<<

Histotoxic Clostridial MASAHIRO NAGAHAMA,1 MASAYA TAKEHARA,1 and JULIAN I. ROOD2 1Department of Microbiology, Faculty of Pharmaceutical Sciences, Tokushima Bunri University, Yamashiro-cho, Tokushima 770-8514, Japan; 2Infection and Immunity Program, Monash Biomedicine Discovery Institute and Department of Microbiology, Monash University, Clayton, Victoria 3800, Australia

ABSTRACT The pathogenesis of clostridial myonecrosis or Although the histotoxic cause severe, gas involves an interruption to the blood supply to rapidly developing infections, these potent toxigenic the infected tissues, often via a traumatic wound, anaerobic still must be regarded as opportunistic patho- growth of the infecting clostridial cells, the production of gens. All of these infections require predisposing extracellular toxins, and toxin-mediated cell and tissue damage. This review focuses on host-pathogen interactions conditions to cause disease, often a traumatic wound in perfringens-mediated and Clostridium that enables the entry of spores from the soil or gastro- septicum-mediated myonecrosis. The major toxins involved intestinal tract to gain access to ischemic internal organs are C. perfringens α-toxin, which has phospholipase C and or soft tissues of the body. Alternatively, in nontrau- sphingomyelinase activity, and C. septicum α-toxin, matic myonecrosis caused by C. septicum, gastrointes- β a -pore-forming toxin that belongs to the aerolysin family. tinal malignancy is commonly associated with the onset ff Although these toxins are cytotoxic, their e ects on host of disease, with tumor development leading to ulceration cells are quite complex, with a range of intracellular cell signaling pathways induced by their action on host cell membranes. and of the gastrointestinal mucosa and entry of the gastrointestinal microbiota into the circulation and subsequent at distal sites (3, 4). Clostridial GENERAL CHARACTERISTICS OF necrotizing soft tissue infections, especially those caused HISTOTOXIC CLOSTRIDIAL INFECTIONS by C. perfringens, , and Clostri- Many pathogenic clostridial species cause potentially dium novyi, also have resulted from the subcutaneous – fatal soft tissue infections in humans and animals be- injection of contaminated black tar heroin (5 7). cause of their ability to produce extracellular protein toxins (1, 2). The ability of these pathogens to produce Received: 19 March 2018, Accepted: 23 March 2018, spores that are resistant to environmental stress is an Published: 26 July 2019 Editors: Vincent A. Fischetti, The Rockefeller University, New York, important factor in the epidemiology of these diseases. NY; Richard P. Novick, Skirball Institute for Molecular Medicine, Disease pathogenesis involves the growth of the clos- NYU Medical Center, New York, NY; Joseph J. Ferretti, Department tridial pathogen in the tissues and extensive tissue de- of Microbiology & Immunology, University of Oklahoma Health Science Center, Oklahoma City, OK; Daniel A. Portnoy, Department struction, which is the result of the action of extracellular of Molecular and Cellular Microbiology, University of California, toxins. Typical histotoxic clostridial diseases include Berkeley, Berkeley, CA; Miriam Braunstein, Department of Microbiology and Immunology, University of North Carolina-Chapel human or myonecrosis and in Hill, Chapel Hill, NC, and Julian I. Rood, Infection and Immunity cattle. Although there are several clostridial species that Program, Monash Biomedicine Discovery Institute, Monash are responsible for these syndromes (Table 1), this University, Melbourne, Australia review will focus on histotoxic infections caused by Citation: Nagahama M, Takehara M, Rood JI. 2019. Histotoxic clostridial infections. Microbiol Spectrum 6(4):GPP3-0024-2018. and , doi:10.1128/microbiolspec.GPP3-0024-2018. primarily because these are the species that have been Correspondence: Masahiro Nagahama, [email protected] the subject of the most extensive molecular and func- -u.ac.jp or Julian I. Rood, [email protected] tional studies. © 2019 American Society for Microbiology. All rights reserved.

ASMscience.org/MicrobiolSpectrum 1 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

TABLE 1 Histotoxic clostridial infections

Syndromea Causative agent Host Major toxinsb References Traumatic gas gangrene C. perfringens type A Humans α-toxin 12, 17 Perfringolysin O (θ-toxin) Traumatic gas gangrene C. novyi Humans α-toxin 151 C. septicum α-toxin 129 C. histolyticum Lethal factor 152 C. sordellii Lethal toxin (TcsL) 153 Hemorrhagic toxin (TcsH) Nontraumatic gas gangrene C. septicum Humans α-toxin 3, 4 C. fallax Humans and animals Not known 154 Necrotizing pancreatitis C. perfringens type A Humans α-toxin 155 Clostridial myonecrosis C. perfringens types A, B, C, and D Sheep, cattle, and other animals α-toxin 29 β-toxin ε-toxin Blackleg C. chauvoei Cattle, sheep, and other ruminants Cytolysin A (CctA) 156

aThe terms “gas gangrene” and “clostridial myonecrosis” are used interchangeably in this review. bNote that the different α-toxins referred to in this column are not necessarily structurally or functionally related.

PATHOGENESIS OF CLOSTRIDIAL coli, and therefore devoid of any other C. perfringens MYONECROSIS CAUSED BY toxins, showed that the C-terminal domain of α-toxin C. PERFRINGENS AND C. SEPTICUM was immunoprotective (14). Second, mutation of the As previously described (8), the pathogenesis of these α-toxin structural gene (plc or cpa) abrogated the ability infections involves three distinct stages, irrespective of of the bacterium to cause clostridial myonecrosis in a whether the source of infection involves a traumatic murine model. was restored by complemen- wound. Stage 1 involves an interruption to the blood tation in trans with a recombinant plasmid containing supply such that the redox potential in the tissues drops the wild-type plc gene (15), thereby fulfilling molecular to a level that facilitates spore germination and/or the Koch’s postulates and providing proof of the essential growth of the infecting clostridial cells. Stage 2 involves role of this toxin in disease. bacterial growth and establishment of the conditions for The other major toxin produced by C. perfringens toxin production. In C. perfringens, regulation of toxin type A is perfringolysin O, or θ-toxin, a pore-forming production primarily utilizes the VirSR two-component toxin that is a member of the cholesterol-dependent cy- signal transduction system (9) and an accessory growth tolysin family (16, 17). Although mutation of the per- regulator-like quorum sensing system (10, 11), as well fringolysin O structural gene, pfoA, did not eliminate as other regulatory networks. Stage 3 encompasses the the ability to cause disease (18, 19) subsequent studies toxin-mediated cell and tissue damage that leads to ne- of plcpfoA double mutants (20) provided evidence that crosis, systemic toxicity, and clinical disease (8). perfringolysin O has a synergistic effect with α-toxin on C. perfringens type A is the major cause of traumatic disease pathogenesis (20, 21). gas gangrene, although other histotoxic clostridia may Recent studies have involved the concurrent analy- also be responsible for this sporadic, but fulminant sis of the transcriptomes of both the host and C. per- and often fatal, disease (Table 1)(1). The major toxin fringens in a murine myonecrosis infection (22). The produced by gas gangrene strains of C. perfringens is results showed that many host genes involved in the α-toxin, a zinc metallophospholipase that has both innate immune response to infection were upregulated phospholipase C and sphingomyelinase activity (12). in C. perfringens-infected muscle tissues. In the C. per- α-Toxin was the first bacterial toxin to be shown to have fringens cells, upregulated genes included those en- enzymatic activity (13). Two independent studies using coding potential adhesins and proteins associated with different experimental approaches have shown that the cell envelope. α-toxin is essential for C. perfringens-mediated myo- C. septicum is the major causative agent of nontrau- necrosis. First, immunization studies in mice using re- matic gas gangrene, often associated with a gastro- combinant α-toxin variants purified from Escherichia intestinal malignancy (4). The major toxin produced

2 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections by C. septicum is also called α-toxin, but it is not related to C. perfringens α-toxin. C. septicum α-toxin is an aerolysin-like pore-forming toxin that is secreted as an inactive prototoxin that is cleaved near the C terminus by membrane-bound furin or furin-like proteases to form the active toxin (23). The toxin then oligomerizes on the membrane and subsequently inserts to form a membrane pore, which ultimately leads to cell (24). Genetic studies have shown that α-toxin is essential for C. septicum-mediated myonecrosis in mice (25). Muta- tion of the α-toxin structural gene, csa, leads to a loss of virulence, which is restored by complementation with the wild-type csa gene.

STRUCTURE AND FUNCTION OF C. PERFRINGENS α-TOXIN FIGURE 1 Structure of α-toxin. The structure of α-toxin was α obtained from the Protein Data Bank (code 1QM6) and was -Toxin is a 43-kDa single polypeptide that is a hemo- created using MacPyMOL. A ribbon representation is shown, lytic, cytotoxic, dermonecrotic, and lethal extracellu- colored from blue (N-terminal) to red (C-terminal). Two zinc lar toxin (26–29). Moreover, it induces contractions metal ions in the active site are shown as red spheres. of blood vessels and smooth muscle, aggregation of platelets, superoxide production, and cytokine release. Importantly, together with perfringolysin O, it is re- domain of pancreatic lipase, the N-terminal domain of sponsible for the hallmark histopathological feature of soybean lipoxygenase, and the C2 phospholipid-binding C. perfringens-mediated myonecrosis, the paucity of the domains of eukaryotic intracellular proteins, such as polymorphonuclear leukocyte (PMNL) influx into the synaptotagmin I, which are involved in signal trans- infected lesion (15, 19–21, 30). duction and vesicular transport (37, 38). Determination of the crystal structure of α-toxin The active site in the N-terminal domain is situated revealed that it has an α-helical N-terminal domain within a solvent-accessible cleft that contains two zinc (amino acids 1 to 246) that is responsible for its enzy- ions and an exchangeable divalent cation. His-148 and matic activity, a flexible linker (amino acids 247 to 255), Glu-152 dock with one zinc ion, which is critical for the and a C-terminal (amino acids 256 to 370) β-sandwich catalytic site of the toxin. His-11 and Asp-130 tightly membrane-binding domain (31)(Fig. 1). The two do- dock with the other zinc ion, which is needed for main- mains also are linked by hydrophobic surface interac- tenance of the structure. His-68, His-126, and His-136 tions between their neighboring faces. The N-terminal dock with an exchangeable divalent cation, which is domain has significant sequence similarity to phospho- required for binding to target cell membranes (39). lipase C enzymes from cereus (32), Clostridium The external loop regions (residues 70 to 90, 135 to 150, bifermentans (33), and monocytogenes (34). It 205 to 215, 265 to 275, 290 to 300, and 330 to 340) comprises nine densely packed α-helices and resembles bind to phospholipids in the host cell membranes (31). the structure of B. cereus phospholipase C (31, 35), Thr-74 is present in one of these loops, and although which has enzymatic activity, but is neither toxic nor the substitution derivative T74I is still active and can hemolytic. There are three loops in the N-terminal do- cleave a water-soluble substrate, its membrane-damaging main: loop 1 (residues 70 to 90), loop 2 (residues 135 activity toward red blood cells and phosphatidylcholine- to 150), and loop 3 (residues 205 to 215). These loops cholesterol liposomes is decreased compared to wild-type serve a vital role in the direct interaction between the toxin (40, 41). However, the structure of the T74I variant toxin and the membrane-binding component. The C- is identical to that of the wild-type toxin (41). The T74I terminal domain comprises eight-stranded antiparallel substitution results in altered recognition of the glycerol β-sheets (31). This domain plays a role in membrane moiety in the phospholipid, without affecting the func- binding and is required for hemolytic and toxic activity tion of the active site, supporting the hypothesis that the (28, 36). It is absent from nontoxic phospholipase C en- amino acid 70 to 90 loop is important for phospholipid zymes (31). It has structural similarity to the C-terminal recognition (42, 43).

