, ADPribosylation) which impairs the intrinsic GAP activity thus yielding permanent active molecules. Bacterial virulence factors mimic GEFs and thus activate RhoGTPases.

Small GTPases

3

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

preventing its interaction with downstream effectors.38,39 The biological effects of C3 interaction with Ral remains to be elucidated. In contrast, EDIN which is produced by certain S. aureus strains, is considered as an important virulence factor, which is notably involved in impetigo, diabetic foot ulcers, and other skin infections.40-43 S. aureus can invade eukaryotic cells and release EDIN intracellularly, which contributes to actin cytoskeleton disorganization and tissue destruction.44 Thus, EDIN facilitates bacterial dissemination through the altered tissues. Indeed, in a mouse model EDIN promotes increased infection foci in deep tissues.45 An original effect triggered by EDIN and related C3 enzymes consists in the perturbation of endothelial permeability by formation of transcellular channels. Thereby, EDIN-mediated RhoA inactivation in endothelial cells induces a reorganization of the actin cytoskeleton resulting in the formation of transient macroapertures termed large transendothelial cell macroaperture tunnels (TEMs).4,46,47 The increased endothelial permeability facilitates the binding of S. aureus to extracellular matrix proteins and its dissemination from the blood stream to underlying tissues. RhoGTPase inactivating toxins and epithelial/endothelial barrier permeability as well as inflammatory response The manipulation of epithelial barrier permeability by RhoGTPase inactivating toxins is well illustrated with the large clostridial glucosylating toxins (LCGTs). These toxins are 250– 300 kDa proteins encompassing Clostridium difficile toxins A and B (TcdA and TcdB), Clostridium sordellii lethal toxin (TcsL), and hemorrhagic toxin (TcsH)), Clostridium novyi alpha-toxin TcnA), and C. perfringens TpeL (toxin C. perfringens large cytotoxin) (Table 1). LCGTs are single protein chains containing at least four functional domains. The one third C-terminal part exhibits multiple repeated sequences (31 short repeats and 7 long repeats in TcdA), which are involved in the recognition of a cell surface receptor. The central part contains a hydrophobic segment and probably mediates the translocation of the toxin across the membrane. The enzymatic site which is characterized by the DxD motif surrounded by a hydrophobic region, and the substrate recognition site are localized within the 543 N-terminal residues forming the enzymatic domain. The overall structure of this domain in TcdB, TcsL and TcnA is conserved and consists of a b-strain central core (about 235 amino acids) forming an active center pocket surrounded by numerous a-helices.48 The first aspartic residue of the DxD motif binds to ribosyl and glucosyl moieties of UDP-glucose and the second aspartic residues binds to divalent cation (mainly Mn2C) which increases the hydrolase activity and/or the binding of UDP-glucose.49 In addition, a cysteine protease domain (543–567) with DHC motif, lies in the vicinity of the autocleavage site.48 LCGTs enter cells by receptor-mediated endocytosis and release the enzymatic domain into the cytosol from acidified endosome by an auto-proteolytic activity stimulated by inositol hexakisphosphate.50-52 LCGTs catalyze the glucosylation of Rho- and/or Ras-GTPases from UDP-glucose, except TcnA, which uses UDP-N-acetylglucosamine as co-substrate. TcdA and

4

TcdB glucosylate Rho, Rac and Cdc42 at Thr-37, whereas TcsL glucosylates Ras at Thr-35, Rap, Ral and Rac at Thr-37.14,53 The large glucosylating clostridial toxins cleave the cosubstrate and transfer the glucose moiety to the acceptor amino acid of the Rho proteins.54-56 The conserved Thr, which is glucosylated, is located in switch I. Thr35/37 is involved in the coordination of Mg2C and subsequently to the binding of the b and g phosphates of GTP. The hydroxyl group of Thr35/37 is exposed to the surface of the molecule in its GDP-bound form, which is the only accessible substrate of glucosylating toxins. The nucleotide binding of the glucosylated G-protein Ras by TcsL is not grossly altered, but the GEF activation of GDP forms is decreased.57 Glucosylation of Thr35 completely prevents the recognition of the downstream effector, blocking the G-protein in the inactive form.57 The crystal structure of Ras modified by TcsL shows that glucosylation prevents the conformational change in the GTP state of the Ras effector loop, which is required for the interaction with the effector Raf.58 Similar results were found with RhoA glucosylated by TcdB.27 In addition, glucosylation of GTPase slightly reduces the intrinsic GTPase activity, completely inhibits GAP-stimulated GTP hydrolysis,57 and leads to accumulation of the GTP-bound form of Rho to the membrane where it is tightly bound.59 By inactivating Rho proteins, LCGTs induce cell rounding, with loss of actin stress fibers, reorganization of the cortical actin, disruption of the intercellular junctions and thus increase in cell barrier permeability. Rho is a major regulator of actin polymerization as well as of tight junction function, whereas Rac is mainly involved in the control of cortical actin and E-cadherin-dependent adherens junctions.10 Rac inactivation by LCGTs seems to be major player in actin cytoskeleton disorganization.60 TcdA and TcdB, which inactivate Rho, Rac and Cdc42, depolymerize apical and basal actin filaments and subsequently disorganize the ultrastructure and component distribution (ZO-1, ZO-2, occludin, claudin) of tight junctions, as well as perturb the organization of adherens junctions resulting in disruption of the epithelial barrier function.61-64 In addition, TcdA and TcdB induce cell detachment, cell death by apoptosis and necrosis, and a severe inflammatory response of the intestinal mucosa.65-68 Indeed, in intestinal epithelial and immune cells, TcdA and TcdB stimulate the secretion of proinflammatory cytokines (TNF-a, IL-1b, IL-6, IL-8, and many other mediators) mainly via activation of p38 MAP kinase.65,67,69-71 The activity of the C. difficile RhoGTPase inactivating toxins results in increased intestinal barrier permeability, fluid secretion, and destruction of the intestinal epithelium. The inflammatory response exacerbates the mucosal damage elicited by the C. difficile toxins. The rupture of the intestinal barrier integrity is a major initial mechanism in the C. difficile infections and facilitates the migration of neutrophils and other immune cells into the intestines. Toxigenic C. difficile colonizes first the intestine of susceptible patients mainly subsequently to a perturbation of the intestinal microbiota induced by antimicrobial agents72 and produces toxins which attack the intestinal barrier. Then, the resulting necrotic intestinal mucosa constitutes a favorable site of further C. difficile growth and subsequent release of toxins.

Small GTPases

Volume 5 Issue 3

Table 1. Examples of bacterial toxins that modulate Rho/RasGTPase signaling and alter epithelial/endothelial barriers Toxin

C3bot

C3lim C3cer C3stau-1, -2, -3 or EDIN-A, -B, -C

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

TcdA, TcdB

TcdBF

Pathogen

Target

Clostridium botulinum C and D

Toxins inhibiting RhoGTPases C3 exoenzymes RhoA, B, C ADP-ribosylation at N41 Cosubstrate NAD

Actin filament depolymerization, macroaperture

32

171 18,172-174

Large clostridial glucosdylating toxins RhoA, B, C, G Glucosylation at T37 of RhoA, Rac, Cdc42 T35 of Rac1 (and equivalent) TC10, TCL Cosubstrate UDP-glucose

Clostridium difficile

Clostridium sordellii

TcsH TcnA

Clostridium sordellii Clostridium novyi

TpeL

Clostridium difficileperfringens

Ras, Ral, Rap

PaTox

Photorhabdus asymbiotica

RhoA,B,C Rac1,2,3, Cdc42

MARTX

Vibrio cholerae, Vibrio sp.

Clostridium difficile 1470

Bordetella pertussis Bordetella parapertussis Bordetella bronchiseptica

53-55

53,56,176

177

N-acetyl-glucosamination at T37 of RhoA Cosubstrate UDP-Nacetylglucosamine N-acetyl-glucosamination and glucosylation at T35 of Ras N-acetyl-glucosamination at Y34 of RhoA, Y32 of Rac and Cdc42

Toxins activating RhoGTPases Deamidating toxins Rho, Rac, Cdc42 Deamidation, transglutamination at Q63 or Q61

Escherichia coli Yersinia pseudotuberculosis

Actin filament depolymerization Alteration of intercellular junctions Apoptosis, necrosis

175

Multifunctional-autoprocessing RTX toxins Rho, Rac, Cdc42 Inactivation (non yet defined mechanism)

Deamidation at Q61 or Q63

178

Apoptosis

179,180

Actin cytoskeleton disorganization Anti-phagocytosis Toxicity toward insect larvae

84

Actin filament depolymerization Intestinal colonization, escape of immune cells

86,87

actin filament polymerization activated-Rho ubiquitinmediated proteosomal degradation

181

182,183 184

LCGTs are responsible for severe diseases in man and animals, the most common of which are the C. difficile infections. Toxigenic C. difficile strains are the causative agent of pseudomembranous colitis and about 30% of the postantibiotic diarrhea, which are the most frequent nosocomial intestinal diseases.73 TcdA, which experimentally induces necrotic and hemorrhagic intestinal lesions was considered as the main virulence factor.2,68 However, in the hamster disease model, TcdB was found to be the essential virulence factor by using genetically modified C. difficile strains.74 Since C. difficile strains producing both TcdA and TcdB or only TcdB cause enteric disease in humans, TcdB might

www.landesbioscience.com

Reference

RhoA, B, C, E

TcsL

CNF1, CNF2, CNF3 CNFY

Effects

169,170

Clostridium limosum Bacillus cereus Staphylococcus aureus

Rac, Cdc42, R-Ras, Ral, Rap RhoG, Rac Cdc42 (variable) Ras, Rap, Ral TC10, TCL R-Ras1, 2, 3 RhoA, Rac, Cdc42 RhoA, Rac, Cdc42

Dermonecrotic toxin (DNT)

Biochemical activity

be also an important enterotoxin.2,68 Both TcdB and TcdA participate in the alteration of the intestinal barrier and in the recruitment of inflammatory cells, which are abundant in the lesions. C. sordellii and C. novyi are involved in gangrene. C. sordellii induces sporadic cases of fulminant toxic shock syndrome which accompanies local infections, mainly of uterus.75 It is also an agent of hemorrhagic enteritis and enterotoxemia in cattle.17 TcsL of C. sordellii, which only modifies Rac among the Rho proteins, also alters the permeability of epithelium, but through a slightly different way than that of C. difficile toxins. Indeed, TcsL mainly causes a redistribution of E-cadherin whereas tight

Small GTPases

5

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

junctions are not significantly affected.76 C. sordellii can cause necrotic and hemorrhagic enteritis, probably in a similar manner than C. difficile colitis, but this pathogen is also responsible for sporadic cases of bacteraemia and toxic shock without myonecrosis of skeletal muscle.77 The portal of entry is presumed to be the gastrointestinal tract. Local increased intestinal barrier permeability due to C. sordellii toxins probably facilitates the translocation of bacteria into the blood stream. C. sordellii is also associated with toxic shock syndrome (hypotension, hemoconcentration, pleural effusion, sero-sanguineous ascite) consecutive to local infection of uterus or perineum.78-82 Diffusion of TcsL in the blood circulation and alteration of vascular endothelial barrier are the main features of the pathogenesis. In experimental mice, TcsL causes a marked edema in the cardio-respiratory system by altering E-cadherin junctions between lung endothelial cells in a Rac modification-dependent manner and a rapid death.83 A novel enzymatically inactivation of RhoGTPases has been recently identified consisting in N-acetyl-glucosamination of a switch I tyrosine. Photorhabdus asymbiotica toxin (PaTox) (334 kDa) contains a C-terminal domain with a conserved glucosyltransferase motif (DxD) with an additional Asp residue (DxDD), and catalyzes mono-glucosylation of RhoA at Tyr-34 as well as Rac and Cdc42 at Tyr-32 using UDP-N-acetylglucosamine as co-substrate. PaTox induces actin filament disruption and cell rounding, inhibits macrophage phagocytosis, and is toxic for insect larvae.84 Modification of the conserved tyrosine in switch I impairs downstream signaling of RhoGTPases as well as activation by RhoGEFs. In addition, PaTox exhibits a second enzymatic activity related to that of RhoGTPase activating toxins (see below). Indeed, PaTox deamidates a crucial glutamine residue of switch II involved in GTP hydrolysis of heterotrimeric G proteins (Gaq/11 and Gai).84 The PaTox effects of heterotrimeric G protein activation remain to be elucidated. They might counterbalance the inhibitory activity of RhoGTPases in a spatiotemporal manner during the host infection. P. asymbiotica is an entomopathogen, and also an emerging pathogen for human causing localized soft tissue infection or disseminated bacteremic infection.85 Another type of toxins, which inactivate RhoGTPases, are produced by bacteria of the Vibrio species. Vibrio cholerae is a potent intestinal pathogen, which secretes in addition to the cholera toxin, multiple virulence factors which are involved in the early colonization step by preventing bacteria killing via the innate immune cells like neutrophils and macrophages. Among the accessory toxins, MARTX (multifunctional autoprocessing repeats-in-toxin) is a large type I-secreted protein (350– 560 kDa), which alters the actin filaments via two mechanisms mediated by two distinct domains. The actin cross-linking domain is responsible for covalent cross-linking of actin monomers, whereas the Rho-inactivation domain (RID) block Rho, Rac and Cdc42 in their GDP inactive form leading to actin filament depolymerization. MARTX do not directly modify RhoGTPases, but rather interact with regulatory proteins of RhoGTPases. RID shows homology with the cysteine proteases of the Clan CE family, but the exact mechanism of action remains to be identified.86,87 V. cholerae MARTX toxin carries

