Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas...

17
Downloaded from www.asmscience.org by IP: 130.194.145.38 On: Fri, 06 Sep 2019 02:16:21 Histotoxic Clostridial Infections MASAHIRO NAGAHAMA, 1 MASAYA TAKEHARA, 1 and JULIAN I. ROOD 2 1 Department of Microbiology, Faculty of Pharmaceutical Sciences, Tokushima Bunri University, Yamashiro-cho, Tokushima 770-8514, Japan; 2 Infection and Immunity Program, Monash Biomedicine Discovery Institute and Department of Microbiology, Monash University, Clayton, Victoria 3800, Australia ABSTRACT The pathogenesis of clostridial myonecrosis or gas gangrene involves an interruption to the blood supply to the infected tissues, often via a traumatic wound, anaerobic growth of the infecting clostridial cells, the production of extracellular toxins, and toxin-mediated cell and tissue damage. This review focuses on host-pathogen interactions in Clostridium perfringens-mediated and Clostridium septicum-mediated myonecrosis. The major toxins involved are C. perfringens α-toxin, which has phospholipase C and sphingomyelinase activity, and C. septicum α-toxin, a β-pore-forming toxin that belongs to the aerolysin family. Although these toxins are cytotoxic, their eects on host cells are quite complex, with a range of intracellular cell signaling pathways induced by their action on host cell membranes. GENERAL CHARACTERISTICS OF HISTOTOXIC CLOSTRIDIAL INFECTIONS Many pathogenic clostridial species cause potentially fatal soft tissue infections in humans and animals be- cause of their ability to produce extracellular protein toxins ( 1, 2). The ability of these pathogens to produce spores that are resistant to environmental stress is an important factor in the epidemiology of these diseases. Disease pathogenesis involves the growth of the clos- tridial pathogen in the tissues and extensive tissue de- struction, which is the result of the action of extracellular toxins. Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in cattle. Although there are several clostridial species that are responsible for these syndromes ( Table 1), this review will focus on histotoxic infections caused by Clostridium perfringens and Clostridium septicum, primarily because these are the species that have been the subject of the most extensive molecular and func- tional studies. Although the histotoxic clostridia cause severe, rapidly developing infections, these potent toxigenic bacteria still must be regarded as opportunistic patho- gens. All of these infections require predisposing conditions to cause disease, often a traumatic wound that enables the entry of spores from the soil or gastro- intestinal tract to gain access to ischemic internal organs or soft tissues of the body. Alternatively, in nontrau- matic myonecrosis caused by C. septicum, gastrointes- tinal malignancy is commonly associated with the onset of disease, with tumor development leading to ulceration and necrosis of the gastrointestinal mucosa and entry of the gastrointestinal microbiota into the circulation and subsequent infection at distal sites ( 3, 4). Clostridial necrotizing soft tissue infections, especially those caused by C. perfringens, Clostridium sordellii, and Clostri- dium novyi, also have resulted from the subcutaneous injection of contaminated black tar heroin ( 57). Received: 19 March 2018, Accepted: 23 March 2018, Published: 26 July 2019 Editors: Vincent A. Fischetti, The Rockefeller University, New York, NY; Richard P. Novick, Skirball Institute for Molecular Medicine, NYU Medical Center, New York, NY; Joseph J. Ferretti, Department of Microbiology & Immunology, University of Oklahoma Health Science Center, Oklahoma City, OK; Daniel A. Portnoy, Department of Molecular and Cellular Microbiology, University of California, Berkeley, Berkeley, CA; Miriam Braunstein, Department of Microbiology and Immunology, University of North Carolina-Chapel Hill, Chapel Hill, NC, and Julian I. Rood, Infection and Immunity Program, Monash Biomedicine Discovery Institute, Monash University, Melbourne, Australia Citation: Nagahama M, Takehara M, Rood JI. Histotoxic clostridial infections. Microbiol Spectrum 6(4):GPP3-0024-2018. doi:10.1128/microbiolspec.GPP3-0024-2018. Correspondence: Masahiro Nagahama, [email protected] -u.ac.jp or Julian I. Rood, [email protected] ASMscience.org/MicrobiolSpectrum 1 © 2019 American Society for Microbiology. All rights reserved. 2019.

Transcript of Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas...

Page 1: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

Histotoxic Clostridial InfectionsMASAHIRO 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 orgas gangrene involves an interruption to the blood supply tothe infected tissues, often via a traumatic wound, anaerobicgrowth of the infecting clostridial cells, the production ofextracellular toxins, and toxin-mediated cell and tissuedamage. This review focuses on host-pathogen interactionsin Clostridium perfringens-mediated and Clostridiumsepticum-mediated myonecrosis. The major toxins involvedare C. perfringens α-toxin, which has phospholipase C andsphingomyelinase activity, and C. septicum α-toxin,a β-pore-forming toxin that belongs to the aerolysin family.Although these toxins are cytotoxic, their effects on hostcells are quite complex, with a range of intracellular cell signalingpathways induced by their action on host cell membranes.

GENERAL CHARACTERISTICS OFHISTOTOXIC CLOSTRIDIAL INFECTIONSMany pathogenic clostridial species cause potentiallyfatal soft tissue infections in humans and animals be-cause of their ability to produce extracellular proteintoxins (1, 2). The ability of these pathogens to producespores that are resistant to environmental stress is animportant factor in the epidemiology of these diseases.Disease pathogenesis involves the growth of the clos-tridial pathogen in the tissues and extensive tissue de-struction, which is the result of the action of extracellulartoxins. Typical histotoxic clostridial diseases includehuman gas gangrene or myonecrosis and blackleg incattle. Although there are several clostridial species thatare responsible for these syndromes (Table 1), thisreview will focus on histotoxic infections caused byClostridium perfringens and Clostridium septicum,primarily because these are the species that have beenthe subject of the most extensive molecular and func-tional studies.

Although the histotoxic clostridia cause severe,rapidly developing infections, these potent toxigenicbacteria still must be regarded as opportunistic patho-gens. All of these infections require predisposingconditions to cause disease, often a traumatic woundthat enables the entry of spores from the soil or gastro-intestinal tract to gain access to ischemic internal organsor soft tissues of the body. Alternatively, in nontrau-matic myonecrosis caused by C. septicum, gastrointes-tinal malignancy is commonly associated with the onsetof disease, with tumor development leading to ulcerationand necrosis of the gastrointestinal mucosa and entry ofthe gastrointestinal microbiota into the circulation andsubsequent infection at distal sites (3, 4). Clostridialnecrotizing soft tissue infections, especially those causedby C. perfringens, Clostridium sordellii, and Clostri-dium novyi, also have resulted from the subcutaneousinjection of contaminated black tar heroin (5–7).

Received: 19 March 2018, Accepted: 23 March 2018,Published: 26 July 2019

Editors: Vincent A. Fischetti, The Rockefeller University, New York,NY; Richard P. Novick, Skirball Institute for Molecular Medicine,NYU Medical Center, New York, NY; Joseph J. Ferretti, Departmentof Microbiology & Immunology, University of Oklahoma HealthScience Center, Oklahoma City, OK; Daniel A. Portnoy, Departmentof Molecular and Cellular Microbiology, University of California,Berkeley, Berkeley, CA; Miriam Braunstein, Department ofMicrobiology and Immunology, University of North Carolina-ChapelHill, Chapel Hill, NC, and Julian I. Rood, Infection and ImmunityProgram, Monash Biomedicine Discovery Institute, MonashUniversity, Melbourne, Australia

Citation: Nagahama M, Takehara M, Rood JI. Histotoxic clostridial infections. Microbiol Spectrum 6(4):GPP3-0024-2018. doi:10.1128/microbiolspec.GPP3-0024-2018.

Correspondence: Masahiro Nagahama, [email protected] or Julian I. Rood, [email protected]

ASMscience.org/MicrobiolSpectrum 1

© 2019 American Society for Microbiology. All rights reserved.

2019.

Page 2: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

PATHOGENESIS OF CLOSTRIDIALMYONECROSIS CAUSED BYC. PERFRINGENS AND C. SEPTICUMAs previously described (8), the pathogenesis of theseinfections involves three distinct stages, irrespective ofwhether the source of infection involves a traumaticwound. Stage 1 involves an interruption to the bloodsupply such that the redox potential in the tissues dropsto a level that facilitates spore germination and/or thegrowth of the infecting clostridial cells. Stage 2 involvesbacterial growth and establishment of the conditions fortoxin production. In C. perfringens, regulation of toxinproduction primarily utilizes the VirSR two-componentsignal transduction system (9) and an accessory growthregulator-like quorum sensing system (10, 11), as wellas other regulatory networks. Stage 3 encompasses thetoxin-mediated cell and tissue damage that leads to ne-crosis, systemic toxicity, and clinical disease (8).