ASMscience.org/MicrobiolSpectrum 3 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

Computational modeling of α-toxin and phosphati- to the membrane and that the amino acid 70 to 90 loop dylcholine revealed that Tyr-57 and Tyr-65, which exist and Tyr-57 and Tyr-65 serve an important role in the re- at an external site near the catalytic cleft and are in cognition of target membranes (Fig. 2). Other studies proximity with the 70 to 90 loop, are responsible for have shown that substitution of the C-terminal domain substrate interaction (44). The replacement of these residues Asp-269, Asp-336, Tyr-275, and Tyr-331, Tyr residues with Ala did not have any effect on phos- which were predicted to be involved in membrane bind- pholipase C and sphingomyelinase activity but had a ing (31), were required for hemolytic and cytotoxic ac- marked effect on the membrane-damaging activity (he- tivity (45). molysis of sheep red cells and disruption of liposomes). These residues exist in or near the entrance of the cata- INTERACTION OF C. PERFRINGENS lytic cleft and are essential for the membrane-damaging α-TOXIN WITH HOST CELLS effect of α-toxin, because they are required for entry of the catalytic cleft into the hydrophobic region of plasma Membrane-Damaging Activity on membranes (44). Phospholipid Bilayers Several studies have shown that the calcium-binding The effect of α-toxin on host tissues is caused by the C-terminal domain is responsible for membrane bind- phospholipase C- and sphingomyelinase-mediated cleav- ing (28, 31, 36, 37). Two acrylodan-labeled C-terminal age of phospholipids in the host cell membranes. The domain derivatives, S263C and S365C, bound to lipid resultant disruption of the phospholipid bilayer together bilayers and displayed a chromogenic shift, demonstrat- with the end products of these cleavage reactions, diacyl- ing entry of the C-domain of the toxin into the hydro- glycerol and ceramide, respectively, activates lipid me- phobic regions of the liposomes (36). The data provide tabolism in host cell membranes and plays a role in host evidence that the C-terminal domain of the toxin binds cell damage (27, 28).

FIGURE 2 Binding model of α-toxin and membrane phospholipids. The structure of α-toxin was obtained from the Protein Data Bank (code 1QM6) and was created using MacPyMOL. Phospholipids are shown as light gray spheres for head groups and light gray lines for tail groups. The head group of a phospholipid, from the outer membrane into the active site of α-toxin, is displayed. Amino acids that may play a role in membrane inter- action are shown.

4 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections

The composition of the phospholipid bilayer modu- Treatment of rabbit with α-toxin pro- lates the extent of the α-toxin-mediated membrane dis- motes the formation of diacylglycerol, which is fol- ruption (46). In another study that measured the release lowed by adhesion of the cells to extracellular matrix of carboxyfluorescein from various liposomes, α-toxin- proteins and the generation of superoxide ions (Fig. 3B) mediated release, or liposomal leakage, decreased with (51). α-Toxin induces diacylglycerol generation through increasing chain length of the hydrocarbon residues activation of intrinsic phospholipase C enzymes by a in the phospholipid bilayer (40). It was concluded that pertussis toxin-sensitive GTP-binding protein. More- membrane disruption was related to the fluidity of the over, it increases the phosphorylation of neurotrophic membrane and double bonds in the acyl chain. Sub- tyrosine kinase receptor type 1 (TrkA), leading to 3- sequently, several phosphatidylcholines (C18:0/C18:1) phosphoinositide-dependent protein kinase 1 activation with a double bond in the sn-2 fatty acyl chain were (52). Therefore, α-toxin independently causes stimula- prepared (47). The phase transition temperature value tion of intrinsic phospholipase C and protein kinase 1 was lowest when the cis-double bond was positioned at activity through TrkA receptor activation. Both phe- the center of the C18:1 sn-2 acyl chain and increased nomena synergistically augment the activity of protein continuously as the double bond was moved toward kinase C θ, leading to the formation of superoxide ions either end of the acyl chain. α-toxin-mediated carboxy- via activation of the mitogen-activated protein kinase fluorescein release from various liposomes was highest (MAPK) cascade (Fig. 3B)(52). In human lung carci- when the cis-double bond was positioned at the center of noma (A549) cells, α-toxin mediates the activation of the the sn-2 fatty acyl chain, resulting in a bell-shaped curve p38 MAPK, NF-κB, and extracellular signal-regulated (47). Accordingly, the membrane-disrupting activity of kinase 1/2 (ERK1/2) through TrkA activation, which the toxin is strongly associated with the membrane flu- leads to interleukin-8 (IL-8) release (53)(Fig. 3B). idity of target membranes. Treatment of rabbit erythrocytes with α-toxin in- There is evidence that α-toxin also promotes per- duces the biphasic formation of phosphatidic acid fringolysin O-induced membrane damage in phospholipid (54)(Fig. 3A). α-Toxin-induced formation of phospha- bilayers. Hydrolysis of phosphatidylcholine residues of tidic acid is activated in the earlier stage (30 s) by in- liposomes that contained cholesterol increased the amount trinsic phospholipase C and in the later stage (ca. 25 offreecholesterolinthelipidbilayerandenhanced min) by intrinsic phospholipase D. Both processes perfringolysin O-dependent cytolytic activity (48). These involve stimulation of the G-protein-signaling pathway data may at least partly explain the synergistic effects of α- by toxin-mediated membrane phospholipid cleavage toxin and perfringolysin O that were previously observed (55)(Fig. 3A). in C. perfringens infections of mice (20). Sheep erythrocytes have higher amounts of sphingo- myelin but do not contain phosphatidylcholine. Cleav- Activation of Phospholipid age of sphingomyelin by α-toxin leads to increased levels Metabolism and of ceramide and sphingosine 1-phosphate (56), which is α-toxin provokes contraction of isolated rat ileum and essential for toxin-induced hemolysis (57) and activates aorta tissues in a dose-dependent manner through stim- the sphingomyelin metabolism system (Fig. 3A)(56). ulation of the phospholipid metabolic pathway (49, 50). The resultant cleavage of N-nervonoic sphingomyelin In isolated rat organs, α-toxin activates the phospha- (C24:1-SM) occurs in a detergent-resistant membrane tidylinositol metabolic pathway and then the arachi- component (57)(Fig. 3A). Pertussis toxin blocks the donic acid cascade, thereby promoting thromboxane production of C24:1-ceramide caused by α-toxin and production. Contraction caused by the toxin was evoked blocks red cell hemolysis, which indicates that intrinsic by thromboxane A2 generation induced by arachidonic sphingomyelinase, which can cleave C24:1-SM to C24:1- acid (49, 50). The metabolism of arachidonic acid results ceramide, is regulated by a pertussis toxin-sensitive G- in the production of bioactive lipid mediators, including protein in plasma membranes. In addition, the activation leukotrienes, thromboxanes, and prostaglandins (27). of Rho protein by α-toxin increases sphingosine kinase These bioactive lipid mediators play key roles in en- activity. Subsequently, the ceramides are quickly me- hancing the local inflammatory reaction and lead to tabolized to sphingosine, indicating that the toxin also vasoconstriction, which is responsible for inducing an- activates ceramidases (57). These results indicate that oxia in the early stages of bacterial infections. These the cleavage of C24:1-SM to sphingosine 1-phosphate by effects presumably contribute to the capacity of C. per- α-toxin in lipid rafts is essential for toxin-induced he- fringens to proliferate in the host and cause disease. molysis of sheep red cells (Fig. 3A).

ASMscience.org/MicrobiolSpectrum 5 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

FIGURE 3 Relationship between phospholipid metabolism and the biological activities of α-toxin. (A) Signaling events involved in α-toxin-induced hemolysis of sheep or rabbit – erythrocytes. (B) Signaling events involved in α-toxin-activated generation of O2 in neutrophils and in α-toxin-mediated release of IL-8 from A549 cells. Abbreviations: SM, sphingomyelin; PIP2, phosphatidylinositol 4,5-bisphosphate; PLC, phospholipase C; DG, diacylglycerol; PC, phosphatidylcholine; SMase, sphingomyelinase; PLD, phospholipase D; PA, phosphatidic acid; PKCθ, protein kinase Cθ; PI3K, phosphatidylinositol 3-kinase; PIP3, phosphatidylinositol 3,4,5-trisphosphate; CDase, ceramidase; CER, ceramide; SPH, sphingosine; S1P, sphingosine 1-phosphate; DGK, DG kinase; PDK1, phosphatidylinositide dependent kinase 1; DRM, detergent-resistant membrane; TrkA, neurotrophic tyrosine kinase receptor type1; MAPKK, mitogen-activated protein kinase kinase; Erk1/2, extra- cellular signal-regulated kinase 1/2.

Interaction of α-Toxin with Gangliosides the ganglioside recognition site of botulinum and teta- Amino acid sequence analysis of botulinum toxin and nus toxins, significantly reduced binding to ganglioside toxin showed that the ganglioside-GT1b- and GM1a-containing liposomes. W84A and Y85A deriva- lactose-binding site is determined by the existence of the tives had markedly decreased capabilities to stimulate conserved peptide motif, H.....SXWY.....G (ca. 30 amino TrkA (59). Ganglioside GM1a assembles with a TrkA acids [aa]), with the Tyr and Trp residues being partic- molecule on the cell surface and stimulates TrkA and ularly critical (58). A similar motif (H.....SWY.....G, ERK1/2 (60). It has been reported that ganglioside GM1 amino acids 68 to 93) is present in α-toxin (53), in the clustering leads to the condensation of TrkA in lipid rafts region (amino acids 55 to 93) that is located at the in- and the stimulation of downstream signaling targets terface between the N-terminal and C-terminal domains (61). α-Toxin-mediated diacylglycerol production leads (42). It has been demonstrated that the 70 to 90 loop to diacylglycerol flip-flop movement that affects the in α-toxin plays a role in binding to the cell surface dynamics of cell membranes (Fig. 4). Therefore, it serves (43)(Fig. 2). Computational docking studies indicated in GM1a clustering and intrinsic phospholipase C-γ1 that Trp-84 binds to the sialic acid residue on ganglio- activation through TrkA, triggering various signal side GM1a via an aromatic stacking interaction and a transduction activities (62–64). hydrogen-bonding interaction and that Tyr-85 associ- α-toxin causes higher cytotoxicity in a ganglioside- ates with the galactosamine residue on GM1a via an deficient cell line (Don Q) (65) and can be internalized α-glycosidic bonding interaction (59). Replacement of by dynamin-mediated endocytosis (66, 67). The toxin Trp-84 and Tyr-85 in α-toxin, which correspond to binds to caveolin-1 both on the cell membranes and in