6

two effector domains which attack the actin cytoskeleton by two mechanisms, one of which involves a downregulation of RhoGTPases. RhoGTPase activating toxins Another class of bacterial toxins enzymatically modify RhoGTPases, but by blocking these molecules in their active form. This toxin family encompasses DNT (dermonecrotic factor) from Bordetella sp, CNF (cytotoxic necrotizing factor) from Escherichia coli, and Yersinia pseudotuberculosis (Table 1). CNF consists of several isoforms which are highly related at the amino acid level. CNF1-producing E. coli strains are pathogens for humans, in which they represent about 30% of uropathogenic strains, and are also involved in enteritis and septicaemia in animals.88-90 CNF2 strains are mainly isolated from calf and piglet, and CNF3 strains from sheep and goat.5,91 DNT is more distantly related to CNF and shares sequence identity mainly in the catalytic domain. CNF is the prototype of the RhoGTPase activating toxins and consists of a single chain protein (about 110 kDa) containing three functional domains: an N-terminal domain (amino acids 1–299), which is involved in the recognition of a cell surface receptor, a central domain (amino acids 299–720) containing two hydrophobic regions which have been proposed to translocate the toxin across the cell membrane, and a C-terminal (720– 1014) catalytic domain. The C-terminal domain of CNF1 has an original protein fold formed by a central b-sandwich, that is composed of two mixed b–sheets, surrounded by a-helices and extensive loop regions.92 CNF1 seems to be secreted by a Type I secretion system in the extracellular medium similarly to a-hemolysin.5 Then, the toxin recognizes laminin receptor (LPR) on the surface of target cells and it is endocytosed in a clathrin-independent pathway. The enzymatic domain is then released from acidic late endosomes to the cytosol.93 CNF1 catalyzes the deamidation of Gln63 in Rho and Gln61 in Rac and Cdc42 to glutamic acid. Gln63/Gln61 are located in the switch II region of the Rho protein, which has an important function in GTP hydrolysis. Thereby, CNF1 blocks the RhoGTPases in their active form linked to GTP and thereby stimulates actin filament polymerization. Indeed, in fibroblasts like Vero cells, CNF1 causes dense actin stress fibers and focal contact point formation, whereas in epithelial cells (Hep2) the formation of lamellipodia and filopodia predominate. In both cell types, CNF1 leads to cell spreading resulting from the increase in actin filament formation at the leading edge and anchorage of actomyosin filaments to focal contact points.94 However, RhoGTPase activation by CNF1 is only transient, since deamidation increases the sensitization of RhoGTPases to ubiquitylation and subsequent proteasome degradation.95 One of the primary CNFinduced disturbances involves modification of intestinal or urinary tract cell barriers. Activation of RhoGTPases results in changes in the dynamic of the actin cytoskeleton and leads subsequently to alteration of intercellular junction assembly. In addition the increased cell motility even within monolayers, and ubiquitin-mediated degradation of RhoGTPases facilitate the disorganization of intercellular junctions.5 Thereby, CNF1

Small GTPases

Volume 5 Issue 3

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

decreases trans epithelial resistance and enhances paracellular permeability from the basal to apical surface of polarized cell monolayers. In T84 cell monolayers, CNF1 induces profound alterations of tight junction components. Occludin, ZO-1, claudin-1, and JAMs are highly redistributed from the apical ring to the cytoplasm, leading to a complete disruption of the continuity of TJ integrity between neighboring cells.96 Disruption of the integrity of the epithelium barrier favors the dissemination of bacteria in the underlying tissues. Another aspect of CNF-dependent RhoGTPase activation concerns the stimulation of bacterial internalization into epithelial cells. Thereby, CNF1 triggers bladder cell invasion by uropathogenic E. coli, mainly through Rac activation and increased membrane ruffling as well as a crosstalk between Rac and b1integrin.95 Thus, bacteria can escape the innate immunity and inflammatory defense, and bacterial persistence into bladder cells might be responsible for recurrence of urinary tract infections.5

Bacterial Pathogens Manipulate RhoGTPase Signaling via Injected Virulence Factors to Enter Cells and Colonize Host Tissues Pathogenic bacteria can invade the host organism at different levels: multiplication in the natural cavities, colonization of the surface of the mucosa, invasion of the extracellular space of tissues, or cell invasion and intracellular life. Here, the term “invasive bacteria” refers to bacteria, which enter and develop inside cells. Invasive bacteria manipulate differently the RhoGTPase pathways to enter the target cells. Instead to secrete toxins into the extracellular medium, they directly inject virulence factors into cells. Thereby, Gram-negative invasive bacteria have developed sophisticated delivery systems allowing the internalization of multiple effectors into a same cell in a temporally and spatially coordinated manner. These delivery systems consist of multicomponent apparatus forming a needle-like structure, which traverses both bacterial envelopes and eukaryotic cell membranes. The effector proteins are delivered through the central pore into the target cell, and in most cases via chaperone proteins which regulate proper folding and translocation. According to the structure and mode of delivery, several types of secretion system have been described. The most commonly used system by invasive bacteria is the type III secretion system, and to a lower extent type IV and type VI. Type III secretion system, also called injectisome, shows similarity with bacterial flagella, whereas type IV secretion system which is also used to deliver DNA has probably evolved from the conjugation machinery. Type VI secretion system retains structural similarity with the T4 bacteriophage tail spike.97-99 The bacterial effectors delivered into the target cell play multiple functions in bacterial infection. A central role concerns the modulation of the actin cytoskeleton mediating the internalization of bacteria into non-phagocytic cells. Type III secretion system is a common mechanism used by various pathogens to inject virulence factors, which modify the RhoGTPase signaling cascade and thus trigger the bacterial invasion.100 The virulence factors involved in bacterial invasion do not enzymatically modify

www.landesbioscience.com

RhoGTPases, but mimic protein regulators which activate or inactivate RhoGTPases. One of the best-documented examples of RhoGTPase manipulation by invasive bacteria is that of Salmonella.98,101-105 Salmonella can bind to intestinal epithelial cells via fimbriae and then enter by a trigger mechanism that induces the formation of large membrane ruffles engulfing the bacteria. The subsequent rearrangements of the actin cytoskeleton and plasma membrane are reminiscent of lamellipodia and filopodia responses stimulated by various agonists such as growth factors, hormones, or activated oncogenes. It has been demonstrated that Rac and Cdc42 are involved in the Salmonella dependent cytoskeletal rearrangements. These effects are mediated by two virulence factors secreted by type III secretion system, SopE and SopE2. Like GEFs, SopE activates Rac1, Rac2, Cdc42, RhoG, and also but to a lesser extent RhoA by catalyzing the exchange of GDP for GTP,106 whereas SopE2, an isoform of SopE, interacts with Cdc42 but not with Rac1.101 Rac1 plays a central role in actin remodeling and membrane ruffling that mediate Salmonella uptake in non-phagocytic cells. Salmonella exploits the Rac signaling pathway mainly via IQGAP, a scaffold protein which stabilizes Rac and Cdc42 in their active form and promotes actin assembly by interacting with N-WASP and Arp2/3. IQGAP also activates the MAPK/ERK pathway, which contributes to efficient Salmonella uptake, resulting in synergistic effect on bacterial infection.107 Since Cdc42 activates Rac, a crosstalk between the two RhoGTPase coordinates the actin cytoskeleton rearrangement mediating Salmonella entry.103 SopE binds to the switch I and switch II regions of GDPloaded RhoGTPases like other RhoGEFs, ultimately leading to switch I and II rearrangements in a conformation of weak GDP affinity binding. This results in GDP release. Then, RhoGTPases interact with GTP, which is in higher concentration (200– 500 mM) in cells vs. GDP, leading to the conformational change in the active GTP-RhoGTPase structure.108,109 This mechanism is similar to that used by eukaryotic GEFs which belong to the Dbl family of proteins containing a Dbl homology domain (DH) and a plekstrin homology domain (PH), like the exchange factor Tiam1 (T-lymphoma invasion and metastasis)-1 protein.110 However, the catalytic domain of SopE has a different structure than that of Tiam1 and interacts with the switch regions via a GAGA motif to reorient switches I and II, whereas the catalytic core of other RhoGEFs consists of an a-helix containing a critical Lys at the active site.110 SopE shows neither amino acid sequence nor structure homology with Dbl proteins, but both types of proteins trigger a similar mechanism of nucleotide exchange. Bacterial and eukaryotic GEFs are examples of convergent evolution. Salmonella produces additional type III secretion factors which trigger redundant SopE effects on the actin cytoskeleton such as SopB, which is involved in the actin cytoskeleton rearrangement including membrane ruffling and bacterial uptake into cells. The effects of SopB are mediated by RhoG and Cdc42 activation. SopB is not a bacterial GEF, it is an inositol phosphate phosphatase which indirectly stimulates Cdc42 by activation of Vav2, a Cdc42-GEF, as well as RhoG via SGEF

Small GTPases

7

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

(SH3-containing guanine nucleotide exchange factor, an exchange factor for RhoG) activation, mediated by the inositol phosphate metabolism.103,111,112 In addition to their effect to facilitate bacterial uptake, SopE, SopE2, and SopB alter the tight junctions and increase the intestinal barrier permeability in a RhoGTPase activation-dependent manner.113 Moreover, one report shows that SopE also acts as a GEF for Rab5 and mediates the recruitment of Rab5 in its GTP form to phagosomes containing Salmonella. This promotes the fusion of these phagosomes with early endosomes, preventing their transport to lysozomes and subsequent destruction.114 Thereby, SopE facilitates intracellular survival of Salmonella. When internalized, Salmonella secretes two additional virulence factors (SifA and SifB) belonging to the WxxxE family of effectors described by Alto.115-117 The WxxxE family which includes IpgB1 and IpgB2 from Shigella flexneri, MAP and EspM2 from Escherichia coli, and EspT from Citrobacter rhodentium (Table 2) have been described as RhoGTPase mimics. Indeed, IpgB2 has been found to interact and activate RhoA effectors such as Rock and mDia, and thus promoting actin filament polymerization.115 However, IpgB2 and IpgB1 previously identified as RhoA and RhoG mimics, respectively,115,118 have been found to activate RhoGTPases through a GEF activity.118120 SifA and SifB contain the WxxxE motif, but also two three helix bundles forming a V-shapped structure which is highly similar to that of SopE, suggesting that they could be GEFs.116 However, no GEF activity of SifA toward RhoA has been evidenced.121 The precise mechanism of action of SifA and SifB leading to Rho or other RhoGTPase activation remains to be elucidated. SifA mediates the formation of filaments which maintain the integrity of vacuoles containing Salmoenlla and coordinates the equilibirum between kinesin and dynein in the control of membrane dynamics.117,122 Thereby, SifA by controling the membrane integritiy of vacuoles containing Salmonella, plays a role in the virulence of this pathogen. In addition, activation of Cdc42 and Rac by SopE leads to stimulation of p21-activated kinase (PAK) and subsequent activation of JNK, the MAP kinase pathways and a number of transcriptional factors, as well as caspase-1 activation resulting in secretion of proinflammatory cytokines.106,123 Indeed SopE promotes a host inflammatory response which changes the microbiota composition allowing intestinal colonization by the pathogen Salmonella.124 Therefore, SopE is involved in cellular invasion of Salmonella but also in tissue dissemination. After entry into cell, Salmonella restores the original cytoskeleton architecture by injecting another type of virulence factor (SptP) which inactivates Rac and Cdc42 and thus antagonizes the effects of SopE/E2 and SopB. SptP contains a N-terminal domain which mediates GAP activity toward Rac and Cdc42, and a C-terminal tyrosine phosphatase domain. Despite the absence of structural homology with eukaryotic GAPs, SptP retains a similar mechanism of stimulation of the GTP hydrolysis reaction via an Arg residue, called Arg finger, supporting a convergent evolution between bacterial and eukaryotic GAPs.125 Thereby, the remodeling of the local actin cytoskeleton mediated

8

by SopE/E2 and SopB allowing the uptake of Salmonella is transient, since this effect is reverted by SptP. A temporal coordination between the bacterial GEFs and GAP is required to induce an efficient bacterial internalization into cells. Indeed, SopE and SptP are equally delivered during the initial step of infection. But SopE is polyubiquitinated and rapidly degraded by the proteasome, whereas SptP is more stable allowing a longer activity and thus a restoration of the actin cytoskeleton.126 Enteropathogenic and enterohemorrhagic E. coli use a different strategy to coordinate RhoGTPase activity by injecting into cells an inhibitor of eukaryotic Rho-GEFs (EspH) and bacterial GEFs (EspT, EspM2, Map). EspH binds to the DH-PH domain of eukaryotic RhoGEFs thus preventing their interaction with RhoGTPases. Thereby, EspH inactivates RhoGTPases and induces focal adhesion disassembly, actin filament depolymerization, cell detachment and cytotoxicity similarly to TcdB. However, EspH is not active on bacterial GEFs, which are structurally different from eukaryotic RhoGEFs.127,128 These invasive pathogens trigger a subtle control of RhoGTPase activity by inhibiting eukaryotic RhoGEFs and by translocating bacterial GEFs insensitive to EspH, which modulates endogenous RhoGTPase for their own benefit. Manipulation of the RhoGTPase signaling by bacterial virulence factors which are injected trough type III secretion system and which mimic eukaryotic Rho-GEFs is used by various enteroinvasive bacteria such as Salmonella, Shigella, enteropathogenic and enterohemorrhagic E. coli, and also other invasive bacteria like Burkholderia pseudomallei (Table 2). Activation of Rac and Cdc42 and subsequent formation of membrane ruffles are key features induced by the bacterial GEFs allowing invasion into non-phagocytic cells and intracellular survival. Indeed, Shigella IpgB1, E. coli EspT, and B. pseudomallei BopE are GEFs having an equivalent role than that of SopE in the local actin remodeling and subsequent bacterial uptake and internalization into vacuoles (Table 2). Manipulation of RhoGTPases by bacterial GEFs support additional functions in the invasion process including stimulation of host immune response via activation of the MAPK and/ or NF-kB signaling pathways, and disorganization of the intercellular junctions leading to increased epithelial barrier permeability (reviewed in108). Chlamydia trachomatis, an obligate intracellular pathogen, also uses a balanced control of the actin cytoskeleton through manipulation of RhoGTPase signaling to invade host cells. The elementary bodies, which are the extracellular infectious forms of C. trachomatis, attach to cells and induce the formation of an actinrich pedestal at the site of entry, which promotes the internalization into endocytic vesicles. The remodeling of the local actin cytoskeleton is Rac-dependent and is mediated by a type III secretion system factor called, TARP. The role of TARP is not yet fully understood. Phosphorylated TARP interacts with GEFs (Vav2, and Sos1/Eps8/Abi1), which activate Rac. TARP exhibits no GEF activity but is a scaffolding protein, which also recruits the PI3-kinase, thus leading to the activation of endogenous Rac GEFs and subsequently of WAVE2, Abi1, and Arp2/3 complex.129-131 Actin filament polymerization is counterbalanced by an additional factor, CT166, secreted inside cells. CT166 contains in its N-terminal part the conserved enzymatic DxD motif

Small GTPases

Volume 5 Issue 3

Table 2. Examples of bacterial effectors injected by type III or IV secretion system that modulate RhoGTPase signaling by mimicking regulators and mediate bacterial entry into target cells Bacterial effectors

SopE, SopE2

Pathogen

Salmonella typhimurium

SopB

Shigella flexneri

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

EspT

EspG1, EspG2

Actin filament polymerization, membrane ruffling, bacterial cell invasion Inositol phosphatase Actin cytoskeleton VAV2 (Cdc42 GEF) activation rearrangement, membrane SGEF (RhoG GEF) activation ruffling GEF? Membrane integrity of Salmonella-containing vacuoles GEF Formation of actin-rich membrane ruffles GEF Formation of actin stress fibers GEF Formation of actin stress fibers

Rac, Cdc42 RhoA, Rac, Cdc42

Enteropathogenic, Enterohemorrhagic Escherichia coli Citrobacter rodentium Enteropathogenic Escherichia coli Citrobacter rodentium Enteropathogenic Escherichia coli

RhoA

Formation of actin-rich membrane ruffles Actin rich pedestal at the site of entry

192

Chlamydia trachomatis

Rac, Cdc42?