C. perfringens type A is the major cause of traumaticgas gangrene, although other histotoxic clostridia mayalso be responsible for this sporadic, but fulminantand often fatal, disease (Table 1) (1). The major toxinproduced by gas gangrene strains of C. perfringens isα-toxin, a zinc metallophospholipase that has bothphospholipase C and sphingomyelinase activity (12).α-Toxin was the first bacterial toxin to be shown to haveenzymatic activity (13). Two independent studies usingdifferent experimental approaches have shown thatα-toxin is essential for C. perfringens-mediated myo-necrosis. First, immunization studies in mice using re-combinant α-toxin variants purified from Escherichia

coli, and therefore devoid of any other C. perfringenstoxins, showed that the C-terminal domain of α-toxinwas immunoprotective (14). Second, mutation of theα-toxin structural gene (plc or cpa) abrogated the abilityof the bacterium to cause clostridial myonecrosis in amurine model. Virulence was restored by complemen-tation in trans with a recombinant plasmid containingthe wild-type plc gene (15), thereby fulfilling molecularKoch’s postulates and providing proof of the essentialrole of this toxin in disease.

The other major toxin produced by C. perfringenstype A is perfringolysin O, or θ-toxin, a pore-formingtoxin that is a member of the cholesterol-dependent cy-tolysin family (16, 17). Although mutation of the per-fringolysin O structural gene, pfoA, did not eliminatethe ability to cause disease (18, 19) subsequent studiesof plcpfoA double mutants (20) provided evidence thatperfringolysin O has a synergistic effect with α-toxin ondisease pathogenesis (20, 21).

Recent studies have involved the concurrent analy-sis of the transcriptomes of both the host and C. per-fringens in a murine myonecrosis infection (22). Theresults showed that many host genes involved in theinnate immune response to infection were upregulatedin C. perfringens-infected muscle tissues. In the C. per-fringens cells, upregulated genes included those en-coding potential adhesins and proteins associated withthe cell envelope.

C. septicum is the major causative agent of nontrau-matic gas gangrene, often associated with a gastro-intestinal malignancy (4). The major toxin produced

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.

2 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 3: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

by C. septicum is also called α-toxin, but it is not relatedto C. perfringens α-toxin. C. septicum α-toxin is anaerolysin-like pore-forming toxin that is secreted as aninactive prototoxin that is cleaved near the C terminusby membrane-bound furin or furin-like proteases toform the active toxin (23). The toxin then oligomerizeson the membrane and subsequently inserts to form amembrane pore, which ultimately leads to cell lysis (24).Genetic studies have shown that α-toxin is essential forC. septicum-mediated myonecrosis in mice (25). Muta-tion of the α-toxin structural gene, csa, leads to a loss ofvirulence, which is restored by complementation withthe wild-type csa gene.

STRUCTURE AND FUNCTIONOF C. PERFRINGENS α-TOXINα-Toxin is a 43-kDa single polypeptide that is a hemo-lytic, cytotoxic, dermonecrotic, and lethal extracellu-lar toxin (26–29). Moreover, it induces contractionsof blood vessels and smooth muscle, aggregation ofplatelets, superoxide production, and cytokine release.Importantly, together with perfringolysin O, it is re-sponsible for the hallmark histopathological feature ofC. perfringens-mediated myonecrosis, the paucity of thepolymorphonuclear leukocyte (PMNL) influx into theinfected lesion (15, 19–21, 30).

Determination of the crystal structure of α-toxinrevealed that it has an α-helical N-terminal domain(amino acids 1 to 246) that is responsible for its enzy-matic activity, a flexible linker (amino acids 247 to 255),and a C-terminal (amino acids 256 to 370) β-sandwichmembrane-binding domain (31) (Fig. 1). The two do-mains also are linked by hydrophobic surface interac-tions between their neighboring faces. The N-terminaldomain has significant sequence similarity to phospho-lipase C enzymes from Bacillus cereus (32), Clostridiumbifermentans (33), and Listeria monocytogenes (34). Itcomprises nine densely packed α-helices and resemblesthe structure of B. cereus phospholipase C (31, 35),which has enzymatic activity, but is neither toxic norhemolytic. There are three loops in the N-terminal do-main: loop 1 (residues 70 to 90), loop 2 (residues 135to 150), and loop 3 (residues 205 to 215). These loopsserve a vital role in the direct interaction between thetoxin and the membrane-binding component. The C-terminal domain comprises eight-stranded antiparallelβ-sheets (31). This domain plays a role in membranebinding and is required for hemolytic and toxic activity(28, 36). It is absent from nontoxic phospholipase C en-zymes (31). It has structural similarity to the C-terminal

domain of pancreatic lipase, the N-terminal domain ofsoybean lipoxygenase, and the C2 phospholipid-bindingdomains of eukaryotic intracellular proteins, such assynaptotagmin I, which are involved in signal trans-duction and vesicular transport (37, 38).

The active site in the N-terminal domain is situatedwithin a solvent-accessible cleft that contains two zincions and an exchangeable divalent cation. His-148 andGlu-152 dock with one zinc ion, which is critical for thecatalytic site of the toxin. His-11 and Asp-130 tightlydock with the other zinc ion, which is needed for main-tenance of the structure. His-68, His-126, and His-136dock with an exchangeable divalent cation, which isrequired for binding to target cell membranes (39).The external loop regions (residues 70 to 90, 135 to 150,205 to 215, 265 to 275, 290 to 300, and 330 to 340)bind to phospholipids in the host cell membranes (31).Thr-74 is present in one of these loops, and althoughthe substitution derivative T74I is still active and cancleave a water-soluble substrate, its membrane-damagingactivity toward red blood cells and phosphatidylcholine-cholesterol liposomes is decreased compared to wild-typetoxin (40, 41). However, the structure of the T74I variantis identical to that of the wild-type toxin (41). The T74Isubstitution results in altered recognition of the glycerolmoiety in the phospholipid, without affecting the func-tion of the active site, supporting the hypothesis that theamino acid 70 to 90 loop is important for phospholipidrecognition (42, 43).

FIGURE 1 Structure of α-toxin. The structure of α-toxin wasobtained from the Protein Data Bank (code 1QM6) and wascreated using MacPyMOL. A ribbon representation is shown,colored from blue (N-terminal) to red (C-terminal). Two zincmetal ions in the active site are shown as red spheres.

ASMscience.org/MicrobiolSpectrum 3

Histotoxic Clostridial Infections

Page 4: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

Computational modeling of α-toxin and phosphati-dylcholine revealed that Tyr-57 and Tyr-65, which existat an external site near the catalytic cleft and are inproximity with the 70 to 90 loop, are responsible forsubstrate interaction (44). The replacement of theseTyr residues with Ala did not have any effect on phos-pholipase C and sphingomyelinase activity but had amarked effect on the membrane-damaging activity (he-molysis of sheep red cells and disruption of liposomes).These residues exist in or near the entrance of the cata-lytic cleft and are essential for the membrane-damagingeffect of α-toxin, because they are required for entry ofthe catalytic cleft into the hydrophobic region of plasmamembranes (44).

Several studies have shown that the calcium-bindingC-terminal domain is responsible for membrane bind-ing (28, 31, 36, 37). Two acrylodan-labeled C-terminaldomain derivatives, S263C and S365C, bound to lipidbilayers and displayed a chromogenic shift, demonstrat-ing entry of the C-domain of the toxin into the hydro-phobic regions of the liposomes (36). The data provideevidence that the C-terminal domain of the toxin binds

to the membrane and that the amino acid 70 to 90 loopand Tyr-57 and Tyr-65 serve an important role in the re-cognition of target membranes (Fig. 2). Other studieshave shown that substitution of the C-terminal domainresidues Asp-269, Asp-336, Tyr-275, and Tyr-331,which were predicted to be involved in membrane bind-ing (31), were required for hemolytic and cytotoxic ac-tivity (45).

INTERACTION OF C. PERFRINGENSα-TOXIN WITH HOST CELLSMembrane-Damaging Activity onPhospholipid BilayersThe effect of α-toxin on host tissues is caused by thephospholipase C- and sphingomyelinase-mediated cleav-age of phospholipids in the host cell membranes. Theresultant disruption of the phospholipid bilayer togetherwith the end products of these cleavage reactions, diacyl-glycerol and ceramide, respectively, activates lipid me-tabolism in host cell membranes and plays a role in hostcell damage (27, 28).