6 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections

FIGURE 4 Model of α-toxin-induced membrane dynamics and clustering of the GM1a and TrkA complex. α-Toxin binds to GM1a and, through the phospholipase C activity of the toxin itself, causes diacylglycerol (DG) production at the outer membrane. The flip-flop movement of DG affects membrane dynamics, facilitating clustering of GM1a and TrkA activation by phosphorylation. Activation of TrkA results in the activation of PLCγ-1, leading to the increased production of DG. Finally, DG formation results in the enhanced activation of TrkA, triggering activation of signal transduction pathways, which induce activation. Abbreviations: PIP2, phosphatidylinositol 4,5-bisphosphate; PLC, phospholipase C; DG, diacylglycerol; PC, phosphatidylcholine; TrkA, neurotrophic tyro- sine kinase receptor type 1. endosomal vesicles, and it also induces ceramide pro- is sustained in a steady state through granulopoiesis, duction in the cell membrane (28), which generates neg- which is accelerated to replenish neutrophils during bac- ative membrane curvature and facilitates endocytosis terial infection (71–73). Pattern recognition receptors, (68). Internalized α-toxin is transported to the early and such as Toll-like receptors (TLRs), are responsible for late endosomes and lysosomes. Lysosomes are damaged the recognition of structural components of microor- during progressive α-toxin internalization into Don Q ganisms (74, 75). TLR2 and TLR4 have been identified cells. In addition, the internalization of α-toxin is also as being pivotal for the recognition of Gram-positive or mediated through the stimulation of MEK/ERK signal- Gram-negative bacteria by recognizing cell wall com- ing (66, 67). It has been reported that the toxin causes ponents such as peptidoglycan and bacterial endotoxin reactive species generation via PKC, MEK/ERK, (lipopolysaccharide), respectively (76–79). During bac- and NF-κB signaling activities in Don Q cells (66, 67). terial infection, these bacterial components stimulate Therefore, α-toxin perturbs the host cell signaling cas- the production of granulocyte colony-stimulating factor cade and acts not only on the cell membrane, but also in (G-CSF), which is a glycoprotein that promotes the the cytoplasm. proliferation, survival, and differentiation of neutrophils and their progenitor cells (80) from endothelial cells and Effect of α-Toxin on Hematopoietic Cells monocytes, resulting in the acceleration of granulopoie- The first line of defense of the innate immune system sis (81). G-CSF-deficient and G-CSF receptor-deficient involves neutrophils, which play an important role in the mice exhibit chronic neutropenia and reduced infection- phagocytosis and subsequent elimination of pathogenic driven granulopoiesis, leading to impaired ability to bacteria (69, 70). Normally, the number of neutrophils eliminate infecting bacteria (82, 83). Thus, granulo-

ASMscience.org/MicrobiolSpectrum 7 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al. poiesis is precisely regulated to replenish neutrophils infected muscle, in an α-toxin-dependent manner (90). during bacterial infection. However, C. perfringens can α-Toxin inhibits neutrophil differentiation and impairs still cause life-threatening infections, and the mecha- the replenishment of differentiated mature neutrophils in nisms by which C. perfringens avoids these host defenses the peripheral circulation, resulting in reduced recruit- are still not completely understood. ment of neutrophils to the C. perfringens infection site. C. perfringens-mediated myonecrosis progresses so The half-life of peripheral neutrophils is less than 12 rapidly that death, which is associated with muscle de- hours in vivo (91), meaning that the number of mature struction, , and multiple organ failure, sometimes neutrophils required to counter a bacterial infection is precedes an accurate diagnosis (84). C. perfringens in- sustained through continuous granulopoiesis. Therefore, fection is characterized by a greatly reduced PMNL in- blockage of both granulopoiesis and PMNL diapedesis flux to the site of the infection (19, 85–87), suggesting into the tissues, as a result of platelet aggregation, leads that the bacteria disrupt neutrophil-based innate immu- to a reduction in cell numbers in a short period of time. nity to evade host defenses. α-Toxin mediates the forma- Thus, C. perfringens infection impairs the host immune tion of platelet-leukocyte aggregates, leading to vascular system, which could explain the both scarcity of PMNL occlusion, and induces a marked reduction in micro- at the site of infection and a propensity for polymicrobial vascular perfusion (45, 86, 88), which is thought to infections in many patients (Fig. 5). impede neutrophil extravasation (89). Recently, it was α-Toxin is essential for disease pathogenesis (18–21, shown using a mouse model of infection that C. per- 30), and blockage of neutrophil differentiation is de- fringens infection reduced mature neutrophils in bone pendent on its phospholipase C and sphingomyelinase marrow, in peripheral blood, and in C. perfringens- activities (90). In human lymphocytes, S. aureus sphin-

FIGURE 5 Inhibition of granulopoiesis and erythropoiesis in a C. perfringens-infected host. Infection with C. perfringens diminishes mature hematopoietic cells in the bone marrow, leading to impairment of replenishment of differentiated neutrophils and eryth- rocytes in the peripheral circulation and resulting in host innate immune deficiency. The major of C. perfringens, α-toxin, plays an important role in this phe- nomenon by interfering with cell differentiation of myeloblasts and erythroblasts.

8 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections gomyelinase disrupts lipid rafts, which are cholesterol- raft integrity in erythroid progenitors, leading to the rich plasma membrane microdomains (92). Lipid rafts impairment of erythroid differentiation. During eryth- function as platforms of signaling molecules involved in roblast enucleation, the clustering of lipid rafts at the the regulation of cell differentiation in many cell types furrow between incipient reticulocytes and pyrenocytes (93, 94). Ceramide, the hydrolysis product of sphingo- is observed, and an inhibitor of lipid rafts impedes myelinase, also is known to be a lipid messenger impli- enucleation (100). Therefore, lipid rafts play an impor- cated in various cellular responses, including apoptosis, tant role in the regulation of erythropoiesis, but the differentiation, and cell growth (95). Recently, C. per- detailed mechanism by which the disturbance in lipid fringens α-toxin was reported to perturb the integrity raft integrity by α-toxin impairs erythroid differentiation of lipid rafts, leading to the blockage of neutrophil dif- remains unclear. ferentiation (90). The detailed molecular mechanism behind this blockage has not been elucidated, but it may be important to focus on the relationship between the STRUCTURE AND HOST CELL role of sphingomyelinase and cell differentiation to shed INTERACTIONS OF C. PERFRINGENS further light on host-pathogen interactions in clostridial PERFRINGOLYSIN O myonecrosis. Perfringolysin O (θ-toxin) is a pore-forming hemolytic toxin that belongs to the cholesterol-dependent cyto- Effect of α-Toxin on Erythropoiesis lysin (CDC) family (16, 101), which includes streptoly- α-Toxin has been reported to lyse erythrocytes from sin O from pyogenes, listeriolysin O from various species. It activates T-type Ca2+ channels and in- L. monocytogenes, pneumolysin from Streptococcus creases intracellular Ca2+, which plays an important role pneumoniae (17), and related cytolysins produced in the hemolysis of horse erythrocytes (96). The toxin by several other clostridia (102). The perfringolysin O activates the sphingomyelin metabolic system, leading structural gene, pfoA, is located on the chromosome to the hemolysis of sheep erythrocytes (56, 57), and the (103, 104). Perfringolysin O is elaborated by most, but lysis of rabbit erythrocytes involves α-toxin-mediated not all, strains of C. perfringens (104–107), and mature activation of endogenous phospholipase C activity (54, perfringolysin O (53 kDa, 473 aa) (108) is secreted 55). Recently, lethal toxin was re- as a water-soluble monomer and binds to host cell ported not only to induce hemolysis, but also to inhibit membranes in a cholesterol-dependent manner (102). erythroid differentiation of cord blood-derived CD34+ Binding to the cell membrane initiates oligomerization hematopoietic stem cells, leading to the suppression and the formation of a pore complex that can contain of erythropoiesis (97). The toxin-induced suppression up to 50 monomeric subunits and is responsible for of erythropoiesis is potentially part of a lethal toxin- cell lysis. mediated pathophysiology in anemia and hypoxia. Sim- Perfringolysin O monomers have a hydrophilic, ilarly, it was reported that mature erythroblasts were β-sheet-rich structure that can be divided into four dis- preferentially decreased when isolated mouse bone tinct domains (109, 110)(Fig. 6). Domain 1 contains a marrow cells were treated with α-toxin, and subsequent seven-stranded antiparallel β-sheet flanked by helices experiments revealed that the blockage of erythroid and loops. Domain 2 forms an elongated β-sheet struc- differentiation was involved in the reduction in mature ture. Domain 3 comprises α-helices and β-sheets, and erythroblasts by α-toxin (98). Massive intravascular domain 4 (the C-terminal region) is divided into two hemolysis and severe anemia have been reported in distinct β-sandwich regions. Domain 4 also contains a C. perfringens-infected patients (99), although it is a highly conserved 11-aa Trp-rich motif (ECTGLAWEW relatively uncommon event. The detailed mechanisms WR) or undecapeptide loop that is near the C-terminus. by which C. perfringens infection causes severe anemia This structure subsequently was shown to be similar to have not been elucidated, but blockage of erythropoiesis that of members of the eukaryotic membrane attack by α-toxin may be important (Fig. 5). complex/perforin family (111). Experiments carried out using an inactive α-toxin Perfringolysin O and most of the other cholesterol- H148G variant revealed that its enzymatic activity was dependent cytolysins bind directly to cholesterol in the required for the α-toxin-mediated reduction of erythro- host cell membrane. Contrary to what was thought blast levels (98). Furthermore, the cell surface expression for many years, the undecapeptide is not the receptor- of a lipid raft marker, GM1, decreased in association binding motif (112). Instead, there is a Thr-Ile pair with erythroid differentiation, and α-toxin affected lipid located in a nearby loop (L1) of domain 4 that is re-

ASMscience.org/MicrobiolSpectrum 9 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

PMNL and macrophages in the blood vessels around the site of infection, which leads to vascular leukostasis and contributes to the paucity of leukocyte infiltration to the site of infection (20, 21). Similarly, intramuscular injection of a crude C. perfringens toxin preparation causes and promotes the formation of intravascular platelet aggregates (87). Leukocytes and fibrin accumu- late rapidly in the aggregate, leading to obstruction of the blood vessels and the production of thrombi that reduce local blood flow (87, 88). Perfringolysin O upregulates the expression of ad- herence molecules, including platelet-activating factor, CD11b, and endothelial adherence molecules (118– 120). In addition, it stimulates the expression of inter- cellular adherence molecule 1 and platelet-activating factor in endothelial cells (19, 118, 119, 121). The toxin triggers platelet/leukocyte and platelet/platelet aggrega- tion by activating the platelet fibrinogen receptor gpIIb/ IIIa synergistically with α-toxin (86, 122). Moreover, it has been reported to impair the polymerization of actin filaments in leukocytes and the migration of neutrophils FIGURE 6 Structure of perfringolysin O. The structure of per- in response to chemoattractants (118, 123). These events fringolysin O was obtained from the Protein Data Bank (code 1PFO) and was created using MacPyMOL. A ribbon repre- lead to the formation of intravascular platelet/leucocyte/ fi sentation is shown, colored from blue (N-terminal) to red brin aggregates, and the adherence of the aggregates (C-terminal). to the vascular endothelial cells results in vascular in- jury, vessel obstruction, hypoxia, and tissue destruction (1, 30, 124). In addition, a direct cytotoxic effect of sponsible for binding to cholesterol (113). The role of perfringolysin O on endothelial cells has been demon- the undecapeptide is to assist in anchoring the toxin strated, which also impairs leukocyte transmigration monomer to the membrane and to couple membrane (30, 88). Disruption of the endothelium contributes to insertion with oligomerization (112). The next step in progressive (30). α-Toxin impairs granulopoiesis, the oligomerization process is a conformational change leading to the reduction of neutrophils, and a direct of two α-helical bundles in domain 3 to form two longer cytolytic effect of perfringolysin O on leukocytes has amphipathic β-hairpins; this extended β-hairpin struc- been reported (123). Therefore, perfringolysin O works ture inserts into the host cell membrane, initially to form synergistically with α-toxin to reduce the level of leuko- toxin dimers and then to form a large oligomeric β-barrel cytes. Finally, C. perfringens cells can escape from the pore that contains 35 to 45 monomers. The biological phagosome of macrophages and survive, and perfringo- effect of large pore formation is to induce cell lysis by lysin O is an important factor in this escape process osmotic stress (114). In addition, perfringolysin O can (125, 126). These data are in agreement with genetic interfere with cell signaling events by activating TLR4 studies that showed that α-toxin and perfringolysin O and inducing inducible nitric oxide synthase expres- act synergistically in the disease process (19, 20). sion and tumor necrosis factor-α (TNFα) secretion in In the latter stage of clostridial gas gangrene, cardio- bone marrow-derived macrophages (115) and by stim- vascular collapse, tachycardia, and multiorgan failure ulating the degradation of Ubc9, thereby reducing the are observed in patients (1). Toxins such as perfringolysin posttranslational SUMOylation of host proteins (116). Oandα-toxin are released into the blood and can act at Therefore, perfringolysin O has effects on both mem- a distance from the site of infection. Perfringolysin O re- brane integrity and internal cell signaling. duces systemic vascular resistance, increases cardiac out- Perfringolysin O works synergistically with α-toxin to put, and decreases the heart rate without a drop in mean cause gas gangrene by promoting leukostasis and in- arterial pressure (117, 127). In addition, it promotes the travascular coagulation (19–21, 117). The result is de- release of inflammatory cytokines, such as TNFα,IL-1, creased blood flow (87, 88) and the accumulation of IL-6, platelet-activating factor, and prostaglandin I2, by