Activation of Rac GEF (Vav2, Sos1/Abi1/Eps8)

SptP

Salmonella typhimurium

CT166

Bacterial effectors inhibiting Rho-GTPases Rac, Cdc42 GAP

Yersinia pseudotuberculosis Rac, RhoG, RhoA, Cdc42 Yersinia enterocolitica Yersinia pestis Pseudomonas aeruginosa Rho, Rac, Cdc42

Yersinia pseudotuberculosis Yersinia enterocolitica Yersinia pestis Chlamydia trachomatis

GAP

GAP

Reversion of SopE-induced actin filament polymerization, bacterial internalization Actin filament depolymerization Ant-iphagoctytosis Actin filament depolymerization, antiphagocytosis

115,191

129,130

125

193-195

149,196

RhoGEFs

Inhibition of eukaryotic RhoGEFs

Actin filament depolymerization

128

Rho, Rac, Cdc42

GAP

147

Rho, Rac

GDI

Rac

Rac glucosylation

Actin filament depolymerization Actin filament depolymerization Anti-phagocytosis Actin filament depolymerization Restoration of the actin architecture

of LCGTs and inactivates Rac1 but not RhoA thus leading to actin filament depolymerization.132 Thereby, Chlamydia controls the host actin cytoskeleton to mediate its internalization by activating Rac through TARP, which stimulates eukaryotic Rac GEFs and by reversing the reorganization of actin filaments by inactivating directly Rac via glucosylation.

www.landesbioscience.com

121

189,190

Tarp

YopO/YpkA

119,120

Microtubule depolymerization Disruption of tight junctions Formation of actin-rich filopodia

GEF

AexT

118,186

Arf RhoA

Rac, Cdc42

Enteropathogenic, enterohemorrhagic Escherichia coli Aeromonas salmonicida

116,117

187,188

BopE

EspH

103,111,112

Formation of actin-rich membrane ruffles

Cdc42

ExoS, ExoT

101,185

GEF

Enteropathogenic, Enterohemorrhagic Escherichia coli Burkholderia pseudomallei

YopE

Reference

Rac, Cdc42

Arf GTPase inhibiton PAK activation GEF-H1 GEF

Map

Main effects

Bacterial effectors activating Rho-GTPases Rac, CDC42 GEF

RhoA

IpgB2 EspM1, EspM2, EspM3

Mode of action

Cdc42, RhoG

SifA, SifB

IpgB1

Target

152

132

Manipulation of RhoGTPase Signaling by Bacterial Adhesins In addition to specific factors interacting with RhoGTPase signaling, numerous bacterial pathogens use membrane proteins or adhesins to trigger signaling pathways which facilitate bacterial

Small GTPases

9

Table 3. Examples of bacterial adhesins and membrane proteins which trigger signaling pathways including activation of RhoGTPase and mediate pathogen internalization into cell Membrane protein CadF, Flagella (FlaA, FlaB)

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

Fibronectin binding protein-A

Pathogen

Activated RhoGTPase

Signaling pathway

Effects

Reference

Campylobacter jejuni

Cdc42 Rac

integrin b1 receptor, FAK and Src, PI3-kinase, VAV2 integrin b1 receptor, FAK, DOCK180/TIAM-1 Integrin signaling

Actin cytoskeleton rearrangement, membrane ruffling,

157

Actin cytoskeleton rearrangement, bacterial invasion Actin cytoskeleton rearrangement, bacterial invasion Bacterial uptake into epithelial cells Bacterial uptake Formation of lamellipodia and filopodia Bacterial uptake

136

Staphylococcus aureus

Rho, Rac, Cdc42

Pneumococcal surface protein C (PspC)

Streptococcus pneumoniae

Cdc42

plgR, PI3-kinase, Akt

Invasin

Yersinia pseudotuberculosis

Rac1

Brucela abortus Bartonella bacillifomis

Rho, Rac, Cdc42 Rho, Rac, Cdc42

Integrin signaling N-WASP, Arp2/3 ? PAK MAPK, SAPK/JNK

? ?

entry into cell and which directly or indirectly involve RhoGTPase activation. Bacterial pathogens have adopted complex strategies to interact with the host. Some bacterial factors induce unique effect on cells, but many other manipulate complex networks of signaling pathways. Examples are shown in Table 3 to illustrate the activation of RhoGTPase signaling by surface bacterial components involved in bacterial entry into cells. This is the case of CadF, a Campylobacter jejuni outer membrane protein, which binds to fibronectin, a major extracellular matrix component. CadF as well as C. jejuni flagella, which are constituted of two major proteins, FlaA and FlaB, mediate membrane ruffling, filopodia formation, and engulfment of bacteria in a RhoGTPase, mainly Cdc42 and Rac, activation-dependent manner. The proposed signaling pathway leading to Cdc42 activation and bacterial invasion includes clustering and activation of b1-integrin receptor, phosphorylation of FAK and Src, activation of EGF and PDGF receptors, PI3-kinase, Vav2 and Cdc42 activation.133,134 FAK can also stimulates DOCK180/TIAM-1, which are GEFs activating Rac1.135 Numerous other invasive bacterial pathogens use the RhoGTPase pathway to induce local actin cytoskeleton rearrangement and their subsequent entry into cells. Indeed, RhoGTPase inhibition significantly prevents the bacterial uptake. However, the molecules, membrane proteins or translocated effectors, which activate the RhoGTPase signaling pathway as well as the mechanism of RhoGTPase activation remain to be defined for many pathogens. For example, the S. aureus adhesin called fibronectin binding protein-A induces b1-integrin clustering and subsequent signaling leading to reorganization of the actin cytoskeleton and staphylococci invasion in a Src tyrosine kinase and RhoGTPase activationdependent manner including N-WASP and Arp3/3 complex (Table 3).136 The pneumococcal surface protein C (PspC), the major adhesin of Streptococcus pneumoniae binds to polymeric immunoglobulin receptor (pIgR) and promotes bacterial invasion via Cdc42, PI3 kinase and Akt activation.137 Activation of b1integrin is also achieved by invasin, an outer membrane protein of

10

135

137

138

139 140

Y. pseudotuberculosis, leading to Rac activation and actin cytoskeleton remodeling via N-WASP and Arp2/3 complex.138 Brucella abortus and Bartonella bacilliformis are intracellular pathogens which also enter cells in a RhoGTPase dependent-manner (Table 3), but the molecules that trigger this cell signaling have not yet been identified.139,140 However, the actin cytoskeleton remodeling by these microorganisms is triggered upon contact between bacteria and target cells suggesting that recognition and activation of a cell surface receptor is probably a required early event in the stimulation of the RhoGTPase cascade.

Bacterial Pathogens Manipulate RhoGTPase Signaling via Injected Virulence Factors to Avoid Phagocytosis Several pathogens replicate in phagocytic cells, such as macrophages, and thus invade the organism or proliferate in local necrotic tissues avoiding phagocytosis. Bacteria do not require a specific equipment to enter phagocytic cells, but in contrast they have an absolute necessity to escape or survive in hostile cells. For this purpose, they inject into phagocytic cells an array of virulence factors which block the phagocytosis or prevent their killing in phagocytic cells. Numerous virulence factors which induce an anti-phagocytic effect, act by modulating the RhoGTPase signaling. RhoGTPases are key players in phagocytosis,37,141 and down modulation of RhoGTPases results in an efficient strategy to impair the phagocytosis. Bacterial virulence factors modulate the RhoGTPase signaling cascade either by mimicking a GAP/ GDI activity or by inducing a biochemical modification which blocks the RhoGTPases in their inactive form (Table 4). Mimicking host regulatory RhoGTPases YopE from Yersinia is secreted by type III secretion system and exerts a GAP activity toward Rac1, RhoG, RhoA and Cdc42 resulting in actin filament depolymerization and inhibition of

Small GTPases

Volume 5 Issue 3

Table 4. Examples of bacterial effectors injected by type III or IV secretion system that modulate RhoGTPase signaling by biochemical modification and mediate bacterial entry into target cells or anti-phagocytosis Bacterial effector

TccC5

YopT

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

LopT VopS

Pathogen

Target

Mode of action

Bacterial effectors activating Rho-GTPases Photorhabdus luminescens Rho, Rac, Cdc42 ADP-ribosylation at Gln61/Gln63

Yersinia enterocolitica

Bacterial effectors inhibiting Rho-GTPases Rho, Rac, Cdc42 Proteolysis at Cys 200 (RhoA)

Photorhabdus luminescens Rho, Rac Vibrio parahemolyticus Rho, Rac, Cdc42

Proteolysis at Cys200 (RhoA) AMPylation at Thr35/Thr37

IbpA

Histophilus somni

Rho, Rac, Cdc42

AMPylation at Tyr32/Tyr34

PfhB2

Pasteurella multocida

Rho, Rac, Cdc42

AMPylation at Tyr32/Tyr34

phagocytosis. YopE shares sequence and structure similarities with the GAP N-terminal domain of SptP and Pseudomonas ExoS (see below), but not with eukaryotic GAPs. However, YopE retains en Arg finger motif like eukaryotic GAPs, which is critical for the activity.142 YopE contains a N-terminal domain (amino acids 54–75), termed membrane localization domain, which mediates the binding to membrane close to its host targets. The subcellular localization of YopE at the membrane seems to drive its specificity toward target RhoGTPases.143 Indeed, Rac is more rapidly inactivated than RhoA in Yersinia-infected cells.144 A key central player in Yersinia invasion has been addressed to RhoG, which is an upstream regulator of RhoGTPases. Indeed, active RhoG stimulates the GEF activity of the complex Elmo. Dock180 toward Rac1. Yersinia adherence to cell promotes RhoG activation through the bacterial surface protein, Invasin, and clustering of b1-integrin, thus facilitating cell invasion. Then, RhoG and subsequently Rac1 are downregulated by YopE GAP activity.145 YopE is one of the most efficient virulence factors produced by Yersinia pestis, Yersinia pseudotuberculosis and Yersinia enterocolitica which are involved in the impairment of phagocytosis by macrophages and neutrophils allowing to these pathogens to invade the host tissues.146 YopE which inhibits immune defenses, is a major determinant of Yersinia virulence. Other pathogens, like Pseudomonas and Aeromonas, exploit the RhoGTPase signaling by injecting effectors, which mimic eukaryotic RhoGAPs to escape phagocytic cells (Table 2). Interestingly, Pseudomonas ExoS and ExoT as well as Aeromonas AexT have a double activity contributing to their antiphagocytic effects. These effectors contain a N-terminal GAP domain and a C-terminal ADP-ribosylating domain. ExoS ADP-ribosylates numerous substrates including Ras and Rab proteins, whereas ExoT modifies CRK proteins and AexT targets actin monomers.147-150 ADPribosylating activity of ExoS toward Rab5 plays an essential role in phagocytosis inhibition.151 ADPribosylation of CRK proteins prevents their interaction with focal adhesion proteins and/ or with DOCK180 which is a GEF for Rac, thus inhibiting Racdependent phagocytosis.150

www.landesbioscience.com

Effects

Reference

Aggregation of actin filaments Anti-phagocytosis

157

Actin filament depolymerization Anti-phagocytosis Anti-phagocytosis Actin filament depolymerization Intestinal invasion Actin filament depolymerization Bacterial dissemination, escape of immune defense Actin filament depolymerization

197

156 158

160,161

162

Instead to mimic GAP function, YopO from Y. enterocolitica and YpkA from Y. pseudotubercuosis and Yersinia pestis contain a C-terminal domain, which interacts with RhoA and Rac to mimic RhoGDI thereby sequestering the RhoGTPases in their inactive form. Indeed, The YpkA C-terminal domain interacts with Rac adopting a structural conformation similar to that of RhoGDI and blocks nucleotide exchange and subsequent activation.152 The N-terminal domain of YpkA/YopO is a kinase domain which is activated by monomeric actin and which phosphorylates Gaq.153 YpkA induces actin filament disruption and inhibits Rac-dependent Fcg receptor- but not complement receptor-3-dependent phagocytosis through its RhoGDI-like activity and has a critical role in Yersinia virulence.144,152,154 Biochemical modification of RhoGTPases Several invasive pathogens use virulence factors which enzymatically modify RhoGTPases and block them in their inactive form (Table 4). Thereby, Yersinia produce a cysteine protease, YopT, which is an additional virulence factor involved in the inhibition of phagocytosis. YopT is structurally related to the papain-like cysteine proteases and removes specifically the C-terminal cysteine of RhoGTPases to which the geranyl-geranyl group is attached. The cleavage of the isoprenoid moiety results in the RhoGTPase release from the membrane and from GDI leading to their inactivation. YopT recognizes Rho, Rac and Cdc42 in their isoprenylated form but independently of their GDP- or GTP- bound state. However, RhoA seems to be the preferred substrate in living cells. YopT disrupts the actin cytoskeleton and prevents the formation of the phagocytic cups thus inhibiting phagocytosis of opsonized and non-opsonized Yersinia by macrophages and neutrophils.144,155 However, Yersiniae use a combination of Yops to interact with the immune cells and the role of YopT alone during infection remains to be defined.144 Other pathogens also use a cysteine protease specific of RhoGTPases as virulence factor. Thereby, the entomopathogenic bacterium, Photorhabdus luminescens, delivers the type III secretion effector, LopT, which induces proteolytic removing of