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

4 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 5: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

The composition of the phospholipid bilayer modu-lates the extent of the α-toxin-mediated membrane dis-ruption (46). In another study that measured the releaseof carboxyfluorescein from various liposomes, α-toxin-mediated release, or liposomal leakage, decreased withincreasing chain length of the hydrocarbon residuesin the phospholipid bilayer (40). It was concluded thatmembrane disruption was related to the fluidity of themembrane and double bonds in the acyl chain. Sub-sequently, several phosphatidylcholines (C18:0/C18:1)with a double bond in the sn-2 fatty acyl chain wereprepared (47). The phase transition temperature valuewas lowest when the cis-double bond was positioned atthe center of the C18:1 sn-2 acyl chain and increasedcontinuously as the double bond was moved towardeither end of the acyl chain. α-toxin-mediated carboxy-fluorescein release from various liposomes was highestwhen the cis-double bond was positioned at the center ofthe sn-2 fatty acyl chain, resulting in a bell-shaped curve(47). Accordingly, the membrane-disrupting activity ofthe toxin is strongly associated with the membrane flu-idity of target membranes.

There is evidence that α-toxin also promotes per-fringolysin O-induced membrane damage in phospholipidbilayers. Hydrolysis of phosphatidylcholine residues ofliposomes that contained cholesterol increased the amountof free cholesterol in the lipid bilayer and enhancedperfringolysin O-dependent cytolytic activity (48). Thesedata may at least partly explain the synergistic effects of α-toxin and perfringolysin O that were previously observedin C. perfringens infections of mice (20).

Activation of PhospholipidMetabolism and Hemolysisα-toxin provokes contraction of isolated rat ileum andaorta tissues in a dose-dependent manner through stim-ulation of the phospholipid metabolic pathway (49, 50).In isolated rat organs, α-toxin activates the phospha-tidylinositol metabolic pathway and then the arachi-donic acid cascade, thereby promoting thromboxaneproduction. Contraction caused by the toxin was evokedby thromboxane A2 generation induced by arachidonicacid (49, 50). The metabolism of arachidonic acid resultsin the production of bioactive lipid mediators, includingleukotrienes, thromboxanes, and prostaglandins (27).These bioactive lipid mediators play key roles in en-hancing the local inflammatory reaction and lead tovasoconstriction, which is responsible for inducing an-oxia in the early stages of bacterial infections. Theseeffects presumably contribute to the capacity of C. per-fringens to proliferate in the host and cause disease.

Treatment of rabbit neutrophils with α-toxin pro-motes the formation of diacylglycerol, which is fol-lowed by adhesion of the cells to extracellular matrixproteins and the generation of superoxide ions (Fig. 3B)(51). α-Toxin induces diacylglycerol generation throughactivation of intrinsic phospholipase C enzymes by apertussis toxin-sensitive GTP-binding protein. More-over, it increases the phosphorylation of neurotrophictyrosine kinase receptor type 1 (TrkA), leading to 3-phosphoinositide-dependent protein kinase 1 activation(52). Therefore, α-toxin independently causes stimula-tion of intrinsic phospholipase C and protein kinase 1activity through TrkA receptor activation. Both phe-nomena synergistically augment the activity of proteinkinase C θ, leading to the formation of superoxide ionsvia activation of the mitogen-activated protein kinase(MAPK) cascade (Fig. 3B) (52). In human lung carci-noma (A549) cells, α-toxin mediates the activation of thep38 MAPK, NF-κB, and extracellular signal-regulatedkinase 1/2 (ERK1/2) through TrkA activation, whichleads to interleukin-8 (IL-8) release (53) (Fig. 3B).

Treatment of rabbit erythrocytes with α-toxin in-duces the biphasic formation of phosphatidic acid(54) (Fig. 3A). α-Toxin-induced formation of phospha-tidic acid is activated in the earlier stage (30 s) by in-trinsic phospholipase C and in the later stage (ca. 25min) by intrinsic phospholipase D. Both processesinvolve stimulation of the G-protein-signaling pathwayby toxin-mediated membrane phospholipid cleavage(55) (Fig. 3A).

Sheep erythrocytes have higher amounts of sphingo-myelin but do not contain phosphatidylcholine. Cleav-age of sphingomyelin by α-toxin leads to increased levelsof ceramide and sphingosine 1-phosphate (56), which isessential for toxin-induced hemolysis (57) and activatesthe sphingomyelin metabolism system (Fig. 3A) (56).The resultant cleavage of N-nervonoic sphingomyelin(C24:1-SM) occurs in a detergent-resistant membranecomponent (57) (Fig. 3A). Pertussis toxin blocks theproduction of C24:1-ceramide caused by α-toxin andblocks red cell hemolysis, which indicates that intrinsicsphingomyelinase, which can cleave C24:1-SM to C24:1-ceramide, is regulated by a pertussis toxin-sensitive G-protein in plasmamembranes. In addition, the activationof Rho protein by α-toxin increases sphingosine kinaseactivity. Subsequently, the ceramides are quickly me-tabolized to sphingosine, indicating that the toxin alsoactivates ceramidases (57). These results indicate thatthe cleavage of C24:1-SM to sphingosine 1-phosphate byα-toxin in lipid rafts is essential for toxin-induced he-molysis of sheep red cells (Fig. 3A).

ASMscience.org/MicrobiolSpectrum 5

Histotoxic Clostridial Infections

Page 6: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

Interaction of α-Toxin with GangliosidesAmino acid sequence analysis of botulinum toxin andtetanus toxin showed that the ganglioside-GT1b- andlactose-binding site is determined by the existence of theconserved peptide motif, H.....SXWY.....G (ca. 30 aminoacids [aa]), with the Tyr and Trp residues being partic-ularly critical (58). A similar motif (H.....SWY.....G,amino acids 68 to 93) is present in α-toxin (53), in theregion (amino acids 55 to 93) that is located at the in-terface between the N-terminal and C-terminal domains(42). It has been demonstrated that the 70 to 90 loopin α-toxin plays a role in binding to the cell surface(43) (Fig. 2). Computational docking studies indicatedthat Trp-84 binds to the sialic acid residue on ganglio-side GM1a via an aromatic stacking interaction and ahydrogen-bonding interaction and that Tyr-85 associ-ates with the galactosamine residue on GM1a via anα-glycosidic bonding interaction (59). Replacement ofTrp-84 and Tyr-85 in α-toxin, which correspond to

the ganglioside recognition site of botulinum and teta-nus toxins, significantly reduced binding to gangliosideGM1a-containing liposomes. W84A and Y85A deriva-tives had markedly decreased capabilities to stimulateTrkA (59). Ganglioside GM1a assembles with a TrkAmolecule on the cell surface and stimulates TrkA andERK1/2 (60). It has been reported that ganglioside GM1clustering leads to the condensation of TrkA in lipid raftsand the stimulation of downstream signaling targets(61). α-Toxin-mediated diacylglycerol production leadsto diacylglycerol flip-flop movement that affects thedynamics of cell membranes (Fig. 4). Therefore, it servesin GM1a clustering and intrinsic phospholipase C-γ1activation through TrkA, triggering various signaltransduction activities (62–64).

α-toxin causes higher cytotoxicity in a ganglioside-deficient cell line (Don Q) (65) and can be internalizedby dynamin-mediated endocytosis (66, 67). The toxinbinds to caveolin-1 both on the cell membranes and in

FIGURE 3 Relationship between phospholipid metabolism and the biological activitiesof α-toxin. (A) Signaling events involved in α-toxin-induced hemolysis of sheep or rabbiterythrocytes. (B) Signaling events involved in α-toxin-activated generation of O2

– inneutrophils 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, phosphatidylinositidedependent kinase 1; DRM, detergent-resistant membrane; TrkA, neurotrophic tyrosinekinase receptor type1; MAPKK, mitogen-activated protein kinase kinase; Erk1/2, extra-cellular signal-regulated kinase 1/2.

6 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 7: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

endosomal vesicles, and it also induces ceramide pro-duction in the cell membrane (28), which generates neg-ative membrane curvature and facilitates endocytosis(68). Internalized α-toxin is transported to the early andlate endosomes and lysosomes. Lysosomes are damagedduring progressive α-toxin internalization into Don Qcells. In addition, the internalization of α-toxin is alsomediated through the stimulation of MEK/ERK signal-ing (66, 67). It has been reported that the toxin causesreactive oxygen species generation via PKC, MEK/ERK,and NF-κB signaling activities in Don Q cells (66, 67).Therefore, α-toxin perturbs the host cell signaling cas-cade and acts not only on the cell membrane, but also inthe cytoplasm.