10 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections acting synergistically with α-toxin, which contributes to formation (23, 102, 133). The resultant cleaved pro- the symptoms of toxic shock, including hypotension, peptide prevents premature oligomerization (134). The hypoxia, and reduced cardiac output (30, 128). 46-kDa prototoxin binds to glycophosphatidylinositol- anchored protein receptors located in lipid rafts in the host cell membrane (135). After protease cleavage on FUNCTION AND HOST CELL INTERACTIONS the cell membrane, the 43-kDa mature toxin undergoes OF C. SEPTICUM α-TOXIN a conformational change and oligomerizes to form a C. septicum is the primary causative agent of nontrau- heptameric prepore complex that is required for pore matic clostridial myonecrosis and enumerates an α-toxin formation (24). Cleavage of the C-terminal extension en- that is essential for disease and is responsible for reduced ables the formation of a transmembrane β-barrel struc- PMNL influx into the infected lesions (25). C. septicum ture, with each monomer contributing one extended α-toxin is a member of the aerolysin family of pore- β-hairpin loop to the β-barrel that forms the trans- forming toxins (129, 130). These toxins include aero- membrane pore (130, 132). lysin from , ε-toxin and C. per- The analysis of substitution derivatives of α-toxin fringens enterotoxin (CPE) from C. perfringens, and has yielded insights into the functional domains of parasporin toxins produced by Bacillus thuringiensis. the toxin (132, 136–138). Y155F and S189C α-toxin These toxins all form small, often heptameric, β-barrel derivatives had reduced toxicity for CHO cells but pores in host cell membranes (130). Although the crystal were still cleaved by host proteases and could bind to structures of several members of the aerolysin family glycophosphatidylinositol-anchored receptors and form have been determined, there is no such structure available oligomers. A S189C/S238C derivative retained its ability for α-toxin. However, molecular modelling of α-toxin to bind, act as a substrate for furin cleavage, and form shows that it does not have the small-lobe lectin-binding oligomers but was not toxic (136). Like aerolysin, domain (D1) of aerolysin, known as the aerolysin per- α-toxin has a Trp-rich motif (amino acids 302 to 312) tussis toxin domain (131), but has homologous domains in its C-terminal domain (138). Several of the residues to the other three aerolysin domains (Fig. 7)(132). in this motif, including three Trp residues, are essential α-Toxin is synthesized as a preprototoxin, with a 31- for binding and/or cytotoxic activity (137, 138). Fluo- aa leader sequence that is cleaved upon secretion from rimetric studies using Cys-substituted derivatives led to the bacterial cell and a 45-aa C-terminal extension that is the identification of the amphipathic transmembrane cleaved by host proteases such as furin prior to pore β-hairpin (Lys-203 to Gln-232) that spans the host cell membrane. Deletion of this region eliminated the toxin’s FIGURE 7 Crystal structure of aerolysin and molecular model ability to form pores and kill cells, but the resultant of C. septicum α-toxin. Domains are indicated. Note that protein could still bind, undergo cleavage, and form domain 1 (D1) of aerolysin is not present in α-toxin. AT, oligomers (132). Subsequently, saturation Ala or Cys C. septicum α-toxin. Reprinted with permission from refer- substitutions were made across the remaining residues in ence 137. Copyright (2006) American Chemical Society. the α-toxin molecule, and the resultant proteins were purified and analyzed for their biological and biophysi- cal properties (137). The results showed that the only α-toxin Cys residue, Cys-86 in domain 1, was required for receptor binding. In summary, derivatives altered in receptor binding all had substitutions in domain 1, but proteins with altered oligomerization properties had changes in domains 1 and 3. Several of these mutants have been examined in the mouse myonecrosis model (139). The results showed that both the amphipathic β-hairpin and Cys-86 were essential for disease, pro- viding evidence that receptor binding and pore forma- tion are important in pathogenesis. Pore formation is essential for C. septicum-mediated myonecrosis (129, 140), but pore-mediated osmotic cell lysis is not the only consequence of intoxication by α-toxin. Treatment of C2C12 mouse myoblast cells

ASMscience.org/MicrobiolSpectrum 11 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al. with α-toxin leads to a rapid Ca2+ influx into the cell but demonstrating that toxin-induced TNF-α release is re- does not induce apoptosis. Instead, α-toxin induces pro- lated to stimulation of the ERK/MAPK signaling path- grammed cellular necrosis (140), which is consistent way through TrkA. Anti-TNF-α antibodies, but not with the fact that it is more active against nucleated cells anti-IL-1β or anti-IL-6 antibodies, blocked the death (141, 142). α-Toxin-mediated Ca2+ influx resulted in the of mice and intravascular hemolysis by α-toxin. Com- activation of proteolytic calpain activity, a reduction in pounds that block the release of TNF-α therefore may lysosomal integrity, leading to cathepsin release, mito- be valuable in treating patients with C. perfringens in- chondrial dysfunction and ATP depletion, and release fection. Recently, experiments with C. perfringens in- from the nucleus of the histone-binding protein HMGB1 fections in mice provided evidence that treatment with (140). Induction of this programmed necrotic pathway opioids such as buprenorphine stop the development was dependent on the ability to form pores; treatment of disease (148), but this treatment option has not been with a substituted α-toxin lacking the β-hairpin loop did tested in clinical settings. not induce a Ca2+ influx or programmed necrosis. All of Several studies have investigated the use of recombi- these changes are consistent with α-toxin-mediated pore nant α-toxin variants to vaccinate against C. perfringens- formation causing the induction of the programmed cel- mediated gas gangrene, primarily from a military or lular necrosis pathway (140). Other studies have shown biosecurity perspective or for elderly patients that are that α-toxin treatment of Madin-Darby Canine Kidney at risk. Initial studies showed that the isolated nontoxic (MCDK) cells causes a rapid K+ efflux, ATP depletion, C-terminal C2 lipid-binding domain of α-toxin (amino and caspase-3-independent necrosis and nonapoptotic acids 247 to 370) was highly immunoprotective in mice cell death (129). Subsequently, it was shown that pore (14). Mice immunized with this formaldehyde-treated formation by α-toxin led to ERK, c-Jun N-terminal pro- recombinant antigen were protected from the effects tein kinase, and p38 phosphorylation and activation of of α-toxin and from challenge with a 10× LD50 dose of their resultant MAPK pathways and the resultant dose- C. perfringens. By contrast, mice immunized with the dependent release of TNFα (143). C. septicum α-toxin equivalent N-terminal enzymatic domain (amino acids 1 also leads to a significant reduction in vascular perfusion to 249) produced immunoreactive antibodies, but these of muscle tissues, as shown by comparative intravital antibodies did not protect against the toxin or disease. microscopy studies using culture supernatants from an These results were confirmed in a subsequent, more de- isogenic series of C. septicum strains (88). tailed, vaccination study that also showed that mice vaccinated with the C-terminal domain and then chal- lenged in a murine infection model had reduced throm- CONTROL AND TREATMENT OF bosis and an increased PMNL influx into the site of HISTOTOXIC CLOSTRIDIAL INFECTIONS infection (149). To further define the immunoprotective The treatment of myonecrotic infections caused by region of the C-terminal domain, the potency of several C. perfringens and C. septicum is complicated by the recombinant fragments (amino acids 251 to 370, 281 to rapid onset and progression of disease. Death can result 370, and 311 to 370) fused to glutathione S-transferase within 24 hours of initial symptoms. Standard treatment (GST) were examined in an active vaccination model regimens involve surgical of the infected (150). Vaccination with the 251 to 370 and 281 to 370 tissues, which needs to occur as early as possible (2), GST-fusions protected against the toxicity of α-toxin and treatment with and (144). and against C. perfringens challenge, whereas the amino The effectiveness of the penicillin on its own has been acid 311 to 370 derivative yielded only partial protec- questioned (145), and there is experimental evidence tion. Despite these three promising studies, no commer- that clindamycin and metronidazole, alone or in com- cial C. perfringens vaccine is currently available. bination with penicillin are more effective in mice (145, 146). Often, surgical removal of an infected limb is ACKNOWLEDGMENTS the only option available in life-threatening infections. Research in Julian Rood’s laboratory was supported by project In other studies, mice lacking TNF-α were protected grant GNT1082401 from the Australian National Health and from C. perfringens α-toxin-induced lethality (147), Medical Research Council. demonstrating that TNF-α release is essential for the lethal effect of the toxin. Treatment with erythromycin REFERENCES α α 1. Stevens DL, Aldape MJ, Bryant AE. 2012. Life-threatening clostridial blocked -toxin-induced release of TNF- from periph- infections. Anaerobe 18:254–259 http://dx.doi.org/10.1016/j.anaerobe eral neutrophils and the activation of TrkA and ERK1/2, .2011.11.001.