Small GTPases

11

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

RhoA and Rac from eukaryotic cell membrane similarly to YopT. LopT plays an essential role in Photorhabdus virulence by preventing phagocytosis from insect macrophages.156 P. luminescens produces additional virulence factors which interfere with the actin cytoskeleton like TccC5 and TccC3. TccC5 ADP-ribosylates RhoA at Gln63 as well as Rac and Cdc42 at Gln61, thus leading to the loss of the GTPase activity and subsequent RhoGTPase activation similarly to the deamidation of Gln61/63 by CNF.157 TccC3 is also an ADP-ribosyltransferase, but which modifies actin monomers at Thr148 preventing the interaction with thymosin b4, a blocker of actin-actin association. The combined effects of TccC5 and TccC3 results in increased actin polymerization but leading to a disorganization of the actin cytoskeleton and impairment of phagocytosis.157 Indeed, Tcc5 and TccC3 induce aggregation of actin filaments which are not assembled in long and continuous stress fibers but form patches of short filaments. These opposite effects on actin polymerization support that an appropriate balance between actin monomers and filaments is required for a properly organized and functional actin cytoskeleton. P. luminescens uses a set of toxins which target the actin cytoskeleton but which exploit different modes of action including activation and inactivation of RhoGTPases to fully prevent the phagocytosis. An additional biochemical mechanism of RhoGTPase inhibition consists of modification of a residue of switch I by adenylation (or AMPylation), which prevents interaction with downstream effectors by steric hindrance. Indeed, VopS, a type III secretion system effector produced by V. parahaemolyticus, catalyzes the adenylation of Rho, Rac, and Cdc42 at Thr35/37. Vops and kinases use the same cosubstrate which is ATP. However, VopS transfers the AMP part to an acceptor amino acid, whereas kinases use the g-phosphate to modify the target residue.158 VopS induces a disruption of actin filament in HeLa cells by inactivating RhoGTPases and contributes with additional effectors like VopQ, a PI3K-independent autophagy inducer, to the suppression of the NLRC4-dependent inflammasome response. Indeed, VopS via inactivation of Cdc42 inhibits NLRC4-dependent caspase-1 activation.158,159 Escape of intestinal NLRC4 expressing macrophages confers an advantage to the invading V. parahaemolyticus. VopS contains the adenylation core motif HPExxGNGR, which is called Fic (filamentous induced by c-AMP) domain. This domain is conserved in other bacterial virulence factors and in one human protein.160 IbpA (immunoglobulin-binding protein A) is secreted by a two-partner system by Histophilus somni which is responsible for pneumonia and septicemia in cattle. IbpA secreted at the bacterial surface is then internalized into cells by a yet undetermined pathway.161 IbpA catalyzes adenylation of Rho, Rac and Cdc42, but at a distinct residue of switch I than VopS, Tyr32/34 instead of Thr35/37.162 IbpA is critical in H. somnus virulence and induces HeLa cell rounding and actin cytoskeleton disruption. It is proposed that cell retraction leads to an increased paracellular permeability of pulmonary alveolar barrier allowing the transmigration of bacteria through the alveolar epithelial cell monolayers and subsequent bacterial dissemination in the underlying tissues and septicemia.161 The exact role of IbpA in pathogenesis is not

12

fully understood but seems important in bacterial dissemination and bacterial escape from the immune defenses.160 It is noteworthy that the adenylating factors including VopS and IbpA recognize equally both GTP- and GDP-bound RhoGTPase forms, in contrast to the other bacterial virulence factors which preferentially interact with RhoGTPases in their GDP state.162 A functionally related virulence factor to IbpA is synthesized by Pasteurella, another Gram-negative bacterium. Indeed, PfhB2 (Pasteurella filamentous hemagglutinin B2), which is produced by Pasteurella multocida, modifies RhoGTPases similarly to IbpA.162 Although IbpA and PfhB2 are not significantly related at the amino acid sequence level, they share similar structure. The role of PfhB in P. multocida virulence is not yet well defined. P. multocida genome contains two genes encoding PfhB1 and PfhB2, which show a high homology with Bordetella pertussis filamentous hemagglutinin (FhaB), and which retain a C-terminal domain similar to that of the serum resistance protein p76 of H. somnus. FhaB mediates bacterial adherence to host cells, and p76 induces bacterial resistance to opsonization.163 PfhB1 and PfhB2 could have similar functions in P. multocida and PfhB2 possibly triggers additional effects linked to its ability in modifying RhoGTPases.

Activation of RhoGTPases and Innate Immunity Host can detect the presence of pathogens and can develop an appropriate defense response trough the innate immunity. RhoGTPases are involved in various signaling pathways controlling innate immune functions such as Toll-like receptor (TLR) signaling, leukocyte chemotaxis and migration, phagocytosis, and NADPH oxidase activation.164 Among the various innate immunity molecules, NODs (nucleotide-binding and oligomerization domain) are intracellular receptors able to detect invasive bacteria by sensing cytosolic microbial products. NOD1 is expressed by many cell types including intestinal epithelial cells whereas NOD2 is restricted to monocytes, macrophages, dendritic cells and Paneth cells.165 Peptidoglycan fragments in the cytosol can activate NOD1 or NOD2 which forms a complex with receptor-interacting protein (RIP2) and other proteins called nodosome. This later activates the NF-kB and mitogenactivated protein (MAP) kinase pathways leading to proinflammatory and antimicrobial responses. Recently, RhoGTPases have been evidenced as components of NOD1 nodosomes and to control their activity. Indeed, activation of Rac and Cdc42 by bacterial virulence factors like SopE triggers the NOD1 signaling pathways and induces an inflammatory response.166,167 Similarly, CNF1 which activates Rho, Rac and Cdc42, triggers host immune response in a Rac2 activation-dependent manner and subsequent Rip (receptor-interacting protein kinase) pathway.168 Inversely, inactivation of RhoGTPases by toxins or virulence factors is probably detected by the NOD signaling pathway. Thereby, changes in RhoGTPase activation state by bacterial virulence factors or toxins are sensed by NOD1, which elicits an adaptive response toward pathogens, thus conferring to host cell an efficient mechanism monitoring the presence of invasive

Small GTPases

Volume 5 Issue 3

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

Figure 4. Main effects of manipulation of RhoGTPase signaling by toxigenic and invasive bacteria on pathogen dissemination. Toxigenic bacteria secrete toxins, which upon biochemical RhoGTPase modification alter the actin cytoskeleton and subsequently epithelial and endothelial barriers opening a breach for bacterial dissemination. Invasive bacteria modulate the RhoGTPase signaling mainly by injecting virulence factors which mimic regulatory proteins of the RhoGTPase cycle or by triggering a cell surface receptor mediated pathway which activates RhoGTpases leading to bacterial invasion in epithelial cells or to prevention of phagocytosis by professional phagocytic cells. Activation of RhoGTPase state is sensed by the NOD1 pathway of innate immunity which controls a proinflammatory defense response.

bacteria or pathogens able to attack an epithelial barrier through manipulation of RhoGTPases.

Concluding Remarks Modulation of the actin cytoskeleton via manipulation of RhoGTPase signaling is a common mechanism used by numerous pathogens to invade and disseminate in host organism. It is noteworthy that regarding the complex pathway of regulation of actin polymerization, bacterial pathogens mainly target RhoGTPases and to a lesser extent actin monomers. Thereby, RhoGTPases are the preferred targets of toxins and virulence factors which modulate the actin cytoskeleton. Bacterial pathogens have developed numerous strategies to interfere with RhoGTPases including direct interaction and subsequent biochemical modification, or mimicry of RhoGTPases or proteins involved in their regulation. Targeting RhoGTPases instead of actin molecules offer two main advantages. First, interacting on an upstream step of the regulation cascade of actin polymerization affords an amplified effect. Secondly, targeting distinct sets of RhoGTPases allows a focused activity on only certain pools of actin like localized cortical actin thus leading to specific modification of actin structure such as formation of phagocytic cups. Based on the mode of secretion of virulence factors, two main mechanism of pathogenicity can be distinguished: secretion of toxins or injection of virulence factors directly into target cells. Toxins are

www.landesbioscience.com

exported and diffuse in the external bacterial medium to all the surrounding cells. Toxins enter target cells using eukaryotic endocytosis machinery and induce biochemical modification of RhoGTPases via their enzymatic activity. This often results in a brutal and global effect of toxins on the actin cytoskeleton leading to severe lesions like destruction of epithelial and endothelial barriers, ultimately tissue necrosis and hemorrhage. In contrast, the virulence factors are selectively injected into the target cells via specialized apparatus (type III, IV, or VI secretion systems) and preferentially control the RhoGTPase signaling cycle by regulatory protein mimicry avoiding cell destruction but hijacking cell function for the own pathogen benefit (bacterial internalization, anti-phagocytosis, escape of immune defenses). Invasive bacteria can also manipulate the RhoGTPase signaling through recognition and activation of cell surface receptor(s) (Fig. 4). The distinct modes of interactions of bacterial effectors with RhoGTPases or eukaryotic protein mimicry by bacteria likely represent different modes of adaptation of pathogens with their respective host. However, the mechanisms of evolution resulting from the co-habitation between bacteria and eukaryotes remain enigmatic. If RhoGTPases are preferred targets for many invasive and toxigenic bacteria, these molecules also constitute sentinels sensed by the host cell to monitor the presence of pathogens. Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

Small GTPases

13

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

References 1. Aktories K. Bacterial protein toxins that modify host regulatory GTPases. Nat Rev Microbiol 2011; 9:48798; PMID:21677684; http://dx.doi.org/10.1038/ nrmicro2592 2. Genth H, Dreger SC, Huelsenbeck J, Just I. Clostridium difficile toxins: more than mere inhibitors of Rho proteins. Int J Biochem Cell Biol 2008; 40:592-7; PMID:18289919; http://dx.doi.org/10.1016/j. biocel.2007.12.014 3. Aktories K, Schwan C, Papatheodorou P, Lang AE. Bidirectional attack on the actin cytoskeleton. Bacterial protein toxins causing polymerization or depolymerization of actin. Toxicon 2012; 60:572-81; PMID:22543189; http://dx.doi.org/10.1016/j. toxicon.2012.04.338 4. Lemichez E, Aktories K. Hijacking of Rho GTPases during bacterial infection. Exp Cell Res 2013; 319:2329-36; PMID:23648569; http://dx.doi.org/ 10.1016/j.yexcr.2013.04.021 5. Lemonnier M, Landraud L, Lemichez E. Rho GTPase-activating bacterial toxins: from bacterial virulence regulation to eukaryotic cell biology. FEMS Microbiol Rev 2007; 31:515-34; PMID:17680807; http://dx.doi.org/10.1111/j.1574-6976.2007.00078.x 6. Boquet P, Lemichez E. Bacterial virulence factors targeting Rho GTPases: parasitism or symbiosis? Trends Cell Biol 2003; 13:238-46; PMID:12742167; http:// dx.doi.org/10.1016/S0962-8924(03)00037-0 7. Fiorentini C, Falzano L, Travaglione S, Fabbri A. Hijacking Rho GTPases by protein toxins and apoptosis: molecular strategies of pathogenic bacteria. Cell Death Differ 2003; 10:147-52; PMID: 12700642; http://dx.doi.org/10.1038/sj.cdd.4401151 8. Gruenheid S, Finlay BB. Microbial pathogenesis and cytoskeletal function. Nature 2003; 422:775-81; PMID:12700772; http://dx.doi.org/10.1038/ nature01603 9. Lemichez E, Lecuit M, Nassif X, Bourdoulous S. Breaking the wall: targeting of the endothelium by pathogenic bacteria. Nat Rev Microbiol 2010; 8:93104; PMID:20040916 10. Popoff MR, Geny B. Multifaceted role of Rho, Rac, Cdc42 and Ras in intercellular junctions, lessons from toxins. Biochim Biophys Acta 2009; 1788:797-812; PMID:19366594; http://dx.doi.org/10.1016/j. bbamem.2009.01.011 11. Henderson B, Poole S, Wilson M. Bacterial modulins: a novel class of virulence factors which cause host tissue pathology by inducing cytokine synthesis. Microbiol Rev 1996; 60:316-41; PMID:8801436 12. Barth H, Aktories K, Popoff MR, Stiles BG. Binary bacterial toxins: biochemistry, biology, and applications of common Clostridium and Bacillus proteins. Microbiol Mol Biol Rev 2004; 68:373-402; PMID:15353562; http://dx.doi.org/10.1128/ MMBR.68.3.373-402.2004 13. Stiles BG, Wigelsworth DJ, Popoff MR, Barth H. Clostridial binary toxins: iota and C2 family portraits. Front Cell Infect Microbiol 2011; 1:11; PMID: 22919577; http://dx.doi.org/10.3389/fcimb.2011. 00011 14. Aktories K, Just I. Clostridial Rho-inhibiting protein toxins. Curr Top Microbiol Immunol 2005; 291:113-45; PMID:15981462; http://dx.doi.org/ 10.1007/3-540-27511-8_7 15. Vogelsgesang M, Pautsch A, Aktories K. C3 exoenzymes, novel insights into structure and action of Rho-ADPribosylating toxins. Naunyn Schmiedebergs Arch Pharmacol 2007; 374:347-60; PMID:17146673; http://dx. doi.org/10.1007/s00210-006-0113-y 16. Hall A. Rho GTPases and the actin cytoskeleton. Science 1998; 279:509-14; PMID:9438836; http://dx. doi.org/10.1126/science.279.5350.509 17. Popoff MR, Bouvet P. Clostridial toxins. Future Microbiol 2009; 4:1021-64; PMID:19824793; http://dx.doi.org/10.2217/fmb.09.72