Effect of α-Toxin on Hematopoietic CellsThe first line of defense of the innate immune systeminvolves neutrophils, which play an important role in thephagocytosis and subsequent elimination of pathogenicbacteria (69, 70). Normally, the number of neutrophils

is sustained in a steady state through granulopoiesis,which is accelerated to replenish neutrophils during bac-terial infection (71–73). Pattern recognition receptors,such as Toll-like receptors (TLRs), are responsible forthe recognition of structural components of microor-ganisms (74, 75). TLR2 and TLR4 have been identifiedas being pivotal for the recognition of Gram-positive orGram-negative bacteria by recognizing cell wall com-ponents such as peptidoglycan and bacterial endotoxin(lipopolysaccharide), respectively (76–79). During bac-terial infection, these bacterial components stimulatethe production of granulocyte colony-stimulating factor(G-CSF), which is a glycoprotein that promotes theproliferation, survival, and differentiation of neutrophilsand their progenitor cells (80) from endothelial cells andmonocytes, resulting in the acceleration of granulopoie-sis (81). G-CSF-deficient and G-CSF receptor-deficientmice exhibit chronic neutropenia and reduced infection-driven granulopoiesis, leading to impaired ability toeliminate infecting bacteria (82, 83). Thus, granulo-

FIGURE 4 Model of α-toxin-induced membrane dynamics and clustering of the GM1aand TrkA complex. α-Toxin binds to GM1a and, through the phospholipase C activity of thetoxin itself, causes diacylglycerol (DG) production at the outer membrane. The flip-flopmovement of DG affects membrane dynamics, facilitating clustering of GM1a and TrkAactivation 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 enhancedactivation of TrkA, triggering activation of signal transduction pathways, which induceneutrophil activation. Abbreviations: PIP2, phosphatidylinositol 4,5-bisphosphate; PLC,phospholipase C; DG, diacylglycerol; PC, phosphatidylcholine; TrkA, neurotrophic tyro-sine kinase receptor type 1.

ASMscience.org/MicrobiolSpectrum 7

Histotoxic Clostridial Infections

Page 8: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

poiesis is precisely regulated to replenish neutrophilsduring bacterial infection. However, C. perfringens canstill cause life-threatening infections, and the mecha-nisms by which C. perfringens avoids these host defensesare still not completely understood.

C. perfringens-mediated myonecrosis progresses sorapidly that death, which is associated with muscle de-struction, shock, and multiple organ failure, sometimesprecedes an accurate diagnosis (84). C. perfringens in-fection is characterized by a greatly reduced PMNL in-flux to the site of the infection (19, 85–87), suggestingthat the bacteria disrupt neutrophil-based innate immu-nity to evade host defenses. α-Toxin mediates the forma-tion of platelet-leukocyte aggregates, leading to vascularocclusion, and induces a marked reduction in micro-vascular perfusion (45, 86, 88), which is thought toimpede neutrophil extravasation (89). Recently, it wasshown using a mouse model of infection that C. per-fringens infection reduced mature neutrophils in bonemarrow, in peripheral blood, and in C. perfringens-

infected muscle, in an α-toxin-dependent manner (90).α-Toxin inhibits neutrophil differentiation and impairsthe replenishment of differentiated mature neutrophils inthe peripheral circulation, resulting in reduced recruit-ment of neutrophils to the C. perfringens infection site.The half-life of peripheral neutrophils is less than 12hours in vivo (91), meaning that the number of matureneutrophils required to counter a bacterial infection issustained through continuous granulopoiesis. Therefore,blockage of both granulopoiesis and PMNL diapedesisinto the tissues, as a result of platelet aggregation, leadsto a reduction in cell numbers in a short period of time.Thus, C. perfringens infection impairs the host immunesystem, which could explain the both scarcity of PMNLat the site of infection and a propensity for polymicrobialinfections in many patients (Fig. 5).

α-Toxin is essential for disease pathogenesis (18–21,30), and blockage of neutrophil differentiation is de-pendent on its phospholipase C and sphingomyelinaseactivities (90). In human lymphocytes, S. aureus sphin-

FIGURE 5 Inhibition of granulopoiesis and erythropoiesis in a C. perfringens-infectedhost. Infection with C. perfringens diminishes mature hematopoietic cells in the bonemarrow, leading to impairment of replenishment of differentiated neutrophils and eryth-rocytes in the peripheral circulation and resulting in host innate immune deficiency. Themajor virulence factor 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

Nagahama et al.

Page 9: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

gomyelinase disrupts lipid rafts, which are cholesterol-rich plasma membrane microdomains (92). Lipid raftsfunction as platforms of signaling molecules involved inthe regulation of cell differentiation in many cell types(93, 94). Ceramide, the hydrolysis product of sphingo-myelinase, also is known to be a lipid messenger impli-cated in various cellular responses, including apoptosis,differentiation, and cell growth (95). Recently, C. per-fringens α-toxin was reported to perturb the integrityof lipid rafts, leading to the blockage of neutrophil dif-ferentiation (90). The detailed molecular mechanismbehind this blockage has not been elucidated, but it maybe important to focus on the relationship between therole of sphingomyelinase and cell differentiation to shedfurther light on host-pathogen interactions in clostridialmyonecrosis.

Effect of α-Toxin on Erythropoiesisα-Toxin has been reported to lyse erythrocytes fromvarious species. It activates T-type Ca2+ channels and in-creases intracellular Ca2+, which plays an important rolein the hemolysis of horse erythrocytes (96). The toxinactivates the sphingomyelin metabolic system, leadingto the hemolysis of sheep erythrocytes (56, 57), and thelysis of rabbit erythrocytes involves α-toxin-mediatedactivation of endogenous phospholipase C activity (54,55). Recently, Bacillus anthracis lethal toxin was re-ported not only to induce hemolysis, but also to inhibiterythroid differentiation of cord blood-derived CD34+

hematopoietic stem cells, leading to the suppressionof erythropoiesis (97). The toxin-induced suppressionof erythropoiesis is potentially part of a lethal toxin-mediated pathophysiology in anemia and hypoxia. Sim-ilarly, it was reported that mature erythroblasts werepreferentially decreased when isolated mouse bonemarrow cells were treated with α-toxin, and subsequentexperiments revealed that the blockage of erythroiddifferentiation was involved in the reduction in matureerythroblasts by α-toxin (98). Massive intravascularhemolysis and severe anemia have been reported inC. perfringens-infected patients (99), although it is arelatively uncommon event. The detailed mechanismsby which C. perfringens infection causes severe anemiahave not been elucidated, but blockage of erythropoiesisby α-toxin may be important (Fig. 5).

Experiments carried out using an inactive α-toxinH148G variant revealed that its enzymatic activity wasrequired for the α-toxin-mediated reduction of erythro-blast levels (98). Furthermore, the cell surface expressionof a lipid raft marker, GM1, decreased in associationwith erythroid differentiation, and α-toxin affected lipid

raft integrity in erythroid progenitors, leading to theimpairment of erythroid differentiation. During eryth-roblast enucleation, the clustering of lipid rafts at thefurrow between incipient reticulocytes and pyrenocytesis observed, and an inhibitor of lipid rafts impedesenucleation (100). Therefore, lipid rafts play an impor-tant role in the regulation of erythropoiesis, but thedetailed mechanism by which the disturbance in lipidraft integrity by α-toxin impairs erythroid differentiationremains unclear.

STRUCTURE AND HOST CELLINTERACTIONS OF C. PERFRINGENSPERFRINGOLYSIN OPerfringolysin O (θ-toxin) is a pore-forming hemolytictoxin that belongs to the cholesterol-dependent cyto-lysin (CDC) family (16, 101), which includes streptoly-sin O from Streptococcus pyogenes, listeriolysin O fromL. monocytogenes, pneumolysin from Streptococcuspneumoniae (17), and related cytolysins producedby several other clostridia (102). The perfringolysin Ostructural gene, pfoA, is located on the chromosome(103, 104). Perfringolysin O is elaborated by most, butnot all, strains of C. perfringens (104–107), and matureperfringolysin O (53 kDa, 473 aa) (108) is secretedas a water-soluble monomer and binds to host cellmembranes in a cholesterol-dependent manner (102).Binding to the cell membrane initiates oligomerizationand the formation of a pore complex that can containup to 50 monomeric subunits and is responsible forcell lysis.

Perfringolysin O monomers have a hydrophilic,β-sheet-rich structure that can be divided into four dis-tinct domains (109, 110) (Fig. 6). Domain 1 contains aseven-stranded antiparallel β-sheet flanked by helicesand loops. Domain 2 forms an elongated β-sheet struc-ture. Domain 3 comprises α-helices and β-sheets, anddomain 4 (the C-terminal region) is divided into twodistinct β-sandwich regions. Domain 4 also contains ahighly conserved 11-aa Trp-rich motif (ECTGLAWEWWR) or undecapeptide loop that is near the C-terminus.This structure subsequently was shown to be similar tothat of members of the eukaryotic membrane attackcomplex/perforin family (111).