12 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections

2. Stevens DL, Bryant AE. 2017. Necrotizing soft-tissue infections. N Engl perfringens-mediated gas gangrene. Infect Immun 69:7904–7910 http:// J Med 377:2253–2265 http://dx.doi.org/10.1056/NEJMra1600673. dx.doi.org/10.1128/IAI.69.12.7904-7910.2001. 3. Stevens DL, Musher DM, Watson DA, Eddy H, Hamill RJ, Gyorkey F, 21. Ellemor DM, Baird RN, Awad MM, Boyd RL, Rood JI, Emmins JJ. Rosen H, Mader J. 1990. Spontaneous, nontraumatic gangrene due to 1999. Use of genetically manipulated strains of Clostridium perfringens Clostridium septicum. Rev Infect Dis 12:286–296 http://dx.doi.org reveals that both alpha-toxin and theta-toxin are required for vascular /10.1093/clinids/12.2.286. leukostasis to occur in experimental gas gangrene. Infect Immun 67:4902– 4. Srivastava I, Aldape MJ, Bryant AE, Stevens DL. 2017. Spontaneous 4907. C. septicum gas gangrene: a literature review. Anaerobe 48:165–171 22. Low L-Y, Harrison PF, Gould J, Powell DR, Choo JM, Forster http://dx.doi.org/10.1016/j.anaerobe.2017.07.008. SC, Chapman R, Gearing LJ, Cheung JK, Hertzog P, Rood JI. 2018. 5. Kimura AC, Higa JI, Levin RM, Simpson G, Vargas Y, Vugia DJ. 2004. Concurrent host-pathogen transcriptional responses in a Clostridium Outbreak of due to Clostridium sordellii among black- perfringens murine myonecrosis infection. MBio 9:e00473-18 http://dx tar heroin users. Clin Infect Dis 38:e87–e91 http://dx.doi.org/10.1086 .doi.org/10.1128/mBio.00473-18. /383471. 23. Gordon VM, Benz R, Fujii K, Leppla SH, Tweten RK. 1997. Clos- 6. Assadian O, Assadian A, Senekowitsch C, Makristathis A, Hagmüller tridium septicum alpha-toxin is proteolytically activated by furin. Infect G. 2004. Gas gangrene due to Clostridium perfringens in two injecting Immun 65:4130–4134. drug users in Vienna, Austria. Wien Klin Wochenschr 116:264–267 http:// 24. Sellman BR, Kagan BL, Tweten RK. 1997. Generation of a dx.doi.org/10.1007/BF03041058. membrane-bound, oligomerized pre-pore complex is necessary for pore 7. Gonzales y Tucker RD, Frazee B. 2014. View from the front lines: an formation by Clostridium septicum alpha toxin. Mol Microbiol 23: emergency medicine perspective on clostridial infections in injection drug 551–558 http://dx.doi.org/10.1046/j.1365-2958.1997.d01-1876.x. users. Anaerobe 30:108–115 http://dx.doi.org/10.1016/j.anaerobe.2014 25. Kennedy CL, Krejany EO, Young LF, O’Connor JR, Awad MM, .09.005. Boyd RL, Emmins JJ, Lyras D, Rood JI. 2005. The alpha-toxin of 8. Stevens DL, Rood JI. 2006. Histotoxic clostridia, p 715–725. Clostridium septicum is essential for virulence. Mol Microbiol 57: In Fischetti VA, Novick RP, Ferretti JJ, Portnoy DA, Rood JI (ed), Gram- 1357–1366 http://dx.doi.org/10.1111/j.1365-2958.2005.04774.x. Positive Pathogens, 2nd ed. ASM Press, Washington, DC. http://dx.doi 26. Titball RW. 1993. Bacterial phospholipases C. Microbiol Rev 57: .org/10.1128/9781555816513.ch58 347–366. 9. Ohtani K. 2016. Gene regulation by the VirS/VirR system in Clostrid- 27. Titball RW, Naylor CE, Basak AK. 1999. The Clostridium per- ium perfringens. Anaerobe 41:5–9 http://dx.doi.org/10.1016/j.anaerobe fringens α-toxin. Anaerobe 5:51–64 http://dx.doi.org/10.1006/anae.1999 .2016.06.003. .0191. 10. Ohtani K, Yuan Y, Hassan S, Wang R, Wang Y, Shimizu T. 2009. 28. Sakurai J, Nagahama M, Oda M. 2004. Clostridium perfringens Virulence gene regulation by the agr system in Clostridium perfringens. alpha-toxin: characterization and mode of action. J Biochem 136:569– J Bacteriol 191:3919–3927 http://dx.doi.org/10.1128/JB.01455-08. 574 http://dx.doi.org/10.1093/jb/mvh161. 11. Ma M, Li J, McClane BA. 2015. Structure-function analysis of peptide signaling in the Clostridium perfringens Agr-like quorum sensing system. 29. Uzal FA, Freedman JC, Shrestha A, Theoret JR, Garcia J, Awad MM, J Bacteriol 197:1807–1818 http://dx.doi.org/10.1128/JB.02614-14. Adams V, Moore RJ, Rood JI, McClane BA. 2014. Towards an under- – standing of the role of Clostridium perfringens toxins in human and 12. Titball R, Rood J. 2000. Bacterial phospholipases, p 529 556. animal disease. Future Microbiol 9:361–377 http://dx.doi.org/10.2217/fmb In Aktories K, Just I (ed), Bacterial Protein Toxins. Springer, Berlin, .13.168. Germany. http://dx.doi.org/10.1007/978-3-662-05971-5_23 30. Stevens DL, Bryant AE. 2002. The role of clostridial toxins in the 13. Macfarlane MG, Knight BCJG. 1941. The biochemistry of bacterial pathogenesis of gas gangrene. Clin Infect Dis 35(Suppl 1):S93–S100 http:// toxins: the lecithinase activity of Cl. welchii toxins. Biochem J 35:884–902 dx.doi.org/10.1086/341928. http://dx.doi.org/10.1042/bj0350884. 14. Williamson ED, Titball RW. 1993. A genetically engineered vaccine 31. Naylor CE, Eaton JT, Howells A, Justin N, Moss DS, Titball RW, against the alpha-toxin of Clostridium perfringens protects mice against Basak AK. 1998. Structure of the key toxin in gas gangrene. Nat Struct – experimental gas gangrene. Vaccine 11:1253–1258 http://dx.doi.org Biol 5:738 746 http://dx.doi.org/10.1038/1447. /10.1016/0264-410X(93)90051-X. 32. Gilmore MS, Cruz-Rodz AL, Leimeister-Wächter M, Kreft J, Goebel 15. Awad MM, Bryant AE, Stevens DL, Rood JI. 1995. Virulence studies W. 1989. A cytolytic determinant, cereolysin AB, which on chromosomal α-toxin and θ-toxin mutants constructed by allelic ex- comprises the phospholipase C and sphingomyelinase genes: nucleotide – change provide genetic evidence for the essential role of α-toxin in Clos- sequence and genetic linkage. J Bacteriol 171:744 753 http://dx.doi.org tridium perfringens-mediated gas gangrene. Mol Microbiol 15:191–202 /10.1128/jb.171.2.744-753.1989. http://dx.doi.org/10.1111/j.1365-2958.1995.tb02234.x. 33. Tso JY, Siebel C. 1989. Cloning and expression of the phospholipase 16. Tweten RK. 2005. Cholesterol-dependent cytolysins, a family of C gene from Clostridium perfringens and Clostridium bifermentans. Infect versatile pore-forming toxins. Infect Immun 73:6199–6209 http://dx.doi Immun 57:468–476. .org/10.1128/IAI.73.10.6199-6209.2005. 34. Vazquez-Boland J-A, Kocks C, Dramsi S, Ohayon H, Geoffroy C, 17. Gilbert RJ. 2010. Cholesterol-dependent cytolysins. Adv Exp Med Mengaud J, Cossart P. 1992. Nucleotide sequence of the lecithinase Biol 677:56–66 http://dx.doi.org/10.1007/978-1-4419-6327-7_5. operon of and possible role of lecithinase in cell- – 18. Awad MM, Rood JI. 1997. Isolation of α-toxin, θ-toxin and κ-toxin to-cell spread. Infect Immun 60:219 230. mutants of Clostridium perfringens by Tn916 mutagenesis. Microb 35. Hough E, Hansen LK, Birknes B, Jynge K, Hansen S, Hordvik A, Pathog 22:275–284 http://dx.doi.org/10.1006/mpat.1996.0115. Little C, Dodson E, Derewenda Z. 1989. High-resolution (1.5 A) crystal 19. Stevens DL, Tweten RK, Awad MM, Rood JI, Bryant AE. 1997. structure of phospholipase C from Bacillus cereus. Nature 338:357–360 Clostridial gas gangrene: evidence that α and θ toxins differentially http://dx.doi.org/10.1038/338357a0. modulate the immune response and induce acute tissue necrosis. J Infect 36. Nagahama M, Mukai M, Morimitsu S, Ochi S, Sakurai J. 2002. Role Dis 176:189–195 http://dx.doi.org/10.1086/514022. of the C-domain in the biological activities of Clostridium perfringens 20. Awad MM, Ellemor DM, Boyd RL, Emmins JJ, Rood JI. 2001. alpha-toxin. Microbiol Immunol 46:647–655 http://dx.doi.org/10.1111 Synergistic effects of α-toxin and perfringolysin O in Clostridium /j.1348-0421.2002.tb02748.x.

ASMscience.org/MicrobiolSpectrum 13 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