14

18. Wilde C, Chhatwal GS, Schmalzing G, Aktories K, Just I. A novel C3-like ADP-ribosyltransferase from Staphylococcus aureus modifying RhoE and Rnd3. J Biol Chem 2001; 276:9537-42; PMID:11124969; http://dx.doi.org/10.1074/jbc.M011035200 19. Han S, Arvai AS, Clancy SB, Tainer JA. Crystal structure and novel recognition motif of rho ADP-ribosylating C3 exoenzyme from Clostridium botulinum: structural insights for recognition specificity and catalysis. J Mol Biol 2001; 305:95-107; PMID:11114250; http://dx.doi.org/10.1006/jmbi.2000.4292 20. Vogelsgesang M, Stieglitz B, Herrmann C, Pautsch A, Aktories K. Crystal structure of the Clostridium limosum C3 exoenzyme. FEBS Lett 2008; 582:1032-6; PMID:18325337; http://dx.doi.org/10.1016/j. febslet.2008.02.051 21. Tsuge H, Nagahama M, Nishimura H, Hisatsune J, Sakaguchi Y, Itogawa Y, Katunuma N, Sakurai J. Crystal structure and site-directed mutagenesis of enzymatic components from Clostridium perfringens iota-toxin. J Mol Biol 2003; 325:471-83; PMID:12498797; http://dx.doi.org/10.1016/S00222836(02)01247-0 22. Evans HR, Holloway DE, Sutton JM, Ayriss J, Shone CC, Acharya KR. C3 exoenzyme from Clostridium botulinum: structure of a tetragonal crystal form and a reassessment of NAD-induced flexure. Acta Crystallogr D Biol Crystallogr 2004; 60:1502-5; PMID:15272191; http://dx.doi.org/10.1107/ S0907444904011680 23. Evans HR, Sutton JM, Holloway DE, Ayriss J, Shone CC, Acharya KR. The crystal structure of C3stau2 from Staphylococcus aureus and its complex with NAD. J Biol Chem 2003; 278:45924-30; PMID:12933793; http://dx.doi.org/10.1074/jbc. M307719200 24. Wei Y, Zhang Y, Derewenda U, Liu X, Minor W, Nakamoto RK, Somlyo AV, Somlyo AP, Derewenda ZS. Crystal structure of RhoA-GDP and its functional implications. Nat Struct Biol 1997; 4:699-703; PMID:9302995; http://dx.doi.org/10.1038/nsb0997699 25. Bourmeyster N, Stasia MJ, Garin J, Gagnon J, Boquet P, Vignais PV. Copurification of rho protein and the rho-GDP dissociation inhibitor from bovine neutrophil cytosol. Effect of phosphoinositides on rho ADPribosylation by the C3 exoenzyme of Clostridium botulinum. Biochemistry 1992; 31:12863-9; PMID: 1334435; http://dx.doi.org/10.1021/bi00166a022 26. Ren XD, Bokoch GM, Traynor-Kaplan A, Jenkins GH, Anderson RA, Schwartz MA. Physical association of the small GTPase Rho with a 68-kDa phosphatidylinositol 4-phosphate 5-kinase in Swiss 3T3 cells. Mol Biol Cell 1996; 7:435-42; PMID:8868471; http://dx.doi.org/10.1091/mbc.7.3.435 27. Sehr P, Joseph G, Genth H, Just I, Pick E, Aktories K. Glucosylation and ADP ribosylation of rho proteins: effects on nucleotide binding, GTPase activity, and effector coupling. Biochemistry 1998; 37:5296-304; PMID:9548761; http://dx.doi.org/10.1021/ bi972592c 28. Genth H, Schmidt M, Gerhard R, Aktories K, Just I. Activation of phospholipase D1 by ADP-ribosylated RhoA. Biochem Biophys Res Commun 2003; 302:127-32; PMID:12593858; http://dx.doi.org/ 10.1016/S0006-291X(03)00112-8 29. Barth H, Olenik C, Sehr P, Schmidt G, Aktories K, Meyer DK. Neosynthesis and activation of Rho by Escherichia coli cytotoxic necrotizing factor (CNF1) reverse cytopathic effects of ADP-ribosylated Rho. J Biol Chem 1999; 274:27407-14; PMID:10488072; http://dx.doi.org/10.1074/jbc.274.39.27407 30. Genth H, Gerhard R, Maeda A, Amano M, Kaibuchi K, Aktories K, Just I. Entrapment of Rho ADP-ribosylated by Clostridium botulinum C3 exoenzyme in the Rho-guanine nucleotide dissociation inhibitor-1 complex. J Biol Chem 2003; 278:28523-7;

Small GTPases

31.

32.

33.

34.

35.

36.

37.

38.

39.

40.

41.

42.

43.

44.

PMID:12750364; http://dx.doi.org/10.1074/jbc. M301915200 Fujihara H, Walker LA, Gong MC, Lemichez E, Boquet P, Somlyo AV, Somlyo AP. Inhibition of RhoA translocation and calcium sensitization by in vivo ADP-ribosylation with the chimeric toxin DC3B. Mol Biol Cell 1997; 8:2437-47; PMID:9398666; http://dx.doi.org/10.1091/ mbc.8.12.2437 Chardin P, Boquet P, Madaule P, Popoff MR, Rubin EJ, Gill DM. The mammalian G protein rhoC is ADP-ribosylated by Clostridium botulinum exoenzyme C3 and affects actin microfilaments in Vero cells. EMBO J 1989; 8:1087-92; PMID:2501082 Aktories K, Barth H, Just I. Clostridium botulinum C3 exoenzyme and C3-like transferases. In: Aktories K, Just I, eds. Bacterial protein toxins. Berlin: Springer, 2000:207-33. Aktories K, Wilde C, Vogelsgesang M. Rho-modifying C3-like ADP-ribosyltransferases. Rev Physiol Biochem Pharmacol 2004; 152:1-22; PMID:15372308; http://dx.doi.org/10.1007/s10254-004-0034-4 Nusrat A, Giry M, Turner JR, Colgan SP, Parkos CA, Carnes D, Lemichez E, Boquet P, Madara JL. Rho protein regulates tight junctions and perijunctional actin organization in polarized epithelia. Proc Natl Acad Sci U S A 1995; 92:10629-33; PMID:7479854; http://dx.doi.org/10.1073/pnas.92.23.10629 Fahrer J, Kuban J, Heine K, Rupps G, Kaiser E, Felder E, Benz R, Barth H. Selective and specific internalization of clostridial C3 ADP-ribosyltransferases into macrophages and monocytes. Cell Microbiol 2010; 12:233-47; PMID:19840027; http://dx.doi. org/10.1111/j.1462-5822.2009.01393.x Caron E, Hall A. Identification of two distinct mechanisms of phagocytosis controlled by different Rho GTPases. Science 1998; 282:1717-21; PMID: 9831565; http://dx.doi.org/10.1126/science.282. 5394.1717 Holbourn KP, Sutton JM, Evans HR, Shone CC, Acharya KR. Molecular recognition of an ADP-ribosylating Clostridium botulinum C3 exoenzyme by RalA GTPase. Proc Natl Acad Sci U S A 2005; 102:5357-62; PMID:15809419; http://dx.doi.org/ 10.1073/pnas.0501525102 Pautsch A, Vogelsgesang M, Tr€ankle J, Herrmann C, Aktories K. Crystal structure of the C3bot-RalA complex reveals a novel type of action of a bacterial exoenzyme. EMBO J 2005; 24:3670-80; PMID:16177825; http://dx.doi.org/10.1038/sj.emboj.7600813 Czech A, Yamaguchi T, Bader L, Linder S, Kaminski K, Sugai M, Aepfelbacher M. Prevalence of Rho-inactivating epidermal cell differentiation inhibitor toxins in clinical Staphylococcus aureus isolates. J Infect Dis 2001; 184:785-8; PMID:11517442; http://dx.doi. org/10.1086/322983 O’Neill AJ, Larsen AR, Skov R, Henriksen AS, Chopra I. Characterization of the epidemic European fusidic acid-resistant impetigo clone of Staphylococcus aureus. J Clin Microbiol 2007; 45:1505-10; PMID:17344365; http://dx.doi.org/10.1128/ JCM.01984-06 Messad N, Landraud L, Canivet B, Lina G, Richard JL, Sotto A, Lavigne JP, Lemichez E; French Study Group on the Diabetic Foot. Distribution of edin in Staphylococcus aureus isolated from diabetic foot ulcers. Clin Microbiol Infect 2013; 19:875-80; PMID:23176291; http://dx.doi.org/10.1111/14690691.12084 Munro P, Clement R, Lavigne JP, Pulcini C, Lemichez E, Landraud L. High prevalence of edin-C encoding RhoA-targeting toxin in clinical isolates of Staphylococcus aureus. Eur J Clin Microbiol Infect Dis 2011; 30:965-72; PMID:21311940; http://dx.doi. org/10.1007/s10096-011-1181-6 Molinari G, Rohde M, Wilde C, Just I, Aktories K, Chhatwal GS. Localization of the C3-Like ADP-

Volume 5 Issue 3

45.

46.

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

47.

48.

49.

50.

51.

52.

53.

54.

55.

56.

57.

ribosyltransferase from Staphylococcus aureus during bacterial invasion of mammalian cells. Infect Immun 2006; 74:3673-7; PMID:16714601; http://dx.doi. org/10.1128/IAI.02013-05 Munro P, Benchetrit M, Nahori MA, Stefani C, Clement R, Michiels JF, Landraud L, Dussurget O, Lemichez E. The Staphylococcus aureus epidermal cell differentiation inhibitor toxin promotes formation of infection foci in a mouse model of bacteremia. Infect Immun 2010; 78:3404-11; PMID:20479081; http:// dx.doi.org/10.1128/IAI.00319-10 Boyer L, Doye A, Rolando M, Flatau G, Munro P, Gounon P, Clement R, Pulcini C, Popoff MR, Mettouchi A, et al. Induction of transient macroapertures in endothelial cells through RhoA inhibition by Staphylococcus aureus factors. J Cell Biol 2006; 173:809-19; PMID:16754962; http://dx.doi.org/10.1083/ jcb.200509009 Maddugoda MP, Stefani C, Gonzalez-Rodriguez D, Saarikangas J, Torrino S, Janel S, Munro P, Doye A, Prodon F, Aurrand-Lions M, et al. cAMP signaling by anthrax edema toxin induces transendothelial cell tunnels, which are resealed by MIM via Arp2/3-driven actin polymerization. Cell Host Microbe 2011; 10:464-74; PMID:22100162; http://dx.doi.org/ 10.1016/j.chom.2011.09.014 Jank T, Aktories K. Structure and mode of action of clostridial glucosylating toxins: the ABCD model. Trends Microbiol 2008; 16:222-9; PMID:18394902; http://dx.doi.org/10.1016/j.tim.2008.01.011 Just I, Hofmann F, Aktories K. Molecular mechanism of action of the large clostridial cytotoxins. In: Aktories K, Just I, eds. Bacterial Protein Toxins. Berlin: Springer, 2000:307-31. Egerer M, Giesemann T, Herrmann C, Aktories K. Autocatalytic processing of Clostridium difficile toxin B. Binding of inositol hexakisphosphate. J Biol Chem 2009; 284:3389-95; PMID:19047051; http://dx.doi. org/10.1074/jbc.M806002200 Egerer M, Giesemann T, Jank T, Satchell KJ, Aktories K. Auto-catalytic cleavage of Clostridium difficile toxins A and B depends on cysteine protease activity. J Biol Chem 2007; 282:25314-21; PMID:17591770; http://dx.doi.org/10.1074/jbc.M703062200 Reineke J, Tenzer S, Rupnik M, Koschinski A, Hasselmayer O, Schrattenholz A, Schild H, von EichelStreiber C. Autocatalytic cleavage of Clostridium difficile toxin B. Nature 2007; 446:415-9; PMID: 17334356; http://dx.doi.org/10.1038/nature05622 Popoff MR, Geny B. Rho/Ras-GTPase-dependent and -independent activity of clostridial glucosylating toxins. J Med Microbiol 2011; 60:1057-69; PMID: 21349986; http://dx.doi.org/10.1099/jmm.0. 029314-0 Just I, Selzer J, Wilm M, von Eichel-Streiber C, Mann M, Aktories K. Glucosylation of Rho proteins by Clostridium difficile toxin B. Nature 1995; 375:500-3; PMID:7777059; http://dx.doi.org/10.1038/ 375500a0 Just I, Wilm M, Selzer J, Rex G, von Eichel-Streiber C, Mann M, Aktories K. The enterotoxin from Clostridium difficile (ToxA) monoglucosylates the Rho proteins. J Biol Chem 1995; 270:13932-6; PMID: 7775453; http://dx.doi.org/10.1074/jbc.270.23. 13932 Popoff MR, Chaves-Olarte E, Lemichez E, von Eichel-Streiber C, Thelestam M, Chardin P, Cussac D, Antonny B, Chavrier P, Flatau G, et al. Ras, Rap, and Rac small GTP-binding proteins are targets for Clostridium sordellii lethal toxin glucosylation. J Biol Chem 1996; 271:10217-24; PMID:8626586; http:// dx.doi.org/10.1074/jbc.271.17.10217 Herrmann C, Ahmadian MR, Hofmann F, Just I. Functional consequences of monoglucosylation of Ha-Ras at effector domain amino acid threonine 35. J Biol Chem 1998; 273:16134-9; PMID:9632667; http://dx.doi.org/10.1074/jbc.273.26.16134

www.landesbioscience.com

58. Vetter IR, Hofmann F, Wohlgemuth S, Herrmann C, Just I. Structural consequences of mono-glucosylation of Ha-Ras by Clostridium sordellii lethal toxin. J Mol Biol 2000; 301:1091-5; PMID:10966807; http://dx. doi.org/10.1006/jmbi.2000.4045 59. Genth H, Aktories K, Just I. Monoglucosylation of RhoA at threonine 37 blocks cytosol-membrane cycling. J Biol Chem 1999; 274:29050-6; PMID: 10506156; http://dx.doi.org/10.1074/jbc.274.41. 29050 60. Halabi-Cabezon I, Huelsenbeck J, May M, Ladwein M, Rottner K, Just I, Genth H. Prevention of the cytopathic effect induced by Clostridium difficile Toxin B by active Rac1. FEBS Lett 2008; 582:37516; PMID:18848548; http://dx.doi.org/10.1016/j. febslet.2008.10.003 61. Chen ML, Pothoulakis C, LaMont JT. Protein kinase C signaling regulates ZO-1 translocation and increased paracellular flux of T84 colonocytes exposed to Clostridium difficile toxin A. J Biol Chem 2002; 277:4247-54; PMID:11729192; http://dx.doi.org/ 10.1074/jbc.M109254200 62. Nusrat A, von Eichel-Streiber C, Turner JR, Verkade P, Madara JL, Parkos CA. Clostridium difficile toxins disrupt epithelial barrier function by altering membrane microdomain localization of tight junction proteins. Infect Immun 2001; 69:1329-36; PMID: 11179295; http://dx.doi.org/10.1128/IAI.69.3.13291336.2001 63. Hecht G, Koutsouris A, Pothoulakis C, LaMont JT, Madara JL. Clostridium difficile toxin B disrupts the barrier function of T84 monolayers. Gastroenterology 1992; 102:416-23; PMID:1732112 64. Hecht G, Pothoulakis C, LaMont JT, Madara JL. Clostridium difficile toxin A perturbs cytoskeletal structure and tight junction permeability of cultured human intestinal epithelial monolayers. J Clin Invest 1988; 82:1516-24; PMID:3141478; http://dx.doi. org/10.1172/JCI113760 65. Pothoulakis C, Lamont JT. Microbes and microbial toxins: paradigms for microbial-mucosal interactions II. The integrated response of the intestine to Clostridium difficile toxins. Am J Physiol Gastrointest Liver Physiol 2001; 280:G178-83; PMID:11208538 66. Savidge TC, Pan WH, Newman P, O’brien M, Anton PM, Pothoulakis C. Clostridium difficile toxin B is an inflammatory enterotoxin in human intestine. Gastroenterology 2003; 125:413-20; PMID:12891543; http://dx.doi.org/10.1016/S0016-5085(03)00902-8 67. Sun X, Savidge T, Feng H. The enterotoxicity of Clostridium difficile toxins. Toxins (Basel) 2010; 2:184880; PMID:22069662; http://dx.doi.org/10.3390/ toxins2071848 68. Voth DE, Ballard JD. Clostridium difficile toxins: mechanism of action and role in disease. Clin Microbiol Rev 2005; 18:247-63; PMID:15831824; http:// dx.doi.org/10.1128/CMR.18.2.247-263.2005 69. Warny M, Keates AC, Keates S, Castagliuolo I, Zacks JK, Aboudola S, Qamar A, Pothoulakis C, LaMont JT, Kelly CP. p38 MAP kinase activation by Clostridium difficile toxin A mediates monocyte necrosis, IL-8 production, and enteritis. J Clin Invest 2000; 105:1147-56; PMID:10772660; http://dx.doi.org/ 10.1172/JCI7545 70. Na X, Zhao D, Koon HW, Kim H, Husmark J, Moyer MP, Pothoulakis C, LaMont JT. Clostridium difficile toxin B activates the EGF receptor and the ERK/MAP kinase pathway in human colonocytes. Gastroenterology 2005; 128:1002-11; PMID: 15825081; http://dx.doi.org/10.1053/j.gastro.2005. 01.053 71. Wershil B, Castagliuolo I, Pothoulakis C. Direct evidence of mast cell involvement in Clostridium difficile toxin A-induced enteritis in mice. Gastroenterology 1998; 114:956-64; PMID:9558284; http://dx.doi. org/10.1016/S0016-5085(98)70315-4 72. Stanley JD, Bartlett JG, Dart BW 4th, Ashcraft JH. Clostridium difficile infection. Curr Probl Surg 2013;