Perfringolysin O and most of the other cholesterol-dependent cytolysins bind directly to cholesterol in thehost cell membrane. Contrary to what was thoughtfor many years, the undecapeptide is not the receptor-binding motif (112). Instead, there is a Thr-Ile pairlocated in a nearby loop (L1) of domain 4 that is re-

ASMscience.org/MicrobiolSpectrum 9

Histotoxic Clostridial Infections

Page 10: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

sponsible for binding to cholesterol (113). The role ofthe undecapeptide is to assist in anchoring the toxinmonomer to the membrane and to couple membraneinsertion with oligomerization (112). The next step inthe oligomerization process is a conformational changeof two α-helical bundles in domain 3 to form two longeramphipathic β-hairpins; this extended β-hairpin struc-ture inserts into the host cell membrane, initially to formtoxin dimers and then to form a large oligomeric β-barrelpore that contains 35 to 45 monomers. The biologicaleffect of large pore formation is to induce cell lysis byosmotic stress (114). In addition, perfringolysin O caninterfere with cell signaling events by activating TLR4and inducing inducible nitric oxide synthase expres-sion and tumor necrosis factor-α (TNFα) secretion inbone marrow-derived macrophages (115) and by stim-ulating the degradation of Ubc9, thereby reducing theposttranslational SUMOylation of host proteins (116).Therefore, perfringolysin O has effects on both mem-brane integrity and internal cell signaling.

Perfringolysin O works synergistically with α-toxin tocause gas gangrene by promoting leukostasis and in-travascular coagulation (19–21, 117). The result is de-creased blood flow (87, 88) and the accumulation of

PMNL and macrophages in the blood vessels around thesite of infection, which leads to vascular leukostasis andcontributes to the paucity of leukocyte infiltration tothe site of infection (20, 21). Similarly, intramuscularinjection of a crude C. perfringens toxin preparationcauses and promotes the formation of intravascularplatelet aggregates (87). Leukocytes and fibrin accumu-late rapidly in the aggregate, leading to obstruction ofthe blood vessels and the production of thrombi thatreduce 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-activatingfactor in endothelial cells (19, 118, 119, 121). The toxintriggers platelet/leukocyte and platelet/platelet aggrega-tion by activating the platelet fibrinogen receptor gpIIb/IIIa synergistically with α-toxin (86, 122). Moreover, ithas been reported to impair the polymerization of actinfilaments in leukocytes and the migration of neutrophilsin response to chemoattractants (118, 123). These eventslead to the formation of intravascular platelet/leucocyte/fibrin aggregates, and the adherence of the aggregatesto the vascular endothelial cells results in vascular in-jury, vessel obstruction, hypoxia, and tissue destruction(1, 30, 124). In addition, a direct cytotoxic effect ofperfringolysin O on endothelial cells has been demon-strated, which also impairs leukocyte transmigration(30, 88). Disruption of the endothelium contributes toprogressive edema (30). α-Toxin impairs granulopoiesis,leading to the reduction of neutrophils, and a directcytolytic effect of perfringolysin O on leukocytes hasbeen reported (123). Therefore, perfringolysin O workssynergistically with α-toxin to reduce the level of leuko-cytes. Finally, C. perfringens cells can escape from thephagosome of macrophages and survive, and perfringo-lysin O is an important factor in this escape process(125, 126). These data are in agreement with geneticstudies that showed that α-toxin and perfringolysin Oact synergistically in the disease process (19, 20).

In the latter stage of clostridial gas gangrene, cardio-vascular collapse, tachycardia, and multiorgan failureare observed in patients (1). Toxins such as perfringolysinO and α-toxin are released into the blood and can act ata distance from the site of infection. Perfringolysin O re-duces systemic vascular resistance, increases cardiac out-put, and decreases the heart rate without a drop in meanarterial pressure (117, 127). In addition, it promotes therelease of inflammatory cytokines, such as TNFα, IL-1,IL-6, platelet-activating factor, and prostaglandin I2, by

FIGURE 6 Structure of perfringolysin O. The structure of per-fringolysin O was obtained from the Protein Data Bank (code1PFO) and was created using MacPyMOL. A ribbon repre-sentation is shown, colored from blue (N-terminal) to red(C-terminal).

10 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 11: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

acting synergistically with α-toxin, which contributes tothe symptoms of toxic shock, including hypotension,hypoxia, and reduced cardiac output (30, 128).

FUNCTION AND HOST CELL INTERACTIONSOF C. SEPTICUM α-TOXINC. septicum is the primary causative agent of nontrau-matic clostridial myonecrosis and enumerates an α-toxinthat is essential for disease and is responsible for reducedPMNL influx into the infected lesions (25). C. septicumα-toxin is a member of the aerolysin family of pore-forming toxins (129, 130). These toxins include aero-lysin from Aeromonas hydrophila, ε-toxin and C. per-fringens enterotoxin (CPE) from C. perfringens, andparasporin toxins produced by Bacillus thuringiensis.These toxins all form small, often heptameric, β-barrelpores in host cell membranes (130). Although the crystalstructures of several members of the aerolysin familyhave been determined, there is no such structure availablefor α-toxin. However, molecular modelling of α-toxinshows that it does not have the small-lobe lectin-bindingdomain (D1) of aerolysin, known as the aerolysin per-tussis toxin domain (131), but has homologous domainsto the other three aerolysin domains (Fig. 7) (132).

α-Toxin is synthesized as a preprototoxin, with a 31-aa leader sequence that is cleaved upon secretion fromthe bacterial cell and a 45-aa C-terminal extension that iscleaved by host proteases such as furin prior to pore

formation (23, 102, 133). The resultant cleaved pro-peptide prevents premature oligomerization (134). The46-kDa prototoxin binds to glycophosphatidylinositol-anchored protein receptors located in lipid rafts in thehost cell membrane (135). After protease cleavage onthe cell membrane, the 43-kDa mature toxin undergoesa conformational change and oligomerizes to form aheptameric prepore complex that is required for poreformation (24). Cleavage of the C-terminal extension en-ables the formation of a transmembrane β-barrel struc-ture, with each monomer contributing one extendedβ-hairpin loop to the β-barrel that forms the trans-membrane pore (130, 132).

The analysis of substitution derivatives of α-toxinhas yielded insights into the functional domains ofthe toxin (132, 136–138). Y155F and S189C α-toxinderivatives had reduced toxicity for CHO cells butwere still cleaved by host proteases and could bind toglycophosphatidylinositol-anchored receptors and formoligomers. A S189C/S238C derivative retained its abilityto bind, act as a substrate for furin cleavage, and formoligomers but was not toxic (136). Like aerolysin,α-toxin has a Trp-rich motif (amino acids 302 to 312)in its C-terminal domain (138). Several of the residuesin this motif, including three Trp residues, are essentialfor binding and/or cytotoxic activity (137, 138). Fluo-rimetric studies using Cys-substituted derivatives led tothe identification of the amphipathic transmembraneβ-hairpin (Lys-203 to Gln-232) that spans the host cellmembrane. Deletion of this region eliminated the toxin’sability to form pores and kill cells, but the resultantprotein could still bind, undergo cleavage, and formoligomers (132). Subsequently, saturation Ala or Cyssubstitutions were made across the remaining residues inthe α-toxin molecule, and the resultant proteins werepurified and analyzed for their biological and biophysi-cal properties (137). The results showed that the onlyα-toxin Cys residue, Cys-86 in domain 1, was requiredfor receptor binding. In summary, derivatives alteredin receptor binding all had substitutions in domain 1,but proteins with altered oligomerization properties hadchanges in domains 1 and 3. Several of these mutantshave been examined in the mouse myonecrosis model(139). The results showed that both the amphipathicβ-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-mediatedmyonecrosis (129, 140), but pore-mediated osmoticcell lysis is not the only consequence of intoxication byα-toxin. Treatment of C2C12 mouse myoblast cells

FIGURE 7 Crystal structure of aerolysin and molecular modelof C. septicum α-toxin. Domains are indicated. Note thatdomain 1 (D1) of aerolysin is not present in α-toxin. AT,C. septicum α-toxin. Reprinted with permission from refer-ence 137. Copyright (2006) American Chemical Society.

ASMscience.org/MicrobiolSpectrum 11

Histotoxic Clostridial Infections

Page 12: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

with α-toxin leads to a rapid Ca2+ influx into the cell butdoes not induce apoptosis. Instead, α-toxin induces pro-grammed cellular necrosis (140), which is consistentwith the fact that it is more active against nucleated cells(141, 142). α-Toxin-mediated Ca2+ influx resulted in theactivation of proteolytic calpain activity, a reduction inlysosomal integrity, leading to cathepsin release, mito-chondrial dysfunction and ATP depletion, and releasefrom the nucleus of the histone-binding protein HMGB1(140). Induction of this programmed necrotic pathwaywas dependent on the ability to form pores; treatmentwith a substituted α-toxin lacking the β-hairpin loop didnot induce a Ca2+ influx or programmed necrosis. All ofthese changes are consistent with α-toxin-mediated poreformation causing the induction of the programmed cel-lular necrosis pathway (140). Other studies have shownthat α-toxin treatment of Madin-Darby Canine Kidney(MCDK) cells causes a rapid K+ efflux, ATP depletion,and caspase-3-independent necrosis and nonapoptoticcell death (129). Subsequently, it was shown that poreformation by α-toxin led to ERK, c-Jun N-terminal pro-tein kinase, and p38 phosphorylation and activation oftheir resultant MAPK pathways and the resultant dose-dependent release of TNFα (143). C. septicum α-toxinalso leads to a significant reduction in vascular perfusionof muscle tissues, as shown by comparative intravitalmicroscopy studies using culture supernatants from anisogenic series of C. septicum strains (88).