37. Guillouard I, Alzari PM, Saliou B, Cole ST. 1997. The carboxy- 52. Oda M, Ikari S, Matsuno T, Morimune Y, Nagahama M, Sakurai J. terminal C2-like domain of the alpha-toxin from Clostridium perfringens 2006. Signal transduction mechanism involved in Clostridium perfringens mediates calcium-dependent membrane recognition. Mol Microbiol 26: alpha-toxin-induced superoxide anion generation in rabbit neutrophils. 867–876 http://dx.doi.org/10.1046/j.1365-2958.1997.6161993.x. Infect Immun 74:2876–2886 http://dx.doi.org/10.1128/IAI.74.5.2876 38. Naylor CE, Jepson M, Crane DT, Titball RW, Miller J, Basak AK, -2886.2006. Bolgiano B. 1999. Characterisation of the calcium-binding C-terminal 53. Oda M, Kabura M, Takagishi T, Suzue A, Tominaga K, Urano S, domain of Clostridium perfringens alpha-toxin. J Mol Biol 294:757–770 Nagahama M, Kobayashi K, Furukawa K, Furukawa K, Sakurai J. 2012. http://dx.doi.org/10.1006/jmbi.1999.3279. Clostridium perfringens alpha-toxin recognizes the GM1a-TrkA com- – 39. Nagahama M, Nakayama T, Kobayashi K, Sakurai J. 1995. Site- plex. J Biol Chem 287:33070 33079 http://dx.doi.org/10.1074/jbc.M112 directed mutagenesis studies on the zinc-binding domain of Clostridium .393801. perfringens alpha toxin. Jpn J Med Sci Biol 48:265–266. 54. Sakurai J, Ochi S, Tanaka H. 1993. Evidence for coupling of Clos- 40. Nagahama M, Michiue K, Sakurai J. 1996. Membrane-damaging tridium perfringens alpha-toxin-induced hemolysis to stimulated phospha- – action of Clostridium perfringens alpha-toxin on phospholipid liposomes. tidic acid formation in rabbit erythrocytes. Infect Immun 61:3711 3718. Biochim Biophys Acta 1280:120–126. 55. Sakurai J, Ochi S, Tanaka H. 1994. Regulation of Clostridium 41. Vachieri SG, Clark GC, Alape-Girón A, Flores-Díaz M, Justin N, perfringens alpha-toxin-activated phospholipase C in rabbit erythrocyte – Naylor CE, Titball RW, Basak AK. 2010. Comparison of a nontoxic membranes. Infect Immun 62:717 721. variant of Clostridium perfringens α-toxin with the toxic wild-type strain. 56. Ochi S, Oda M, Matsuda H, Ikari S, Sakurai J. 2004. Clostridium Acta Crystallogr D Biol Crystallogr 66:1067–1074 http://dx.doi.org/10 perfringens alpha-toxin activates the sphingomyelin metabolism system .1107/S090744491003369X. in sheep erythrocytes. J Biol Chem 279:12181–12189 http://dx.doi.org 42. Eaton JT, Naylor CE, Howells AM, Moss DS, Titball RW, Basak AK. /10.1074/jbc.M307046200. 2002. Crystal structure of the C. perfringens alpha-toxin with the active 57. Oda M, Matsuno T, Shiihara R, Ochi S, Yamauchi R, Saito Y, site closed by a flexible loop region. J Mol Biol 319:275–281 http://dx.doi Imagawa H, Nagahama M, Nishizawa M, Sakurai J. 2008. The rela- .org/10.1016/S0022-2836(02)00290-5. tionship between the metabolism of sphingomyelin species and the he- 43. Clark GC, Briggs DC, Karasawa T, Wang X, Cole AR, Maegawa T, molysis of sheep erythrocytes induced by Clostridium perfringens alpha- Jayasekera PN, Naylor CE, Miller J, Moss DS, Nakamura S, Basak AK, toxin. J Lipid Res 49:1039–1047 http://dx.doi.org/10.1194/jlr.M700587 Titball RW. 2003. Clostridium absonum alpha-toxin: new insights into -JLR200. clostridial phospholipase C substrate binding and specificity. J Mol Biol 58. Tsukamoto K, Kozai Y, Ihara H, Kohda T, Mukamoto M, Tsuji T, 333:759–769 http://dx.doi.org/10.1016/j.jmb.2003.07.016. Kozaki S. 2008. Identification of the receptor-binding sites in the carboxyl- 44. Nagahama M, Otsuka A, Sakurai J. 2006. Role of tyrosine-57 and -65 terminal half of the heavy chain of botulinum neurotoxin types C and D. in membrane-damaging and sphingomyelinase activities of Clostridium Microb Pathog 44:484–493 http://dx.doi.org/10.1016/j.micpath.2007.12 perfringens alpha-toxin. Biochim Biophys Acta 1762:110–114 http://dx .003. .doi.org/10.1016/j.bbadis.2005.10.002. 59. Oda M, Shiihara R, Ohmae Y, Kabura M, Takagishi T, Kobayashi K, 45. Alape-Girón A, Flores-Díaz M, Guillouard I, Naylor CE, Titball Nagahama M, Inoue M, Abe T, Setsu K, Sakurai J. 2012. Clostridium RW, Rucavado A, Lomonte B, Basak AK, Gutiérrez JM, Cole ST, perfringens alpha-toxin induces the release of IL-8 through a dual path- Thelestam M. 2000. Identification of residues critical for toxicity in way via TrkA in A549 cells. Biochim Biophys Acta 1822:1581–1589 Clostridium perfringens phospholipase C, the key toxin in gas gangrene. http://dx.doi.org/10.1016/j.bbadis.2012.06.007. – Eur J Biochem 267:5191 5197 http://dx.doi.org/10.1046/j.1432-1327 60. Yamazaki Y, Horibata Y, Nagatsuka Y, Hirabayashi Y, Hashikawa .2000.01588.x. T. 2007. Fucoganglioside alpha-fucosyl(alpha-galactosyl)-GM1: a 46. Urbina P, Flores-Díaz M, Alape-Girón A, Alonso A, Goñi FM. 2011. novel member of lipid membrane microdomain components involved in Effects of bilayer composition and physical properties on the phospholi- PC12 cell neuritogenesis. Biochem J 407:31–40 http://dx.doi.org/10.1042 pase C and sphingomyelinase activities of Clostridium perfringens α-toxin. /BJ20070090. – Biochim Biophys Acta 1808:279 286 http://dx.doi.org/10.1016/j.bbamem 61. Ichikawa N, Iwabuchi K, Kurihara H, Ishii K, Kobayashi T, Sasaki T, .2010.08.011. Hattori N, Mizuno Y, Hozumi K, Yamada Y, Arikawa-Hirasawa E. 2009. 47. Nagahama M, Otsuka A, Oda M, Singh RK, Ziora ZM, Imagawa H, Binding of laminin-1 to monosialoganglioside GM1 in lipid rafts is crucial Nishizawa M, Sakurai J. 2007. Effect of unsaturated bonds in the sn-2 for neurite outgrowth. J Cell Sci 122:289–299 http://dx.doi.org/10.1242 acyl chain of phosphatidylcholine on the membrane-damaging action /jcs.030338. of Clostridium perfringens alpha-toxin toward liposomes. Biochim Bio- 62. Takagishi T, Oda M, Kabura M, Kurosawa M, Tominaga K, Urano – phys Acta 1768:2940 2945 http://dx.doi.org/10.1016/j.bbamem.2007 S, Ueda Y, Kobayashi K, Kobayashi T, Sakurai J, Terao Y, Nagahama M. .08.016. 2015. Clostridium perfringens alpha-toxin induces Gm1a clustering and 48. Moe PC, Heuck AP. 2010. Phospholipid hydrolysis caused by Clos- Trka phosphorylation in the host cell membrane. PLoS One 10:e0120497 tridium perfringens α-toxin facilitates the targeting of perfringolysin O to http://dx.doi.org/10.1371/journal.pone.0120497. – membrane bilayers. Biochemistry 49:9498 9507 http://dx.doi.org/10.1021 63. Ueda Y, Makino A, Murase-Tamada K, Sakai S, Inaba T, Hullin- /bi1013886. Matsuda F, Kobayashi T. 2013. Sphingomyelin regulates the transbilayer 49. Fujii Y, Sakurai J. 1989. Contraction of the rat isolated aorta caused movement of diacylglycerol in the plasma membrane of Madin-Darby by Clostridium perfringens alpha toxin (phospholipase C): evidence for canine kidney cells. FASEB J 27:3284–3297 http://dx.doi.org/10.1096 the involvement of arachidonic acid metabolism. Br J Pharmacol 97: /fj.12-226548. – 119 124 http://dx.doi.org/10.1111/j.1476-5381.1989.tb11931.x. 64. Oda M, Terao Y, Sakurai J, Nagahama M. 2015. Membrane-binding 50. Sakurai J, Fujii Y, Shirotani M. 1990. Contraction induced by Clos- mechanism of Clostridium perfringens alpha-toxin. Toxins (Basel) 7: tridium perfringens alpha toxin in the isolated rat ileum. Toxicon 28:411– 5268–5275 http://dx.doi.org/10.3390/toxins7124880. 418 http://dx.doi.org/10.1016/0041-0101(90)90079-M. 65. Flores-Díaz M, Alape-Girón A, Clark G, Catimel B, Hirabayashi Y, 51. Ochi S, Miyawaki T, Matsuda H, Oda M, Nagahama M, Sakurai J. Nice E, Gutiérrez JM, Titball R, Thelestam M. 2005. A cellular deficiency 2002. Clostridium perfringens alpha-toxin induces rabbit neutrophil ad- of gangliosides causes hypersensitivity to Clostridium perfringens phos- hesion. Microbiology 148:237–245 http://dx.doi.org/10.1099/00221287 pholipase C. J Biol Chem 280:26680–26689 http://dx.doi.org/10.1074 -148-1-237. /jbc.M500278200.

14 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections

66. Monturiol-Gross L, Flores-Díaz M, Pineda-Padilla MJ, Castro-Castro and Clostridium perfringens. Clin Microbiol Rev 16:451–462 http://dx AC, Alape-Giron A. 2014. Clostridium perfringens phospholipase C in- .doi.org/10.1128/CMR.16.3.451-462.2003. duced ROS production and cytotoxicity require PKC, MEK1 and NFκB 85. McNee JW, Dunn JS. 1917. The method of spread of gas gangrene activation. PLoS One 9:e86475 http://dx.doi.org/10.1371/journal.pone into living tissue. BMJ 1:4–729 http://dx.doi.org/10.1136/bmj.1.2944 .0086475. .726. 67. Flores-Díaz M, Monturiol-Gross L, Naylor C, Alape-Girón A, Flieger 86. Bryant AE, Chen RY, Nagata Y, Wang Y, Lee CH, Finegold S, A. 2016. Bacterial sphingomyelinases and phospholipases as virulence Guth PH, Stevens DL. 2000. Clostridial gas gangrene. II. Phospholipase factors. Microbiol Mol Biol Rev 80:597–628 http://dx.doi.org/10.1128 C-induced activation of platelet gpIIbIIIa mediates vascular occlusion /MMBR.00082-15. and myonecrosis in Clostridium perfringens gas gangrene. J Infect Dis 68. Andrews NW, Almeida PE, Corrotte M. 2014. Damage control: cel- 182:808–815 http://dx.doi.org/10.1086/315757. lular mechanisms of plasma membrane repair. Trends Cell Biol 24:734– 87. Bryant AE, Chen RY, Nagata Y, Wang Y, Lee CH, Finegold S, 742 http://dx.doi.org/10.1016/j.tcb.2014.07.008. Guth PH, Stevens DL. 2000. Clostridial gas gangrene. I. Cellular and 69. Amulic B, Cazalet C, Hayes GL, Metzler KD, Zychlinsky A. 2012. molecular mechanisms of microvascular dysfunction induced by exo- Neutrophil function: from mechanisms to disease. Annu Rev Immunol toxins of Clostridium perfringens. J Infect Dis 182:799–807 http://dx.doi 30:459–489 http://dx.doi.org/10.1146/annurev-immunol-020711-074942. .org/10.1086/315756. 70. Dale DC, Boxer L, Liles WC. 2008. The phagocytes: neutrophils and 88. Hickey MJ, Kwan RY, Awad MM, Kennedy CL, Young LF, Hall P, monocytes. Blood 112:935–945 http://dx.doi.org/10.1182/blood-2007 Cordner LM, Lyras D, Emmins JJ, Rood JI. 2008. Molecular and cellu- -12-077917. lar basis of microvascular perfusion deficits induced by Clostridium 71. Manz MG, Boettcher S. 2014. Emergency granulopoiesis. Nat Rev perfringens and Clostridium septicum. PLoS Pathog 4:e1000045 http:// Immunol 14:302–314 http://dx.doi.org/10.1038/nri3660. dx.doi.org/10.1371/journal.ppat.1000045. 72. Wirths S, Bugl S, Kopp HG. 2014. Neutrophil homeostasis and its 89. Bryant AE, Bayer CR, Aldape MJ, Wallace RJ, Titball RW, Stevens regulation by danger signaling. Blood 123:3563–3566 http://dx.doi.org DL. 2006. Clostridium perfringens phospholipase C-induced platelet/ /10.1182/blood-2013-11-516260. leukocyte interactions impede neutrophil diapedesis. J Med Microbiol 55:495–504 http://dx.doi.org/10.1099/jmm.0.46390-0. 73. Nauseef WM, Borregaard N. 2014. Neutrophils at work. Nat Immunol 15:602–611 http://dx.doi.org/10.1038/ni.2921. 90. Takehara M, Takagishi T, Seike S, Ohtani K, Kobayashi K, Miyamoto K, Shimizu T, Nagahama M. 2016. Clostridium perfringens 74. Martinon F, Tschopp J. 2005. NLRs join TLRs as innate sensors of alpha-toxin impairs innate immunity via inhibition of neutrophil differ- pathogens. Trends Immunol 26:447–454 http://dx.doi.org/10.1016/j.it entiation. Sci Rep 6:28192 http://dx.doi.org/10.1038/srep28192. .2005.06.004. 91. Basu S, Hodgson G, Katz M, Dunn AR. 2002. Evaluation of role of 75. Hajishengallis G, Lambris JD. 2011. Microbial manipulation of G-CSF in the production, survival, and release of neutrophils from bone receptor crosstalk in innate immunity. Nat Rev Immunol 11:187–200 marrow into circulation. Blood 100:854–861 http://dx.doi.org/10.1182 http://dx.doi.org/10.1038/nri2918. /blood.V100.3.854. 76. Beutler B. 2000. Tlr4: central component of the sole mammalian LPS 92. Diaz O, Mébarek-Azzam S, Benzaria A, Dubois M, Lagarde M, sensor. Curr Opin Immunol 12:20–26 http://dx.doi.org/10.1016/S0952 Némoz G, Prigent AF. 2005. Disruption of lipid rafts stimulates phos- -7915(99)00046-1. pholipase D activity in human lymphocytes: implication in the regulation 77. Texereau J, Chiche JD, Taylor W, Choukroun G, Comba B, Mira JP. of immune function. J Immunol 175:8077–8086 http://dx.doi.org/10.4049 2005. The importance of Toll-like receptor 2 polymorphisms in severe /jimmunol.175.12.8077. – infections. Clin Infect Dis 41(Suppl 7):S408 S415 http://dx.doi.org 93. Ostrom RS, Insel PA. 2004. The evolving role of lipid rafts and caveolae /10.1086/431990. in G protein-coupled receptor signaling: implications for molecular phar- 78. Dziarski R. 2003. Recognition of bacterial peptidoglycan by the innate macology. Br J Pharmacol 143:235–245 http://dx.doi.org/10.1038/sj.bjp immune system. Cell Mol Life Sci 60:1793–1804 http://dx.doi.org .0705930. /10.1007/s00018-003-3019-6. 94. Simons K, Toomre D. 2000. Lipid rafts and signal transduction. Nat 79. Takeuchi O, Hoshino K, Kawai T, Sanjo H, Takada H, Ogawa T, Rev Mol Cell Biol 1:31–39 http://dx.doi.org/10.1038/35036052. Takeda K, Akira S. 1999. Differential roles of TLR2 and TLR4 in rec- 95. Kolesnick RN, Krönke M. 1998. Regulation of ceramide production ognition of Gram-negative and Gram-positive bacterial cell wall com- – – and apoptosis. Annu Rev Physiol 60:643 665 http://dx.doi.org/10.1146 ponents. Immunity 11:443 451 http://dx.doi.org/10.1016/S1074-7613 /annurev.physiol.60.1.643. (00)80119-3. 96. Ochi S, Oda M, Nagahama M, Sakurai J. 2003. Clostridium 80. Demetri GD, Griffin JD. 1991. Granulocyte colony-stimulating factor – perfringens alpha-toxin-induced hemolysis of horse erythrocytes is de- and its receptor. Blood 78:2791 2808. pendent on Ca2+ uptake. Biochim Biophys Acta 1613:79–86 http://dx 81. Boettcher S, Gerosa RC, Radpour R, Bauer J, Ampenberger F, .doi.org/10.1016/S0005-2736(03)00140-8. Heikenwalder M, Kopf M, Manz MG. 2014. Endothelial cells translate 97. Chang HH, Wang TP, Chen PK, Lin YY, Liao CH, Lin TK, pathogen signals into G-CSF-driven emergency granulopoiesis. Blood Chiang YW, Lin WB, Chiang CY, Kau JH, Huang HH, Hsu HL, Liao – 124:1393 1403 http://dx.doi.org/10.1182/blood-2014-04-570762. CY, Sun DS. 2013. Erythropoiesis suppression is associated with 82. Lieschke GJ, Grail D, Hodgson G, Metcalf D, Stanley E, Cheers C, lethal toxin-mediated pathogenic progression. PLoS One 8:e71718 http:// Fowler KJ, Basu S, Zhan YF, Dunn AR. 1994. Mice lacking granulocyte dx.doi.org/10.1371/journal.pone.0071718. colony-stimulating factor have chronic neutropenia, granulocyte and 98. Takagishi T, Takehara M, Seike S, Miyamoto K, Kobayashi K, macrophage progenitor cell deficiency, and impaired neutrophil mobili- Nagahama M. 2017. Clostridium perfringens α-toxin impairs erythro- zation. Blood 84:1737–1746. poiesis by inhibition of erythroid differentiation. Sci Rep 7:5217 http://dx 83. Liu F, Wu HY, Wesselschmidt R, Kornaga T, Link DC. 1996. Im- .doi.org/10.1038/s41598-017-05567-8. paired production and increased apoptosis of neutrophils in granulocyte 99. Simon TG, Bradley J, Jones A, Carino G. 2014. Massive intravascular colony-stimulating factor receptor-deficient mice. Immunity 5:491–501 hemolysis from Clostridium perfringens septicemia: a review. J Intensive http://dx.doi.org/10.1016/S1074-7613(00)80504-X. Care Med 29:327–333 http://dx.doi.org/10.1177/0885066613498043. 84. Bryant AE. 2003. Biology and pathogenesis of and 100. Konstantinidis DG, Pushkaran S, Johnson JF, Cancelas JA, procoagulant activity in invasive infections caused by group A streptococci Manganaris S, Harris CE, Williams DA, Zheng Y, Kalfa TA. 2012. Sig-