Small GTPases

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

83.

84.

85.

86.

50:302-37; PMID:23764494; http://dx.doi.org/ 10.1067/j.cpsurg.2013.02.004 Bassetti M, Villa G, Pecori D, Arzese A, Wilcox M. Epidemiology, diagnosis and treatment of Clostridium difficile infection. Expert Rev Anti Infect Ther 2012; 10:1405-23; PMID:23253319; http://dx.doi.org/ 10.1586/eri.12.135 Lyras D, O’Connor JR, Howarth PM, Sambol SP, Carter GP, Phumoonna T, Poon R, Adams V, Vedantam G, Johnson S, et al. Toxin B is essential for virulence of Clostridium difficile. Nature 2009; 458:11769; PMID:19252482; http://dx.doi.org/10.1038/ nature07822 Aronoff DM, Ballard JD. Clostridium sordellii toxic shock syndrome. Lancet Infect Dis 2009; 9:725-6; PMID:19926032; http://dx.doi.org/10.1016/S14733099(09)70303-2 Boehm C, Gibert M, Geny B, Popoff MR, Rodriguez P. Modification of epithelial cell barrier permeability and intercellular junctions by Clostridium sordellii lethal toxins. Cell Microbiol 2006; 8:1070-85; PMID:16819961; http://dx.doi.org/10.1111/j.14625822.2006.00687.x Abdulla A, Yee L. The clinical spectrum of Clostridium sordellii bacteraemia: two case reports and a review of the literature. J Clin Pathol 2000; 53:70912; PMID:11041062; http://dx.doi.org/10.1136/ jcp.53.9.709 Rørbye C, Petersen IS, Nilas L. Postpartum Clostridium sordellii infection associated with fatal toxic shock syndrome. Acta Obstet Gynecol Scand 2000; 79:1134-5; PMID:11130102 Sinave C, Le Templier G, Blouin D, Leveille F, Deland E. Toxic shock syndrome due to Clostridium sordellii: a dramatic postpartum and postabortion disease. Clin Infect Dis 2002; 35:1441-3; PMID:12439811; http://dx.doi.org/10.1086/344464 Ho CS, Bhatnagar J, Cohen AL, Hacker JK, Zane SB, Reagan S, Fischer M, Shieh WJ, Guarner J, Ahmad S, et al. Undiagnosed cases of fatal Clostridium-associated toxic shock in Californian women of childbearing age. Am J Obstet Gynecol 2009; 201:e1-7; PMID:19628200; http://dx.doi.org/10.1016/j. ajog.2009.05.023 Soper DE. Clostridial myonecrosis arising from an episiotomy. Obstet Gynecol 1986; 68(Suppl):26S-8S; PMID:3737071 McGregor JA, Soper DE, Lovell G, Todd JK. Maternal deaths associated with Clostridium sordellii infection. Am J Obstet Gynecol 1989; 161:987-95; PMID:2801850; http://dx.doi.org/10.1016/00029378(89)90768-0 Geny B, Khun H, Fitting C, Zarantonelli L, Mazuet C, Cayet N, Szatanik M, Prevost MC, Cavaillon JM, Huerre M, et al. Clostridium sordellii lethal toxin kills mice by inducing a major increase in lung vascular permeability. Am J Pathol 2007; 170:1003-17; PMID:17322384; http://dx.doi.org/10.2353/ ajpath.2007.060583 Jank T, Bogdanovic X, Wirth C, Haaf E, Spoerner M, B€ohmer KE, Steinemann M, Orth JH, Kalbitzer HR, Warscheid B, et al. A bacterial toxin catalyzing tyrosine glycosylation of Rho and deamidation of Gq and Gi proteins. Nat Struct Mol Biol 2013; 20:1273-80; PMID:24141704; http://dx.doi.org/10.1038/ nsmb.2688 Weissfeld AS, Halliday RJ, Simmons DE, Trevino EA, Vance PH, O’Hara CM, Sowers EG, Kern R, Koy RD, Hodde K, et al. Photorhabdus asymbiotica, a pathogen emerging on two continents that proves that there is no substitute for a well-trained clinical microbiologist. J Clin Microbiol 2005; 43:4152-5; PMID:16081963; http://dx.doi.org/10.1128/ JCM.43.8.4152-4155.2005 Ahrens S, Geissler B, Satchell KJ. Identification of a His-Asp-Cys catalytic triad essential for function of the Rho inactivation domain (RID) of Vibrio cholerae MARTX toxin. J Biol Chem 2013; 288:1397-408;

15

87.

88.

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

89.

90.

91.

92.

93.

94.

95.

96.

97.

98.

99.

100.

101.

16

PMID:23184949; http://dx.doi.org/10.1074/jbc. M112.396309 Sheahan KL, Satchell KJ. Inactivation of small Rho GTPases by the multifunctional RTX toxin from Vibrio cholerae. Cell Microbiol 2007; 9:1324-35; PMID:17474905; http://dx.doi.org/10.1111/j.14625822.2006.00876.x Bingen-Bidois M, Clermont O, Bonacorsi S, Terki M, Brahimi N, Loukil C, Barraud D, Bingen E. Phylogenetic analysis and prevalence of urosepsis strains of Escherichia coli bearing pathogenicity island-like domains. Infect Immun 2002; 70:3216-26; PMID:12011017; http://dx.doi.org/10.1128/ IAI.70.6.3216-3226.2002 Blanco J, Alonso MP, Gonzalez EA, Blanco M, Garabal JI. Virulence factors of bacteraemic Escherichia coli with particular reference to production of cytotoxic necrotising factor (CNF) by P-fimbriate strains. J Med Microbiol 1990; 31:175-83; PMID:1968978; http://dx.doi.org/10.1099/00222615-31-3-175 Landraud L, Gauthier M, Fosse T, Boquet P. Frequency of Escherichia coli strains producing the cytotoxic necrotizing factor (CNF1) in nosocomial urinary tract infections. Lett Appl Microbiol 2000; 30:213-6; PMID:10747253; http://dx.doi.org/ 10.1046/j.1472-765x.2000.00698.x Munro P, Lemichez E. Bacterial toxins activating Rho GTPases. Curr Top Microbiol Immunol 2005; 291:177-90; PMID:15981464; http://dx.doi.org/ 10.1007/3-540-27511-8_10 Buetow L, Flatau G, Chiu K, Boquet P, Ghosh P. Structure of the Rho-activating domain of Escherichia coli cytotoxic necrotizing factor 1. Nat Struct Biol 2001; 8:584-8; PMID:11427886; http://dx.doi.org/ 10.1038/89610 Chung JW, Hong SJ, Kim KJ, Goti D, Stins MF, Shin S, Dawson VL, Dawson TM, Kim KS. 37-kDa laminin receptor precursor modulates cytotoxic necrotizing factor 1-mediated RhoA activation and bacterial uptake. J Biol Chem 2003; 278:16857-62; PMID:12615923; http://dx.doi.org/10.1074/jbc. M301028200 Boquet P, Fiorentini C. The cytotoxic necrotizing factor 1 from Escherichia coli. In: Aktories K, Just I, eds. Bacterial protein toxins. Berlin: Springer, 2000:36184. Doye A, Mettouchi A, Bossis G, Clement R, BuissonTouati C, Flatau G, Gagnoux L, Piechaczyk M, Boquet P, Lemichez E. CNF1 exploits the ubiquitinproteasome machinery to restrict Rho GTPase activation for bacterial host cell invasion. Cell 2002; 111:553-64; PMID:12437928; http://dx.doi.org/ 10.1016/S0092-8674(02)01132-7 Hopkins AM, Walsh SV, Verkade P, Boquet P, Nusrat A. Constitutive activation of Rho proteins by CNF-1 influences tight junction structure and epithelial barrier function. J Cell Sci 2003; 116:725-42; PMID:12538773; http://dx.doi.org/10.1242/ jcs.00300 Chatterjee S, Chaudhury S, McShan AC, Kaur K, De Guzman RN. Structure and biophysics of type III secretion in bacteria. Biochemistry 2013; 52:2508-17; PMID:23521714; http://dx.doi.org/10.1021/ bi400160a Hicks SW, Galan JE. Exploitation of eukaryotic subcellular targeting mechanisms by bacterial effectors. Nat Rev Microbiol 2013; 11:316-26; PMID: 23588250; http://dx.doi.org/10.1038/nrmicro3009 Voth DE, Broederdorf LJ, Graham JG. Bacterial Type IV secretion systems: versatile virulence machines. Future Microbiol 2012; 7:241-57; PMID:22324993; http://dx.doi.org/10.2217/fmb.11.150 Cossart P, Sansonetti PJ. Bacterial invasion: the paradigms of enteroinvasive pathogens. Science 2004; 304:242-8; PMID:15073367; http://dx.doi.org/ 10.1126/science.1090124 Friebel A, Ilchmann H, Aepfelbacher M, Ehrbar K, Machleidt W, Hardt WD. SopE and SopE2 from

102.

103.

104.

105.

106.

107.

108.

109.

110.

111.

112.

113.

114.

115.

Salmonella typhimurium activate different sets of RhoGTPases of the host cell. J Biol Chem 2001; 276:34035-40; PMID:11440999; http://dx.doi.org/ 10.1074/jbc.M100609200 Patel JC, Galan JE. Manipulation of the host actin cytoskeleton by Salmonella–all in the name of entry. Curr Opin Microbiol 2005; 8:10-5; PMID: 15694851; http://dx.doi.org/10.1016/j.mib.2004. 09.001 Patel JC, Galan JE. Differential activation and function of Rho GTPases during Salmonella-host cell interactions. J Cell Biol 2006; 175:453-63; PMID: 17074883; http://dx.doi.org/10.1083/jcb.200605144 Valdez Y, Ferreira RB, Finlay BB. Molecular mechanisms of Salmonella virulence and host resistance. Curr Top Microbiol Immunol 2009; 337:93-127; PMID:19812981; http://dx.doi.org/10.1007/978-3642-01846-6_4 van der Heijden J, Finlay BB. Type III effector-mediated processes in Salmonella infection. Future Microbiol 2012; 7:685-703; PMID:22702524; http://dx. doi.org/10.2217/fmb.12.49 Hardt WD, Chen LM, Schuebel KE, Bustelo XR, Galan JE. S. typhimurium encodes an activator of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell 1998; 93:815-26; PMID:9630225; http://dx.doi.org/10.1016/S00928674(00)81442-7 Kim H, White CD, Li Z, Sacks DB. Salmonella enterica serotype Typhimurium usurps the scaffold protein IQGAP1 to manipulate Rac1 and MAPK signalling. Biochem J 2011; 440:309-18; PMID:21851337; http://dx.doi.org/10.1042/BJ20110419 Bulgin R, Raymond B, Garnett JA, Frankel G, Crepin VF, Berger CN, Arbeloa A. Bacterial guanine nucleotide exchange factors SopE-like and WxxxE effectors. Infect Immun 2010; 78:1417-25; PMID:20123714; http://dx.doi.org/10.1128/IAI.01250-09 Erickson JW, Cerione RA. Structural elements, mechanism, and evolutionary convergence of Rho proteinguanine nucleotide exchange factor complexes. Biochemistry 2004; 43:837-42; PMID:14744125; http:// dx.doi.org/10.1021/bi036026v Buchwald G, Friebel A, Galan JE, Hardt WD, Wittinghofer A, Scheffzek K. Structural basis for the reversible activation of a Rho protein by the bacterial toxin SopE. EMBO J 2002; 21:3286-95; PMID:12093730; http://dx.doi.org/10.1093/emboj/ cdf329 Norris FA, Wilson MP, Wallis TS, Galyov EE, Majerus PW. SopB, a protein required for virulence of Salmonella dublin, is an inositol phosphate phosphatase. Proc Natl Acad Sci U S A 1998; 95:14057-9; PMID:9826652; http://dx.doi.org/10.1073/ pnas.95.24.14057 Zhou D, Chen LM, Hernandez L, Shears SB, Galan JE. A Salmonella inositol polyphosphatase acts in conjunction with other bacterial effectors to promote host cell actin cytoskeleton rearrangements and bacterial internalization. Mol Microbiol 2001; 39:248-59; PMID:11136447; http://dx.doi.org/10.1046/j.13652958.2001.02230.x Boyle EC, Brown NF, Finlay BB. Salmonella enterica serovar Typhimurium effectors SopB, SopE, SopE2 and SipA disrupt tight junction structure and function. Cell Microbiol 2006; 8:1946-57; PMID:16869830; http://dx.doi.org/10.1111/j.14625822.2006.00762.x Mukherjee K, Parashuraman S, Raje M, Mukhopadhyay A. SopE acts as an Rab5-specific nucleotide exchange factor and recruits non-prenylated Rab5 on Salmonella-containing phagosomes to promote fusion with early endosomes. J Biol Chem 2001; 276:2360715; PMID:11316807; http://dx.doi.org/10.1074/jbc. M101034200 Alto NM, Shao F, Lazar CS, Brost RL, Chua G, Mattoo S, McMahon SA, Ghosh P, Hughes TR, Boone C, et al. Identification of a bacterial type III effector

Small GTPases

116.