CONTROL AND TREATMENT OFHISTOTOXIC CLOSTRIDIAL INFECTIONSThe treatment of myonecrotic infections caused byC. perfringens and C. septicum is complicated by therapid onset and progression of disease. Death can resultwithin 24 hours of initial symptoms. Standard treatmentregimens involve surgical debridement of the infectedtissues, which needs to occur as early as possible (2),and treatment with penicillin and clindamycin (144).The effectiveness of the penicillin on its own has beenquestioned (145), and there is experimental evidencethat clindamycin and metronidazole, alone or in com-bination with penicillin are more effective in mice (145,146). Often, surgical removal of an infected limb isthe only option available in life-threatening infections.

In other studies, mice lacking TNF-α were protectedfrom C. perfringens α-toxin-induced lethality (147),demonstrating that TNF-α release is essential for thelethal effect of the toxin. Treatment with erythromycinblocked α-toxin-induced release of TNF-α from periph-eral neutrophils and the activation of TrkA and ERK1/2,

demonstrating that toxin-induced TNF-α release is re-lated to stimulation of the ERK/MAPK signaling path-way through TrkA. Anti-TNF-α antibodies, but notanti-IL-1β or anti-IL-6 antibodies, blocked the deathof mice and intravascular hemolysis by α-toxin. Com-pounds that block the release of TNF-α therefore maybe valuable in treating patients with C. perfringens in-fection. Recently, experiments with C. perfringens in-fections in mice provided evidence that treatment withopioids such as buprenorphine stop the developmentof disease (148), but this treatment option has not beentested in clinical settings.

Several studies have investigated the use of recombi-nant α-toxin variants to vaccinate againstC. perfringens-mediated gas gangrene, primarily from a military orbiosecurity perspective or for elderly patients that areat risk. Initial studies showed that the isolated nontoxicC-terminal C2 lipid-binding domain of α-toxin (aminoacids 247 to 370) was highly immunoprotective in mice(14). Mice immunized with this formaldehyde-treatedrecombinant antigen were protected from the effectsof α-toxin and from challenge with a 10× LD50 dose ofC. perfringens. By contrast, mice immunized with theequivalent N-terminal enzymatic domain (amino acids 1to 249) produced immunoreactive antibodies, but theseantibodies did not protect against the toxin or disease.These results were confirmed in a subsequent, more de-tailed, vaccination study that also showed that micevaccinated with the C-terminal domain and then chal-lenged in a murine infection model had reduced throm-bosis and an increased PMNL influx into the site ofinfection (149). To further define the immunoprotectiveregion of the C-terminal domain, the potency of severalrecombinant fragments (amino acids 251 to 370, 281 to370, and 311 to 370) fused to glutathione S-transferase(GST) were examined in an active vaccination model(150). Vaccination with the 251 to 370 and 281 to 370GST-fusions protected against the toxicity of α-toxinand against C. perfringens challenge, whereas the aminoacid 311 to 370 derivative yielded only partial protec-tion. Despite these three promising studies, no commer-cial C. perfringens vaccine is currently available.

ACKNOWLEDGMENTSResearch in Julian Rood’s laboratory was supported by projectgrant GNT1082401 from the Australian National Health andMedical Research Council.

REFERENCES1. Stevens DL, Aldape MJ, Bryant AE. 2012. Life-threatening clostridialinfections. Anaerobe 18:254–259 http://dx.doi.org/10.1016/j.anaerobe.2011.11.001.

12 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 13: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

2. Stevens DL, Bryant AE. 2017. Necrotizing soft-tissue infections.NEnglJ Med 377:2253–2265 http://dx.doi.org/10.1056/NEJMra1600673.3. Stevens DL, Musher DM, Watson DA, Eddy H, Hamill RJ, Gyorkey F,Rosen H, Mader J. 1990. Spontaneous, nontraumatic gangrene due toClostridium septicum. Rev Infect Dis 12:286–296 http://dx.doi.org/10.1093/clinids/12.2.286.4. Srivastava I, Aldape MJ, Bryant AE, Stevens DL. 2017. SpontaneousC. septicum gas gangrene: a literature review. Anaerobe 48:165–171http://dx.doi.org/10.1016/j.anaerobe.2017.07.008.5. Kimura AC, Higa JI, Levin RM, Simpson G, Vargas Y, Vugia DJ. 2004.Outbreak of necrotizing fasciitis due toClostridium sordellii among black-tar heroin users. Clin Infect Dis 38:e87–e91 http://dx.doi.org/10.1086/383471.6. Assadian O, Assadian A, Senekowitsch C, Makristathis A, HagmüllerG. 2004. Gas gangrene due to Clostridium perfringens in two injectingdrug users in Vienna, Austria.Wien KlinWochenschr 116:264–267 http://dx.doi.org/10.1007/BF03041058.7. Gonzales y Tucker RD, Frazee B. 2014. View from the front lines: anemergency medicine perspective on clostridial infections in injection drugusers. Anaerobe 30:108–115 http://dx.doi.org/10.1016/j.anaerobe.2014.09.005.8. Stevens DL, Rood JI. 2006. Histotoxic clostridia, p 715–725.In Fischetti VA, Novick RP, Ferretti JJ, Portnoy DA, Rood JI (ed), Gram-Positive Pathogens, 2nd ed. ASM Press, Washington, DC. http://dx.doi.org/10.1128/9781555816513.ch589. Ohtani K. 2016. Gene regulation by the VirS/VirR system in Clostrid-ium perfringens. Anaerobe 41:5–9 http://dx.doi.org/10.1016/j.anaerobe.2016.06.003.10. Ohtani K, Yuan Y, Hassan S, Wang R, Wang Y, Shimizu T. 2009.Virulence gene regulation by the agr system in Clostridium perfringens.J Bacteriol 191:3919–3927 http://dx.doi.org/10.1128/JB.01455-08.11.MaM, Li J, McClane BA. 2015. Structure-function analysis of peptidesignaling in the Clostridium perfringens Agr-like quorum sensing system.J Bacteriol 197:1807–1818 http://dx.doi.org/10.1128/JB.02614-14.12. Titball R, Rood J. 2000. Bacterial phospholipases, p 529–556.In Aktories K, Just I (ed), Bacterial Protein Toxins. Springer, Berlin,Germany. http://dx.doi.org/10.1007/978-3-662-05971-5_2313. Macfarlane MG, Knight BCJG. 1941. The biochemistry of bacterialtoxins: the lecithinase activity ofCl. welchii toxins. Biochem J 35:884–902http://dx.doi.org/10.1042/bj0350884.14. Williamson ED, Titball RW. 1993. A genetically engineered vaccineagainst the alpha-toxin of Clostridium perfringens protects mice againstexperimental gas gangrene. Vaccine 11:1253–1258 http://dx.doi.org/10.1016/0264-410X(93)90051-X.15. Awad MM, Bryant AE, Stevens DL, Rood JI. 1995. Virulence studieson chromosomal α-toxin and θ-toxin mutants constructed by allelic ex-change provide genetic evidence for the essential role of α-toxin in Clos-tridium perfringens-mediated gas gangrene. Mol Microbiol 15:191–202http://dx.doi.org/10.1111/j.1365-2958.1995.tb02234.x.16. Tweten RK. 2005. Cholesterol-dependent cytolysins, a family ofversatile pore-forming toxins. Infect Immun 73:6199–6209 http://dx.doi.org/10.1128/IAI.73.10.6199-6209.2005.17. Gilbert RJ. 2010. Cholesterol-dependent cytolysins. Adv Exp MedBiol 677:56–66 http://dx.doi.org/10.1007/978-1-4419-6327-7_5.18. Awad MM, Rood JI. 1997. Isolation of α-toxin, θ-toxin and κ-toxinmutants of Clostridium perfringens by Tn916 mutagenesis. MicrobPathog 22:275–284 http://dx.doi.org/10.1006/mpat.1996.0115.19. Stevens DL, Tweten RK, Awad MM, Rood JI, Bryant AE. 1997.Clostridial gas gangrene: evidence that α and θ toxins differentiallymodulate the immune response and induce acute tissue necrosis. J InfectDis 176:189–195 http://dx.doi.org/10.1086/514022.20. Awad MM, Ellemor DM, Boyd RL, Emmins JJ, Rood JI. 2001.Synergistic effects of α-toxin and perfringolysin O in Clostridium

perfringens-mediated gas gangrene. Infect Immun 69:7904–7910 http://dx.doi.org/10.1128/IAI.69.12.7904-7910.2001.21. Ellemor DM, Baird RN, Awad MM, Boyd RL, Rood JI, Emmins JJ.1999. Use of genetically manipulated strains of Clostridium perfringensreveals that both alpha-toxin and theta-toxin are required for vascularleukostasis to occur in experimental gas gangrene. Infect Immun 67:4902–4907.22. Low L-Y, Harrison PF, Gould J, Powell DR, Choo JM, ForsterSC, Chapman R, Gearing LJ, Cheung JK, Hertzog P, Rood JI. 2018.Concurrent host-pathogen transcriptional responses in a Clostridiumperfringens murine myonecrosis infection. MBio 9:e00473-18 http://dx.doi.org/10.1128/mBio.00473-18.23. Gordon VM, Benz R, Fujii K, Leppla SH, Tweten RK. 1997. Clos-tridium septicum alpha-toxin is proteolytically activated by furin. InfectImmun 65:4130–4134.