ASMscience.org/MicrobiolSpectrum 15 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Nagahama et al.

naling and cytoskeletal requirements in erythroblast enucleation. Blood 114. Harris RW, Sims PJ, Tweten RK. 1991. Evidence that Clostridium 119:6118–6127 http://dx.doi.org/10.1182/blood-2011-09-379263. perfringens theta-toxin induces colloid-osmotic lysis of erythrocytes. In- 101. Verherstraeten S, Goossens E, Valgaeren B, Pardon B, Timbermont fect Immun 59:2499–2501. L, Haesebrouck F, Ducatelle R, Deprez P, Wade KR, Tweten R, Van 115. Park JM, Ng VH, Maeda S, Rest RF, Karin M. 2004. Anthrolysin O Immerseel F. 2015. Perfringolysin O: the underrated Clostridium per- and other Gram-positive cytolysins are toll-like receptor 4 agonists. J Exp fringens toxin? Toxins (Basel) 7:1702–1721 http://dx.doi.org/10.3390 Med 200:1647–1655 http://dx.doi.org/10.1084/jem.20041215. /toxins7051702. 116. Ribet D, Hamon M, Gouin E, Nahori MA, Impens F, Neyret-Kahn 102. Popoff MR. 2014. Clostridial pore-forming toxins: powerful viru- H, Gevaert K, Vandekerckhove J, Dejean A, Cossart P. 2010. Listeria lence factors. Anaerobe 30:220–238 http://dx.doi.org/10.1016/j.anaerobe monocytogenes impairs SUMOylation for efficient infection. Nature .2014.05.014. 464:1192–1195 http://dx.doi.org/10.1038/nature08963. 103. Katayama S, Dupuy B, Garnier T, Cole ST. 1995. Rapid expansion 117. Asmuth DM, Olson RD, Hackett SP, Bryant AE, Tweten RK, of the physical and genetic map of the chromosome of Clostridium Tso JY, Zollman T, Stevens DL. 1995. Effects of Clostridium perfringens perfringens CPN50. J Bacteriol 177:5680–5685 http://dx.doi.org/10.1128 recombinant and crude phospholipase C and theta-toxin on rabbit he- /jb.177.19.5680-5685.1995. modynamic parameters. J Infect Dis 172:1317–1323 http://dx.doi.org 104. Shimizu T, Ohtani K, Hirakawa H, Ohshima K, Yamashita A, Shiba /10.1093/infdis/172.5.1317. T, Ogasawara N, Hattori M, Kuhara S, Hayashi H. 2002. Complete 118. Bryant AE, Bergstrom R, Zimmerman GA, Salyer JL, Hill HR, genome sequence of Clostridium perfringens, an anaerobic flesh-eater. Tweten RK, Sato H, Stevens DL. 1993. Clostridium perfringens inva- Proc Natl Acad Sci USA 99:996–1001 http://dx.doi.org/10.1073/pnas siveness is enhanced by effects of theta toxin upon PMNL structure and .022493799. function: the roles of leukocytotoxicity and expression of CD11/CD18 105. Myers GS, Rasko DA, Cheung JK, Ravel J, Seshadri R, DeBoy RT, adherence glycoprotein. FEMS Immunol Med Microbiol 7:321–336 Ren Q, Varga J, Awad MM, Brinkac LM, Daugherty SC, Haft DH, http://dx.doi.org/10.1111/j.1574-695X.1993.tb00414.x. Dodson RJ, Madupu R, Nelson WC, Rosovitz MJ, Sullivan SA, Khouri H, 119. Bryant AE, Stevens DL. 1996. Phospholipase C and perfringolysin Dimitrov GI, Watkins KL, Mulligan S, Benton J, Radune D, Fisher DJ, O from Clostridium perfringens upregulate endothelial cell-leukocyte Atkins HS, Hiscox T, Jost BH, Billington SJ, Songer JG, McClane BA, adherence molecule 1 and intercellular leukocyte adherence molecule 1 Titball RW, Rood JI, Melville SB, Paulsen IT. 2006. Skewed genomic expression and induce interleukin-8 synthesis in cultured human umbilical variability in strains of the toxigenic bacterial pathogen, Clostridium per- vein endothelial cells. Infect Immun 64:358–362. fringens. Genome Res 16:1031–1040 http://dx.doi.org/10.1101/gr.5238106. 120. Bunting M, Lorant DE, Bryant AE, Zimmerman GA, McIntyre TM, 106. Hassan KA, Elbourne LD, Tetu SG, Melville SB, Rood JI, Paulsen IT. Stevens DL, Prescott SM. 1997. Alpha toxin from Clostridium perfringens 2014. Genomic analyses of Clostridium perfringens isolates from five induces proinflammatory changes in endothelial cells. J Clin Invest 100: toxinotypes. Res Microbiol 166:255–263. 565–574 http://dx.doi.org/10.1172/JCI119566. 107. Kiu R, Caim S, Alexander S, Pachori P, Hall LJ. 2017. Probing 121. Whately R, Zimmerman G, Stevens D, Parker C, McIntyre T, genomic aspects of the multi-host pathogen Clostridium perfringens Prescott S. 1989. The regulation of platelet activating factor synthesis reveals significant pangenome diversity, and a diverse array of virulence in endothelial cells: the role of calcium and protein kinase. J Biol Chem factors. Front Microbiol 8:2485 http://dx.doi.org/10.3389/fmicb.2017 11:6325–6333. .02485. 122. Bryant AE, Bayer CR, Hayes-Schroer SM, Stevens DL. 2003. Acti- 108. Tweten RK. 1988. Nucleotide sequence of the gene for perfringolysin vation of platelet gpIIbIIIa by phospholipase C from Clostridium per- O (theta-toxin) from Clostridium perfringens: significant homology with fringens involves store-operated calcium entry. J Infect Dis 187:408–417 the genes for streptolysin O and pneumolysin. Infect Immun 56:3235– http://dx.doi.org/10.1086/367964. 3240. 123. Stevens DL, Mitten J, Henry C. 1987. Effects of α and θ toxins from 109. Rossjohn J, Feil SC, McKinstry WJ, Tweten RK, Parker MW. 1997. Clostridium perfringens on human polymorphonuclear leukocytes. J In- Structure of a cholesterol-binding, thiol-activated cytolysin and a model fect Dis 156:324–333 http://dx.doi.org/10.1093/infdis/156.2.324. of its membrane form. Cell 89:685–692 http://dx.doi.org/10.1016/S0092 124. Stevens DL. 2000. The pathogenesis of clostridial myonecrosis. Int J -8674(00)80251-2. Med Microbiol 290:497–502 http://dx.doi.org/10.1016/S1438-4221(00) 110. Shatursky O, Heuck AP, Shepard LA, Rossjohn J, Parker MW, 80074-0. Johnson AE, Tweten RK. 1999. The mechanism of membrane insertion 125. O’Brien DK, Melville SB. 2000. The anaerobic pathogen can escape for a cholesterol-dependent cytolysin: a novel paradigm for pore-forming the phagosome of macrophages under aerobic conditions. Cell Microbiol toxins. Cell 99:293–299 http://dx.doi.org/10.1016/S0092-8674(00) 2:505–519 http://dx.doi.org/10.1046/j.1462-5822.2000.00074.x. 81660-8. 126. O’Brien DK, Melville SB. 2004. Effects of Clostridium perfringens 111. Rosado CJ, Buckle AM, Law RH, Butcher RE, Kan WT, Bird CH, alpha-toxin (PLC) and perfringolysin O (PFO) on cytotoxicity to macro- Ung K, Browne KA, Baran K, Bashtannyk-Puhalovich TA, Faux NG, phages, on escape from the phagosomes of macrophages, and on persis- Wong W, Porter CJ, Pike RN, Ellisdon AM, Pearce MC, Bottomley SP, tence of C. perfringens in host tissues. Infect Immun 72:5204–5215 http:// Emsley J, Smith AI, Rossjohn J, Hartland EL, Voskoboinik I, Trapani JA, dx.doi.org/10.1128/IAI.72.9.5204-5215.2004. Bird PI, Dunstone MA, Whisstock JC. 2007. A common fold mediates 127. Stevens DL, Troyer BE, Merrick DT, Mitten JE, Olson RD. 1988. vertebrate defense and bacterial attack. Science 317:1548–1551 http://dx Lethal effects and cardiovascular effects of purified α- and θ-toxins from .doi.org/10.1126/science.1144706. Clostridium perfringens. J Infect Dis 157:272–279 http://dx.doi.org 112. Tweten RK, Hotze EM, Wade KR. 2015. The unique molecular /10.1093/infdis/157.2.272. choreography of giant pore formation by the cholesterol-dependent 128. Stevens DL, Bryant AE. 1997. Pathogenesis of Clostridium per- cytolysins of Gram-positive bacteria. Annu Rev Microbiol 69:323–340 fringens infection: mechanisms and mediators of shock. Clin Infect Dis 25 http://dx.doi.org/10.1146/annurev-micro-091014-104233. (Suppl 2):S160–S164 http://dx.doi.org/10.1086/516249. 113. Farrand AJ, LaChapelle S, Hotze EM, Johnson AE, Tweten RK. 129. Knapp O, Maier E, Mkaddem SB, Benz R, Bens M, Chenal A, Geny 2010. Only two amino acids are essential for cytolytic toxin recognition of B, Vandewalle A, Popoff MR. 2010. Clostridium septicum alpha-toxin cholesterol at the membrane surface. Proc Natl Acad Sci U S A 107:4341– forms pores and induces rapid cell necrosis. Toxicon 55:61–72 http://dx 4346 http://dx.doi.org/10.1073/pnas.0911581107. .doi.org/10.1016/j.toxicon.2009.06.037.