117.

118.

119.

120.

121.

122.

123.

124.

125.

126.

127.

128.

family with G protein mimicry functions. Cell 2006; 124:133-45; PMID:16413487; http://dx.doi.org/ 10.1016/j.cell.2005.10.031 Ohlson MB, Huang Z, Alto NM, Blanc MP, Dixon JE, Chai J, Miller SI. Structure and function of Salmonella SifA indicate that its interactions with SKIP, SseJ, and RhoA family GTPases induce endosomal tubulation. Cell Host Microbe 2008; 4:434-46; PMID:18996344; http://dx.doi.org/10.1016/j. chom.2008.08.012 Orchard RC, Alto NM. Mimicking GEFs: a common theme for bacterial pathogens. Cell Microbiol 2012; 14:10-8; PMID:21951829; http://dx.doi.org/ 10.1111/j.1462-5822.2011.01703.x Handa Y, Suzuki M, Ohya K, Iwai H, Ishijima N, Koleske AJ, Fukui Y, Sasakawa C. Shigella IpgB1 promotes bacterial entry through the ELMO-Dock180 machinery. Nat Cell Biol 2007; 9:121-8; PMID:17173036; http://dx.doi.org/10.1038/ ncb1526 Hachani A, Biskri L, Rossi G, Marty A, Menard R, Sansonetti P, Parsot C, Van Nhieu GT, Bernardini ML, Allaoui A. IpgB1 and IpgB2, two homologous effectors secreted via the Mxi-Spa type III secretion apparatus, cooperate to mediate polarized cell invasion and inflammatory potential of Shigella flexenri. Microbes Infect 2008; 10:260-8; PMID:18316224; http://dx.doi.org/10.1016/j.micinf.2007.11.011 Klink BU, Barden S, Heidler TV, Borchers C, Ladwein M, Stradal TE, Rottner K, Heinz DW. Structure of Shigella IpgB2 in complex with human RhoA: implications for the mechanism of bacterial guanine nucleotide exchange factor mimicry. J Biol Chem 2010; 285:17197-208; PMID:20363740; http://dx. doi.org/10.1074/jbc.M110.107953 Arbeloa A, Garnett J, Lillington J, Bulgin RR, Berger CN, Lea SM, Matthews S, Frankel G. EspM2 is a RhoA guanine nucleotide exchange factor. Cell Microbiol 2010; 12:654-64; PMID:20039879; http://dx.doi.org/10.1111/j.1462-5822.2009.01423.x Leone P, Meresse S. Kinesin regulation by Salmonella. Virulence 2011; 2:63-6; PMID:21217202; http://dx. doi.org/10.4161/viru.2.1.14603 M€uller AJ, Hoffmann C, Galle M, Van Den Broeke A, Heikenwalder M, Falter L, Misselwitz B, Kremer M, Beyaert R, Hardt WD. The S. Typhimurium effector SopE induces caspase-1 activation in stromal cells to initiate gut inflammation. Cell Host Microbe 2009; 6:125-36; PMID:19683679; http://dx.doi.org/ 10.1016/j.chom.2009.07.007 Stecher B, Robbiani R, Walker AW, Westendorf AM, Barthel M, Kremer M, Chaffron S, Macpherson AJ, Buer J, Parkhill J, et al. Salmonella enterica serovar typhimurium exploits inflammation to compete with the intestinal microbiota. PLoS Biol 2007; 5:2177-89; PMID:17760501; http://dx.doi.org/10.1371/journal. pbio.0050244 Stebbins CE, Galan JE. Modulation of host signaling by a bacterial mimic: structure of the Salmonella effector SptP bound to Rac1. Mol Cell 2000; 6:1449-60; PMID:11163217; http://dx.doi.org/10.1016/S10972765(00)00141-6 Kubori T, Galan JE. Temporal regulation of salmonella virulence effector function by proteasomedependent protein degradation. Cell 2003; 115:33342; PMID:14636560; http://dx.doi.org/10.1016/ S0092-8674(03)00849-3 Dong N, Liu L, Shao F. A bacterial effector targets host DH-PH domain RhoGEFs and antagonizes macrophage phagocytosis. EMBO J 2010; 29:1363-76; PMID:20300064; http://dx.doi.org/10.1038/ emboj.2010.33 Wong AR, Clements A, Raymond B, Crepin VF, Frankel G. The interplay between the Escherichia coli Rho guanine nucleotide exchange factor effectors and the mammalian RhoGEF inhibitor EspH. MBio 2012; 3:e00250-11; PMID:22251971; http://dx.doi. org/10.1128/mBio.00250-11

Volume 5 Issue 3

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

129. Clifton DR, Fields KA, Grieshaber SS, Dooley CA, Fischer ER, Mead DJ, Carabeo RA, Hackstadt T. A chlamydial type III translocated protein is tyrosinephosphorylated at the site of entry and associated with recruitment of actin. Proc Natl Acad Sci U S A 2004; 101:10166-71; PMID:15199184; http://dx.doi.org/ 10.1073/pnas.0402829101 130. Lane BJ, Mutchler C, Al Khodor S, Grieshaber SS, Carabeo RA. Chlamydial entry involves TARP binding of guanine nucleotide exchange factors. PLoS Pathog 2008; 4:e1000014; PMID:18383626; http:// dx.doi.org/10.1371/journal.ppat.1000014 131. Subtil A, Wyplosz B, Bala~ na ME, Dautry-Varsat A. Analysis of Chlamydia caviae entry sites and involvement of Cdc42 and Rac activity. J Cell Sci 2004; 117:3923-33; PMID:15265988; http://dx.doi.org/ 10.1242/jcs.01247 132. Thalmann J, Janik K, May M, Sommer K, Ebeling J, Hofmann F, Genth H, Klos A. Actin re-organization induced by Chlamydia trachomatis serovar D–evidence for a critical role of the effector protein CT166 targeting Rac. PLoS One 2010; 5:e9887; PMID:20360858; http://dx.doi.org/10.1371/journal.pone.0009887 133. Krause-Gruszczynska M, Boehm M, Rohde M, Tegtmeyer N, Takahashi S, Buday L, Oyarzabal OA, Backert S. The signaling pathway of Campylobacter jejuniinduced Cdc42 activation: Role of fibronectin, integrin beta1, tyrosine kinases and guanine exchange factor Vav2. Cell Commun Signal 2011; 9:32; PMID:22204307; http://dx.doi.org/10.1186/1478811X-9-32 134. Krause-Gruszczynska M, Rohde M, Hartig R, Genth H, Schmidt G, Keo T, K€onig W, Miller WG, Konkel ME, Backert S. Role of the small Rho GTPases Rac1 and Cdc42 in host cell invasion of Campylobacter jejuni. Cell Microbiol 2007; 9:2431-44; PMID:17521326; http://dx.doi.org/10.1111/j.14625822.2007.00971.x 135. Boehm M, Krause-Gruszczynska M, Rohde M, Tegtmeyer N, Takahashi S, Oyarzabal OA, Backert S. Major host factors involved in epithelial cell invasion of Campylobacter jejuni: role of fibronectin, integrin beta1, FAK, Tiam-1, and DOCK180 in activating Rho GTPase Rac1. Front Cell Infect Microbiol 2011; 1:17; PMID:22919583; http://dx.doi.org/10.3389/ fcimb.2011.00017 136. Schr€ oder A, Schr€ oder B, Roppenser B, Linder S, Sinha B, F€assler R, Aepfelbacher M. Staphylococcus aureus fibronectin binding protein-A induces motile attachment sites and complex actin remodeling in living endothelial cells. Mol Biol Cell 2006; 17:5198-210; PMID:17021255; http://dx.doi.org/10.1091/mbc. E06-05-0463 137. Agarwal V, Hammerschmidt S. Cdc42 and the phosphatidylinositol 3-kinase-Akt pathway are essential for PspC-mediated internalization of pneumococci by respiratory epithelial cells. J Biol Chem 2009; 284:19427-36; PMID:19473971; http://dx.doi.org/ 10.1074/jbc.M109.003442 138. Alrutz MA, Srivastava A, Wong KW, D’SouzaSchorey C, Tang M, Ch’Ng LE, Snapper SB, Isberg RR. Efficient uptake of Yersinia pseudotuberculosis via integrin receptors involves a Rac1-Arp 2/3 pathway that bypasses N-WASP function. Mol Microbiol 2001; 42:689-703; PMID:11722735; http://dx.doi. org/10.1046/j.1365-2958.2001.02676.x 139. Guzman-Verri C, Chaves-Olarte E, von EichelStreiber C, L opez-Go~ ni I, Thelestam M, Arvidson S, Gorvel JP, Moreno E. GTPases of the Rho subfamily are required for Brucella abortus internalization in nonprofessional phagocytes: direct activation of Cdc42. J Biol Chem 2001; 276:44435-43; PMID:11579087; http://dx.doi.org/10.1074/jbc. M105606200 140. Verma A, Ihler GM. Activation of Rac, Cdc42 and other downstream signalling molecules by Bartonella bacilliformis during entry into human endothelial cells. Cell Microbiol 2002; 4:557-69;

www.landesbioscience.com

141.

142.

143.

144.

145.

146.

147.

148.

149.

150.

151.

152.

153.

154.

155.

156.

PMID:12390349; http://dx.doi.org/10.1046/j.14625822.2002.00217.x Niedergang F, Chavrier P. Regulation of phagocytosis by Rho GTPases. Curr Top Microbiol Immunol 2005; 291:43-60; PMID:15981459; http://dx.doi. org/10.1007/3-540-27511-8_4 Evdokimov AG, Tropea JE, Routzahn KM, Waugh DS. Crystal structure of the Yersinia pestis GTPase activator YopE. Protein Sci 2002; 11:401-8; PMID:11790850; http://dx.doi.org/10.1110/ ps.34102 Aepfelbacher M, Roppenser B, Hentschke M, Ruckdeschel K. Activity modulation of the bacterial Rho GAP YopE: an inspiration for the investigation of mammalian Rho GAPs. Eur J Cell Biol 2011; 90:951-4; PMID:21255863; http://dx.doi.org/ 10.1016/j.ejcb.2010.12.004 Aepfelbacher M, Trasak C, Ruckdeschel K. Effector functions of pathogenic Yersinia species. Thromb Haemost 2007; 98:521-9; PMID:17849040 Roppenser B, R€oder A, Hentschke M, Ruckdeschel K, Aepfelbacher M. Yersinia enterocolitica differentially modulates RhoG activity in host cells. J Cell Sci 2009; 122:696-705; PMID:19208761; http://dx.doi.org/ 10.1242/jcs.040345 Matsumoto H, Young GM. Translocated effectors of Yersinia. Curr Opin Microbiol 2009; 12:94-100; PMID:19185531; http://dx.doi.org/10.1016/j. mib.2008.12.005 Fehr D, Burr SE, Gibert M, d’Alayer J, Frey J, Popoff MR. Aeromonas exoenzyme T of Aeromonas salmonicida is a bifunctional protein that targets the host cytoskeleton. J Biol Chem 2007; 282:28843-52; PMID:17656370; http://dx.doi.org/10.1074/jbc. M704797200 Garrity-Ryan L, Shafikhani S, Balachandran P, Nguyen L, Oza J, Jakobsen T, Sargent J, Fang X, Cordwell S, Matthay MA, et al. The ADP ribosyltransferase domain of Pseudomonas aeruginosa ExoT contributes to its biological activities. Infect Immun 2004; 72:546-58; PMID:14688136; http://dx.doi. org/10.1128/IAI.72.1.546-558.2004 Krall R, Schmidt G, Aktories K, Barbieri JT. Pseudomonas aeruginosa ExoT is a Rho GTPase-activating protein. Infect Immun 2000; 68:6066-8; PMID:10992524; http://dx.doi.org/10.1128/ IAI.68.10.6066-6068.2000 Barbieri JT, Sun J. Pseudomonas aeruginosa ExoS and ExoT. Rev Physiol Biochem Pharmacol 2004; 152:79-92; PMID:15375697; http://dx.doi.org/ 10.1007/s10254-004-0031-7 Mustafi S, Rivero N, Olson JC, Stahl PD, Barbieri MA. Regulation of Rab5 function during phagocytosis of live Pseudomonas aeruginosa in macrophages. Infect Immun 2013; 81:2426-36; PMID:23630954; http://dx.doi.org/10.1128/IAI.00387-13 Prehna G, Ivanov MI, Bliska JB, Stebbins CE. Yersinia virulence depends on mimicry of host Rho-family nucleotide dissociation inhibitors. Cell 2006; 126:869-80; PMID:16959567; http://dx.doi.org/ 10.1016/j.cell.2006.06.056 Navarro L, Koller A, Nordfelth R, Wolf-Watz H, Taylor S, Dixon JE. Identification of a molecular target for the Yersinia protein kinase A. Mol Cell 2007; 26:465-77; PMID:17531806; http://dx.doi.org/ 10.1016/j.molcel.2007.04.025 Groves E, Rittinger K, Amstutz M, Berry S, Holden DW, Cornelis GR, Caron E. Sequestering of Rac by the Yersinia effector YopO blocks Fcgamma receptormediated phagocytosis. J Biol Chem 2010; 285:408798; PMID:19926792; http://dx.doi.org/10.1074/jbc. M109.071035 Schmidt G. Yersinia enterocolitica outer protein T (YopT). Eur J Cell Biol 2011; 90:955-8; PMID:21255864; http://dx.doi.org/10.1016/j. ejcb.2010.12.005 Brugirard-Ricaud K, Duchaud E, Givaudan A, Girard PA, Kunst F, Boemare N, Brehelin M, Zumbihl R.

Small GTPases

157.

158.

159.

160.

161.

162.

163.

164.

165.

166.

167.

168.

169.