24. Sellman BR, Kagan BL, Tweten RK. 1997. Generation of amembrane-bound, oligomerized pre-pore complex is necessary for poreformation by Clostridium septicum alpha toxin. Mol Microbiol 23:551–558 http://dx.doi.org/10.1046/j.1365-2958.1997.d01-1876.x.

25. Kennedy CL, Krejany EO, Young LF, O’Connor JR, Awad MM,Boyd RL, Emmins JJ, Lyras D, Rood JI. 2005. The alpha-toxin ofClostridium septicum is essential for virulence. Mol Microbiol 57:1357–1366 http://dx.doi.org/10.1111/j.1365-2958.2005.04774.x.

26. Titball RW. 1993. Bacterial phospholipases C. Microbiol Rev 57:347–366.

27. Titball RW, Naylor CE, Basak AK. 1999. The Clostridium per-fringens α-toxin. Anaerobe 5:51–64 http://dx.doi.org/10.1006/anae.1999.0191.

28. Sakurai J, Nagahama M, Oda M. 2004. Clostridium perfringensalpha-toxin: characterization and mode of action. J Biochem 136:569–574 http://dx.doi.org/10.1093/jb/mvh161.

29. Uzal FA, Freedman JC, Shrestha A, Theoret JR, Garcia J, Awad MM,Adams V, Moore RJ, Rood JI, McClane BA. 2014. Towards an under-standing of the role of Clostridium perfringens toxins in human andanimal disease. FutureMicrobiol 9:361–377 http://dx.doi.org/10.2217/fmb.13.168.

30. Stevens DL, Bryant AE. 2002. The role of clostridial toxins in thepathogenesis of gas gangrene.Clin Infect Dis 35(Suppl 1):S93–S100 http://dx.doi.org/10.1086/341928.

31. Naylor CE, Eaton JT, Howells A, Justin N, Moss DS, Titball RW,Basak AK. 1998. Structure of the key toxin in gas gangrene. Nat StructBiol 5:738–746 http://dx.doi.org/10.1038/1447.

32. Gilmore MS, Cruz-Rodz AL, Leimeister-Wächter M, Kreft J, GoebelW. 1989. A Bacillus cereus cytolytic determinant, cereolysin AB, whichcomprises the phospholipase C and sphingomyelinase genes: nucleotidesequence and genetic linkage. J Bacteriol 171:744–753 http://dx.doi.org/10.1128/jb.171.2.744-753.1989.

33. Tso JY, Siebel C. 1989. Cloning and expression of the phospholipaseC gene fromClostridium perfringens andClostridium bifermentans. InfectImmun 57:468–476.

34. Vazquez-Boland J-A, Kocks C, Dramsi S, Ohayon H, Geoffroy C,Mengaud J, Cossart P. 1992. Nucleotide sequence of the lecithinaseoperon of Listeria monocytogenes and possible role of lecithinase in cell-to-cell spread. Infect Immun 60:219–230.

35. Hough E, Hansen LK, Birknes B, Jynge K, Hansen S, Hordvik A,Little C, Dodson E, Derewenda Z. 1989. High-resolution (1.5 A) crystalstructure of phospholipase C from Bacillus cereus. Nature 338:357–360http://dx.doi.org/10.1038/338357a0.

36. Nagahama M, Mukai M, Morimitsu S, Ochi S, Sakurai J. 2002. Roleof the C-domain in the biological activities of Clostridium perfringensalpha-toxin. Microbiol Immunol 46:647–655 http://dx.doi.org/10.1111/j.1348-0421.2002.tb02748.x.

ASMscience.org/MicrobiolSpectrum 13

Histotoxic Clostridial Infections

Page 14: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

37. Guillouard I, Alzari PM, Saliou B, Cole ST. 1997. The carboxy-terminal C2-like domain of the alpha-toxin from Clostridium perfringensmediates calcium-dependent membrane recognition. Mol Microbiol 26:867–876 http://dx.doi.org/10.1046/j.1365-2958.1997.6161993.x.

38. Naylor CE, Jepson M, Crane DT, Titball RW, Miller J, Basak AK,Bolgiano B. 1999. Characterisation of the calcium-binding C-terminaldomain of Clostridium perfringens alpha-toxin. J Mol Biol 294:757–770http://dx.doi.org/10.1006/jmbi.1999.3279.

39. Nagahama M, Nakayama T, Kobayashi K, Sakurai J. 1995. Site-directed mutagenesis studies on the zinc-binding domain of Clostridiumperfringens alpha toxin. Jpn J Med Sci Biol 48:265–266.

40. Nagahama M, Michiue K, Sakurai J. 1996. Membrane-damagingaction of Clostridium perfringens alpha-toxin on phospholipid liposomes.Biochim Biophys Acta 1280:120–126.

41. Vachieri SG, Clark GC, Alape-Girón A, Flores-Díaz M, Justin N,Naylor CE, Titball RW, Basak AK. 2010. Comparison of a nontoxicvariant of Clostridium perfringens α-toxin with the toxic wild-type strain.Acta Crystallogr D Biol Crystallogr 66:1067–1074 http://dx.doi.org/10.1107/S090744491003369X.42. Eaton JT, Naylor CE, Howells AM,Moss DS, Titball RW, Basak AK.2002. Crystal structure of the C. perfringens alpha-toxin with the activesite closed by a flexible loop region. J Mol Biol 319:275–281 http://dx.doi.org/10.1016/S0022-2836(02)00290-5.43. Clark GC, Briggs DC, Karasawa T, Wang X, Cole AR, Maegawa T,Jayasekera PN, Naylor CE, Miller J, Moss DS, Nakamura S, Basak AK,Titball RW. 2003. Clostridium absonum alpha-toxin: new insights intoclostridial phospholipase C substrate binding and specificity. J Mol Biol333:759–769 http://dx.doi.org/10.1016/j.jmb.2003.07.016.44. NagahamaM, Otsuka A, Sakurai J. 2006. Role of tyrosine-57 and -65in membrane-damaging and sphingomyelinase activities of Clostridiumperfringens alpha-toxin. Biochim Biophys Acta 1762:110–114 http://dx.doi.org/10.1016/j.bbadis.2005.10.002.45. Alape-Girón A, Flores-Díaz M, Guillouard I, Naylor CE, TitballRW, Rucavado A, Lomonte B, Basak AK, Gutiérrez JM, Cole ST,Thelestam M. 2000. Identification of residues critical for toxicity inClostridium perfringens phospholipase C, the key toxin in gas gangrene.Eur J Biochem 267:5191–5197 http://dx.doi.org/10.1046/j.1432-1327.2000.01588.x.46. Urbina P, Flores-Díaz M, Alape-Girón A, Alonso A, Goñi FM. 2011.Effects of bilayer composition and physical properties on the phospholi-pase C and sphingomyelinase activities of Clostridium perfringens α-toxin.Biochim Biophys Acta 1808:279–286 http://dx.doi.org/10.1016/j.bbamem.2010.08.011.47. Nagahama M, Otsuka A, Oda M, Singh RK, Ziora ZM, Imagawa H,Nishizawa M, Sakurai J. 2007. Effect of unsaturated bonds in the sn-2acyl chain of phosphatidylcholine on the membrane-damaging actionof Clostridium perfringens alpha-toxin toward liposomes. Biochim Bio-phys Acta 1768:2940–2945 http://dx.doi.org/10.1016/j.bbamem.2007.08.016.48. Moe PC, Heuck AP. 2010. Phospholipid hydrolysis caused by Clos-tridium perfringens α-toxin facilitates the targeting of perfringolysin O tomembrane bilayers.Biochemistry 49:9498–9507 http://dx.doi.org/10.1021/bi1013886.