16 ASMscience.org/MicrobiolSpectrum Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections

130. Dal Peraro M, van der Goot FG. 2016. Pore-forming toxins: ancient, 144. Stevens DL, Bisno AL, Chambers HF, Dellinger EP, Goldstein EJ, but never really out of fashion. Nat Rev Microbiol 14:77–92 http://dx.doi Gorbach SL, Hirschmann JV, Kaplan SL, Montoya JG, Wade JC, Infec- .org/10.1038/nrmicro.2015.3. tious Diseases Society of America. 2014. Practice guidelines for the diag- 131. Rossjohn J, Buckley JT, Hazes B, Murzin AG, Read RJ, Parker MW. nosis and management of skin and soft tissue infections: 2014 update by – 1997. Aerolysin and pertussis toxin share a common receptor-binding the Infectious Diseases Society of America. Clin Infect Dis 59:e10 e52 domain. EMBO J 16:3426–3434 http://dx.doi.org/10.1093/emboj/16.12 http://dx.doi.org/10.1093/cid/ciu296. .3426. 145. Stevens DL, Maier KA, Laine BM, Mitten JE. 1987. Comparison 132. Melton JA, Parker MW, Rossjohn J, Buckley JT, Tweten RK. 2004. of clindamycin, rifampin, tetracycline, metronidazole, and penicillin for fi The identification and structure of the membrane-spanning domain of ef cacy in prevention of experimental gas gangrene due to Clostridium – the Clostridium septicum alpha toxin. J Biol Chem 279:14315–14322 perfringens. J Infect Dis 155:220 228 http://dx.doi.org/10.1093/infdis http://dx.doi.org/10.1074/jbc.M313758200. /155.2.220. 133. Ballard J, Sokolov Y, Yuan W-L, Kagan BL, Tweten RK. 1993. 146. Stevens DL, Laine BM, Mitten JE. 1987. Comparison of single Activation and mechanism of Clostridium septicum alpha toxin. Mol and combination antimicrobial agents for prevention of experimental Microbiol 10:627–634 http://dx.doi.org/10.1111/j.1365-2958.1993 gas gangrene caused by Clostridium perfringens. Antimicrob Agents – .tb00934.x. Chemother 31:312 316 http://dx.doi.org/10.1128/AAC.31.2.312. 147. Oda M, Kihara A, Yoshioka H, Saito Y, Watanabe N, Uoo K, 134. Sellman BR, Tweten RK. 1997. The propeptide of Clostridium septicum alpha toxin functions as an intramolecular chaperone and is a Higashihara M, Nagahama M, Koide N, Yokochi T, Sakurai J. 2008. Effect of erythromycin on biological activities induced by Clostridium potent inhibitor of alpha toxin-dependent cytolysis. Mol Microbiol 25: α – 429–440 http://dx.doi.org/10.1046/j.1365-2958.1997.4541820.x. perfringens -toxin. J Pharmacol Exp Ther 327:934 940 http://dx.doi .org/10.1124/jpet.108.143677. 135. Gordon VM, Nelson KL, Buckley JT, Stevens VL, Tweten RK, 148. Chakravorty A, Awad MM, Hiscox TJ, Cheung JK, Choo JM, Lyras Elwood PC, Leppla SH. 1999. Clostridium septicum alpha toxin uses D, Rood JI. 2014. Opioid analgesics stop the development of clostridial glycosylphosphatidylinositol-anchored protein receptors. J Biol Chem gas gangrene. J Infect Dis 210:483–492 http://dx.doi.org/10.1093/infdis 274:27274–27280 http://dx.doi.org/10.1074/jbc.274.38.27274. /jiu101. 136. Shin DJ, Lee JJ, Choy HE, Hong Y. 2004. Generation and charac- 149. Stevens DL, Titball RW, Jepson M, Bayer CR, Hayes-Schroer SM, terization of Clostridium septicum alpha toxin mutants and their use in Bryant AE. 2004. Immunization with the C-domain of alpha-toxin diagnosing paroxysmal nocturnal hemoglobinuria. Biochem Biophys Res prevents lethal infection, localizes tissue injury, and promotes host re- Commun 324:753–760 http://dx.doi.org/10.1016/j.bbrc.2004.09.104. sponse to challenge with Clostridium perfringens. J Infect Dis 190:767– fi 137. Melton-Witt JA, Bentsen LM, Tweten RK. 2006. Identi cation of 773 http://dx.doi.org/10.1086/422691. functional domains of Clostridium septicum alpha toxin. Biochemistry 150. Nagahama M, Oda M, Kobayashi K, Ochi S, Takagishi T, Shibutani 45:14347–14354 http://dx.doi.org/10.1021/bi061334p. M, Sakurai J. 2013. A recombinant carboxy-terminal domain of alpha- 138. Mukamoto M, Kimura R, Hang’ombe MB, Kohda T, Kozaki S. toxin protects mice against Clostridium perfringens. Microbiol Immunol 2013. Analysis of tryptophan-rich region in Clostridium septicum alpha- 57:340–345 http://dx.doi.org/10.1111/1348-0421.12036. toxin involved with binding to glycosylphosphatidylinositol-anchored 151. Aronoff DM. 2013. , sordellii,andtetani:mecha- proteins. Microbiol Immunol 57:163–169 http://dx.doi.org/10.1111/1348 nisms of disease. Anaerobe 24:98–101 http://dx.doi.org/10.1016/j.anaerobe -0421.12017. .2013.08.009. 139. Kennedy CL, Lyras D, Cordner LM, Melton-Witt J, Emmins JJ, 152. Jóźwiak J, Grzela T, Jankowska-Steifer E, Komar A, Moskalewski S, Tweten RK, Rood JI. 2009. Pore-forming activity of alpha-toxin is Martirosian G. 2007. Lethal factor of kills cells essential for Clostridium septicum-mediated myonecrosis. Infect Immun by apoptosis. FEMS Immunol Med Microbiol 49:296–303 http://dx.doi – 77:943 951 http://dx.doi.org/10.1128/IAI.01267-08. .org/10.1111/j.1574-695X.2006.00204.x. 140. Kennedy CL, Smith DJ, Lyras D, Chakravorty A, Rood JI. 2009. 153. Genth H, Pauillac S, Schelle I, Bouvet P, Bouchier C, Varela-Chavez Programmed cellular necrosis mediated by the pore-forming alpha-toxin C, Just I, Popoff MR. 2014. Haemorrhagic toxin and lethal toxin from from Clostridium septicum. PLoS Pathog 5:e1000516 http://dx.doi.org Clostridium sordellii strain vpi9048: molecular characterization and /10.1371/journal.ppat.1000516. comparative analysis of substrate specificity of the large clostridial 141. Hang’ombe MB, Kohda T, Mukamoto M, Kozaki S. 2005. Rela- glucosylating toxins. Cell Microbiol 16:1706–1721 http://dx.doi.org tionship between Clostridium septicum alpha-toxin activity and binding /10.1111/cmi.12321. to erythrocyte membranes. J Vet Med Sci 67:69–74 http://dx.doi.org 154. Hausmann R, Albert F, Geissdörfer W, Betz P. 2004. Clostridium /10.1292/jvms.67.69. fallax associated with sudden death in a 16-year-old boy. J Med Microbiol 142. Hang’ombe MB, Mukamoto M, Kohda T, Sugimoto N, Kozaki S. 53:581–583 http://dx.doi.org/10.1099/jmm.0.05495-0. 2004. Cytotoxicity of Clostridium septicum alpha-toxin: its oligomeriza- 155. Biswas R, K D, Sistla S, Chandra Sistla S, Amaranathan A. 2017. tion in detergent resistant membranes of mammalian cells. Microb Pathog Clostridium perfringens: a bacterial pathogen gaining recognition in 37:279–286 http://dx.doi.org/10.1016/j.micpath.2004.09.001. necrotizing pancreatitis. Anaerobe 47:111–114 http://dx.doi.org/10.1016 143. Chakravorty A, Awad MM, Cheung JK, Hiscox TJ, Lyras D, /j.anaerobe.2017.05.006. Rood JI. 2015. The pore-forming α-toxin from Clostridium septicum 156. Rychener L, InAlbon S, Djordjevic SP, Chowdhury PR, Ziech RE, activates the MAPK pathway in a Ras-c-Raf-dependent and independent de Vargas AC, Frey J, Falquet L. 2017. Clostridium chauvoei, an evolu- manner. Toxins (Basel) 7:516–534 http://dx.doi.org/10.3390/toxins tionary dead-end pathogen. Front Microbiol 8:1054 http://dx.doi.org 7020516. /10.3389/fmicb.2017.01054.

ASMscience.org/MicrobiolSpectrum 17 Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21