Site-specific antiphagocytic function of the Photorhabdus luminescens type III secretion system during insect colonization. Cell Microbiol 2005; 7:363-71; PMID:15679839; http://dx.doi.org/10.1111/j.14625822.2004.00466.x Lang AE, Schmidt G, Schlosser A, Hey TD, Larrinua IM, Sheets JJ, Mannherz HG, Aktories K. Photorhabdus luminescens toxins ADP-ribosylate actin and RhoA to force actin clustering. Science 2010; 327:1139-42; PMID:20185726; http://dx.doi.org/10.1126/ science.1184557 Yarbrough ML, Li Y, Kinch LN, Grishin NV, Ball HL, Orth K. AMPylation of Rho GTPases by Vibrio VopS disrupts effector binding and downstream signaling. Science 2009; 323:269-72; PMID:19039103; http://dx.doi.org/10.1126/science.1166382 Higa N, Toma C, Koizumi Y, Nakasone N, Nohara T, Masumoto J, Kodama T, Iida T, Suzuki T. Vibrio parahaemolyticus effector proteins suppress inflammasome activation by interfering with host autophagy signaling. PLoS Pathog 2013; 9:e1003142; PMID:23357873; http://dx.doi.org/10.1371/journal. ppat.1003142 Worby CA, Mattoo S, Kruger RP, Corbeil LB, Koller A, Mendez JC, Zekarias B, Lazar C, Dixon JE. The fic domain: regulation of cell signaling by adenylylation. Mol Cell 2009; 34:93-103; PMID:19362538; http:// dx.doi.org/10.1016/j.molcel.2009.03.008 Zekarias B, Mattoo S, Worby C, Lehmann J, Rosenbusch RF, Corbeil LB. Histophilus somni IbpA DR2/ Fic in virulence and immunoprotection at the natural host alveolar epithelial barrier. Infect Immun 2010; 78:1850-8; PMID:20176790; http://dx.doi.org/ 10.1128/IAI.01277-09 Mattoo S, Durrant E, Chen MJ, Xiao J, Lazar CS, Manning G, Dixon JE, Worby CA. Comparative analysis of Histophilus somni immunoglobulin-binding protein A (IbpA) with other fic domain-containing enzymes reveals differences in substrate and nucleotide specificities. J Biol Chem 2011; 286:32834-42; PMID:21795713; http://dx.doi.org/10.1074/jbc. M111.227603 May BJ, Zhang Q, Li LL, Paustian ML, Whittam TS, Kapur V. Complete genomic sequence of Pasteurella multocida, Pm70. Proc Natl Acad Sci U S A 2001; 98:3460-5; PMID:11248100; http://dx.doi.org/ 10.1073/pnas.051634598 Bokoch GM. Regulation of innate immunity by Rho GTPases. Trends Cell Biol 2005; 15:163-71; PMID:15752980; http://dx.doi.org/10.1016/j. tcb.2005.01.002 Shaw MH, Reimer T, Kim YG, Nu~ nez G. NOD-like receptors (NLRs): bona fide intracellular microbial sensors. Curr Opin Immunol 2008; 20:377-82; PMID:18585455; http://dx.doi.org/10.1016/j. coi.2008.06.001 Keestra AM, Baumler AJ. Detection of enteric pathogens by the nodosome. Trends Immunol 2014; 35:123-30; PMID:24268520; http://dx.doi.org/ 10.1016/j.it.2013.10.009 Keestra AM, Winter MG, Auburger JJ, Fr€assle SP, Xavier MN, Winter SE, Kim A, Poon V, Ravesloot MM, Waldenmaier JF, et al. Manipulation of small Rho GTPases is a pathogen-induced process detected by NOD1. Nature 2013; 496:233-7; PMID: 23542589; http://dx.doi.org/10.1038/nature12025 Boyer L, Magoc L, Dejardin S, Cappillino M, Paquette N, Hinault C, Charriere GM, Ip WK, Fracchia S, Hennessy E, et al. Pathogen-derived effectors trigger protective immunity via activation of the Rac2 enzyme and the IMD or Rip kinase signaling pathway. Immunity 2011; 35:536-49; PMID:22018470; http://dx.doi.org/10.1016/j.immuni.2011.08.015 B€ohmer J, Jung M, Sehr P, Fritz G, Popoff M, Just I, Aktories K. Active site mutation of the C3-like ADPribosyltransferase from Clostridium limosum–analysis of glutamic acid 174. Biochemistry 1996; 35:282-9;

17

170.

171.

172.

Downloaded by [University of Southern Queensland] at 10:01 13 March 2015

173.

174.

175.

176.

177.

178.

18

PMID:8555186; http://dx.doi.org/10.1021/ bi951784C Just I, Mohr C, Schallehn G, Menard L, Didsbury JR, Vandekerckhove J, van Damme J, Aktories K. Purification and characterization of an ADP-ribosyltransferase produced by Clostridium limosum. J Biol Chem 1992; 267:10274-80; PMID:1587816 Just I, Schallehn G, Aktories K. ADP-ribosylation of small GTP-binding proteins by Bacillus cereus. Biochem Biophys Res Commun 1992; 183:931-6; PMID:1567406; http://dx.doi.org/10.1016/S0006291X(05)80279-7 Sugai M, Enomoto T, Hashimoto K, Matsumoto K, Matsuo Y, Ohgai H, Hong YM, Inoue S, Yoshikawa K, Suginaka H. A novel epidermal cell differentiation inhibitor (EDIN): purification and characterization from Staphylococcus aureus. Biochem Biophys Res Commun 1990; 173:92-8; PMID:2256941; http:// dx.doi.org/10.1016/S0006-291X(05)81026-5 Yamaguchi T, Hayashi T, Takami H, Ohnishi M, Murata T, Nakayama K, Asakawa K, Ohara M, Komatsuzawa H, Sugai M. Complete nucleotide sequence of a Staphylococcus aureus exfoliative toxin B plasmid and identification of a novel ADP-ribosyltransferase, EDIN-C. Infect Immun 2001; 69:776071; PMID:11705958; http://dx.doi.org/10.1128/ IAI.69.12.7760-7771.2001 Yamaguchi T, Nishifuji K, Sasaki M, Fudaba Y, Aepfelbacher M, Takata T, Ohara M, Komatsuzawa H, Amagai M, Sugai M. Identification of the Staphylococcus aureus etd pathogenicity island which encodes a novel exfoliative toxin, ETD, and EDIN-B. Infect Immun 2002; 70:5835-45; PMID:12228315; http:// dx.doi.org/10.1128/IAI.70.10.5835-5845.2002 Chaves-Olarte E, L€ow P, Freer E, Norlin T, Weidmann M, von Eichel-Streiber C, Thelestam M. A novel cytotoxin from Clostridium difficile serogroup F is a functional hybrid between two other large clostridial cytotoxins. J Biol Chem 1999; 274:11046-52; PMID:10196187; http://dx.doi.org/10.1074/ jbc.274.16.11046 Just I, Selzer J, Hofmann F, Green GA, Aktories K. Inactivation of Ras by Clostridium sordellii lethal toxin-catalyzed glucosylation. J Biol Chem 1996; 271:10149-53; PMID:8626575; http://dx.doi.org/ 10.1074/jbc.271.17.10149 Genth H, Hofmann F, Selzer J, Rex G, Aktories K, Just I. Difference in protein substrate specificity between hemorrhagic toxin and lethal toxin from Clostridium sordellii. Biochem Biophys Res Commun 1996; 229:370-4; PMID:8954906; http://dx.doi.org/ 10.1006/bbrc.1996.1812 Selzer J, Hofmann F, Rex G, Wilm M, Mann M, Just I, Aktories K. Clostridium novyi alpha-toxin-catalyzed incorporation of GlcNAc into Rho subfamily proteins. J Biol Chem 1996; 271:25173-7; PMID:

179.

180.

181.

182.

183.

184.

185.

186.

187.

188.

8810274; http://dx.doi.org/10.1074/jbc.271.41. 25173 Nagahama M, Ohkubo A, Oda M, Kobayashi K, Amimoto K, Miyamoto K, Sakurai J. Clostridium perfringens TpeL glycosylates the Rac and Ras subfamily proteins. Infect Immun 2011; 79:905-10; PMID: 21098103; http://dx.doi.org/10.1128/IAI.01019-10 Guttenberg G, Hornei S, Jank T, Schwan C, L€ u W, Einsle O, Papatheodorou P, Aktories K. Molecular characteristics of Clostridium perfringens TpeL toxin and consequences of mono-O-GlcNAcylation of Ras in living cells. J Biol Chem 2012; 287:24929-40; PMID:22665487; http://dx.doi.org/10.1074/jbc. M112.347773 Horiguchi Y, Inoue N, Masuda M, Kashimoto T, Katahira J, Sugimoto N, Matsuda M. Bordetella bronchiseptica dermonecrotizing toxin induces reorganization of actin stress fibers through deamidation of Gln63 of the GTP-binding protein Rho. Proc Natl Acad Sci U S A 1997; 94:11623-6; PMID:9326660; http:// dx.doi.org/10.1073/pnas.94.21.11623 Flatau G, Lemichez E, Gauthier M, Chardin P, Paris S, Fiorentini C, Boquet P. Toxin-induced activation of the G protein p21 Rho by deamidation of glutamine. Nature 1997; 387:729-33; PMID:9192901; http://dx.doi.org/10.1038/42743 Schmidt G, Sehr P, Wilm M, Selzer J, Mann M, Aktories K. Gln 63 of Rho is deamidated by Escherichia coli cytotoxic necrotizing factor-1. Nature 1997; 387:725-9; PMID:9192900; http://dx.doi.org/ 10.1038/42735 Hoffmann C, Schmidt G. CNF and DNT. Rev Physiol Biochem Pharmacol 2004; 152:49-63; PMID:15549605; http://dx.doi.org/10.1007/s10254004-0026-4 Hardt WD, Urlaub H, Galan JE. A substrate of the centisome 63 type III protein secretion system of Salmonella typhimurium is encoded by a cryptic bacteriophage. Proc Natl Acad Sci U S A 1998; 95:2574-9; PMID:9482928; http://dx.doi.org/10.1073/ pnas.95.5.2574 Ohya K, Handa Y, Ogawa M, Suzuki M, Sasakawa C. IpgB1 is a novel Shigella effector protein involved in bacterial invasion of host cells. Its activity to promote membrane ruffling via Rac1 and Cdc42 activation. J Biol Chem 2005; 280:24022-34; PMID:15849186; http://dx.doi.org/10.1074/jbc.M502509200 Bulgin RR, Arbeloa A, Chung JC, Frankel G. EspT triggers formation of lamellipodia and membrane ruffles through activation of Rac-1 and Cdc42. Cell Microbiol 2009; 11:217-29; PMID:19016787; http://dx.doi.org/10.1111/j.1462-5822.2008.01248.x Bulgin R, Arbeloa A, Goulding D, Dougan G, Crepin VF, Raymond B, Frankel G. The T3SS effector EspT defines a new category of invasive enteropathogenic E. coli (EPEC) which form intracellular actin pedestals.

Small GTPases

189.

190.

191.

192.

193.

194.

195.

196.

197.

PLoS Pathog 2009; 5:e1000683; PMID:20011125; http://dx.doi.org/10.1371/journal.ppat.1000683 Matsuzawa T, Kuwae A, Yoshida S, Sasakawa C, Abe A. Enteropathogenic Escherichia coli activates the RhoA signaling pathway via the stimulation of GEFH1. EMBO J 2004; 23:3570-82; PMID:15318166; http://dx.doi.org/10.1038/sj.emboj.7600359 Selyunin AS, Sutton SE, Weigele BA, Reddick LE, Orchard RC, Bresson SM, Tomchick DR, Alto NM. The assembly of a GTPase-kinase signalling complex by a bacterial catalytic scaffold. Nature 2011; 469:107-11; PMID:21170023; http://dx.doi.org/ 10.1038/nature09593 Huang Z, Sutton SE, Wallenfang AJ, Orchard RC, Wu X, Feng Y, Chai J, Alto NM. Structural insights into host GTPase isoform selection by a family of bacterial GEF mimics. Nat Struct Mol Biol 2009; 16:853-60; PMID:19620963; http://dx.doi.org/ 10.1038/nsmb.1647 Upadhyay A, Wu HL, Williams C, Field T, Galyov EE, van den Elsen JM, Bagby S. The guanine-nucleotide-exchange factor BopE from Burkholderia pseudomallei adopts a compact version of the Salmonella SopE/SopE2 fold and undergoes a closed-to-open conformational change upon interaction with Cdc42. Biochem J 2008; 411:485-93; PMID:18052936; http://dx.doi.org/10.1042/BJ20071546 Black DS, Bliska JB. The RhoGAP activity of the Yersinia pseudotuberculosis cytotoxin YopE is required for antiphagocytic function and virulence. Mol Microbiol 2000; 37:515-27; PMID:10931345; http://dx.doi. org/10.1046/j.1365-2958.2000.02021.x Von Pawel-Rammingen U, Telepnev MV, Schmidt G, Aktories K, Wolf-Watz H, Rosqvist R. GAP activity of the Yersinia YopE cytotoxin specifically targets the Rho pathway: a mechanism for disruption of actin microfilament structure. Mol Microbiol 2000; 36:737-48; PMID:10844661; http://dx.doi.org/ 10.1046/j.1365-2958.2000.01898.x Mohammadi S, Isberg RR. Yersinia pseudotuberculosis virulence determinants invasin, YopE, and YopT modulate RhoG activity and localization. Infect Immun 2009; 77:4771-82; PMID:19720752; http:// dx.doi.org/10.1128/IAI.00850-09 Goehring UM, Schmidt G, Pederson KJ, Aktories K, Barbieri JT. The N-terminal domain of Pseudomonas aeruginosa exoenzyme S is a GTPase-activating protein for Rho GTPases. J Biol Chem 1999; 274:36369-72; PMID:10593930; http://dx.doi.org/10.1074/ jbc.274.51.36369 Shao F, Merritt PM, Bao Z, Innes RW, Dixon JE. A Yersinia effector and a Pseudomonas avirulence protein define a family of cysteine proteases functioning in bacterial pathogenesis. Cell 2002; 109:575-88; PMID:12062101; http://dx.doi.org/10.1016/S00928674(02)00766-3

Volume 5 Issue 3

Bacterial factors exploit eukaryotic Rho GTPase signaling cascades to promote invasion and proliferation within their host.

Actin cytoskeleton is a main target of many bacterial pathogens. Among the multiple regulation steps of the actin cytoskeleton, bacterial factors inte...
713KB Sizes 5 Downloads 5 Views