49. Fujii Y, Sakurai J. 1989. Contraction of the rat isolated aorta causedby Clostridium perfringens alpha toxin (phospholipase C): evidence forthe involvement of arachidonic acid metabolism. Br J Pharmacol 97:119–124 http://dx.doi.org/10.1111/j.1476-5381.1989.tb11931.x.

50. Sakurai J, Fujii Y, Shirotani M. 1990. Contraction induced by Clos-tridium perfringens alpha toxin in the isolated rat ileum. Toxicon 28:411–418 http://dx.doi.org/10.1016/0041-0101(90)90079-M.

51. Ochi S, Miyawaki T, Matsuda H, Oda M, Nagahama M, Sakurai J.2002. Clostridium perfringens alpha-toxin induces rabbit neutrophil ad-hesion. Microbiology 148:237–245 http://dx.doi.org/10.1099/00221287-148-1-237.

52. Oda M, Ikari S, Matsuno T, Morimune Y, Nagahama M, Sakurai J.2006. Signal transduction mechanism involved in Clostridium perfringensalpha-toxin-induced superoxide anion generation in rabbit neutrophils.Infect Immun 74:2876–2886 http://dx.doi.org/10.1128/IAI.74.5.2876-2886.2006.

53. Oda M, Kabura M, Takagishi T, Suzue A, Tominaga K, Urano S,Nagahama M, Kobayashi K, Furukawa K, Furukawa K, Sakurai J. 2012.Clostridium perfringens alpha-toxin recognizes the GM1a-TrkA com-plex. J Biol Chem 287:33070–33079 http://dx.doi.org/10.1074/jbc.M112.393801.

54. Sakurai J, Ochi S, Tanaka H. 1993. Evidence for coupling of Clos-tridium perfringens alpha-toxin-induced hemolysis to stimulated phospha-tidic acid formation in rabbit erythrocytes. Infect Immun 61:3711–3718.

55. Sakurai J, Ochi S, Tanaka H. 1994. Regulation of Clostridiumperfringens alpha-toxin-activated phospholipase C in rabbit erythrocytemembranes. Infect Immun 62:717–721.56. Ochi S, Oda M, Matsuda H, Ikari S, Sakurai J. 2004. Clostridiumperfringens alpha-toxin activates the sphingomyelin metabolism systemin sheep erythrocytes. J Biol Chem 279:12181–12189 http://dx.doi.org/10.1074/jbc.M307046200.

57. Oda M, Matsuno T, Shiihara R, Ochi S, Yamauchi R, Saito Y,Imagawa H, Nagahama M, Nishizawa M, Sakurai J. 2008. The rela-tionship between the metabolism of sphingomyelin species and the he-molysis of sheep erythrocytes induced by Clostridium perfringens alpha-toxin. J Lipid Res 49:1039–1047 http://dx.doi.org/10.1194/jlr.M700587-JLR200.

58. Tsukamoto K, Kozai Y, Ihara H, Kohda T, Mukamoto M, Tsuji T,Kozaki S. 2008. Identification of the receptor-binding sites in the carboxyl-terminal half of the heavy chain of botulinum neurotoxin types C and D.Microb Pathog 44:484–493 http://dx.doi.org/10.1016/j.micpath.2007.12.003.

59. OdaM, Shiihara R, Ohmae Y, Kabura M, Takagishi T, Kobayashi K,Nagahama M, Inoue M, Abe T, Setsu K, Sakurai J. 2012. Clostridiumperfringens alpha-toxin induces the release of IL-8 through a dual path-way via TrkA in A549 cells. Biochim Biophys Acta 1822:1581–1589http://dx.doi.org/10.1016/j.bbadis.2012.06.007.

60. Yamazaki Y, Horibata Y, Nagatsuka Y, Hirabayashi Y, HashikawaT. 2007. Fucoganglioside alpha-fucosyl(alpha-galactosyl)-GM1: anovel member of lipid membrane microdomain components involved inPC12 cell neuritogenesis. Biochem J 407:31–40 http://dx.doi.org/10.1042/BJ20070090.

61. Ichikawa N, Iwabuchi K, Kurihara H, Ishii K, Kobayashi T, Sasaki T,Hattori N,Mizuno Y, Hozumi K, Yamada Y, Arikawa-Hirasawa E. 2009.Binding of laminin-1 to monosialoganglioside GM1 in lipid rafts is crucialfor neurite outgrowth. J Cell Sci 122:289–299 http://dx.doi.org/10.1242/jcs.030338.62. Takagishi T, Oda M, Kabura M, Kurosawa M, Tominaga K, UranoS, Ueda Y, Kobayashi K, Kobayashi T, Sakurai J, Terao Y, NagahamaM.2015. Clostridium perfringens alpha-toxin induces Gm1a clustering andTrka phosphorylation in the host cell membrane. PLoS One 10:e0120497http://dx.doi.org/10.1371/journal.pone.0120497.63. Ueda Y, Makino A, Murase-Tamada K, Sakai S, Inaba T, Hullin-Matsuda F, Kobayashi T. 2013. Sphingomyelin regulates the transbilayermovement of diacylglycerol in the plasma membrane of Madin-Darbycanine kidney cells. FASEB J 27:3284–3297 http://dx.doi.org/10.1096/fj.12-226548.64. Oda M, Terao Y, Sakurai J, NagahamaM. 2015. Membrane-bindingmechanism of Clostridium perfringens alpha-toxin. Toxins (Basel) 7:5268–5275 http://dx.doi.org/10.3390/toxins7124880.65. Flores-Díaz M, Alape-Girón A, Clark G, Catimel B, Hirabayashi Y,Nice E, Gutiérrez JM, Titball R, ThelestamM. 2005. A cellular deficiencyof gangliosides causes hypersensitivity to Clostridium perfringens phos-pholipase C. J Biol Chem 280:26680–26689 http://dx.doi.org/10.1074/jbc.M500278200.

14 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 15: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

66. Monturiol-Gross L, Flores-Díaz M, Pineda-Padilla MJ, Castro-CastroAC, Alape-Giron A. 2014. Clostridium perfringens phospholipase C in-duced ROS production and cytotoxicity require PKC, MEK1 and NFκBactivation. PLoS One 9:e86475 http://dx.doi.org/10.1371/journal.pone.0086475.

67. Flores-Díaz M, Monturiol-Gross L, Naylor C, Alape-Girón A, FliegerA. 2016. Bacterial sphingomyelinases and phospholipases as virulencefactors. Microbiol Mol Biol Rev 80:597–628 http://dx.doi.org/10.1128/MMBR.00082-15.

68. Andrews NW, Almeida PE, Corrotte M. 2014. Damage control: cel-lular mechanisms of plasma membrane repair. Trends Cell Biol 24:734–742 http://dx.doi.org/10.1016/j.tcb.2014.07.008.

69. Amulic B, Cazalet C, Hayes GL, Metzler KD, Zychlinsky A. 2012.Neutrophil function: from mechanisms to disease. Annu Rev Immunol30:459–489 http://dx.doi.org/10.1146/annurev-immunol-020711-074942.70. Dale DC, Boxer L, Liles WC. 2008. The phagocytes: neutrophils andmonocytes. Blood 112:935–945 http://dx.doi.org/10.1182/blood-2007-12-077917.71. Manz MG, Boettcher S. 2014. Emergency granulopoiesis. Nat RevImmunol 14:302–314 http://dx.doi.org/10.1038/nri3660.72. Wirths S, Bugl S, Kopp HG. 2014. Neutrophil homeostasis and itsregulation by danger signaling. Blood 123:3563–3566 http://dx.doi.org/10.1182/blood-2013-11-516260.73. Nauseef WM, Borregaard N. 2014. Neutrophils at work. NatImmunol 15:602–611 http://dx.doi.org/10.1038/ni.2921.74. Martinon F, Tschopp J. 2005. NLRs join TLRs as innate sensors ofpathogens. Trends Immunol 26:447–454 http://dx.doi.org/10.1016/j.it.2005.06.004.75. Hajishengallis G, Lambris JD. 2011. Microbial manipulation ofreceptor crosstalk in innate immunity. Nat Rev Immunol 11:187–200http://dx.doi.org/10.1038/nri2918.76. Beutler B. 2000. Tlr4: central component of the sole mammalian LPSsensor. Curr Opin Immunol 12:20–26 http://dx.doi.org/10.1016/S0952-7915(99)00046-1.77. Texereau J, Chiche JD, Taylor W, Choukroun G, Comba B, Mira JP.2005. The importance of Toll-like receptor 2 polymorphisms in severeinfections. Clin Infect Dis 41(Suppl 7):S408–S415 http://dx.doi.org/10.1086/431990.78. Dziarski R. 2003. Recognition of bacterial peptidoglycan by the innateimmune system. Cell Mol Life Sci 60:1793–1804 http://dx.doi.org/10.1007/s00018-003-3019-6.79. Takeuchi O, Hoshino K, Kawai T, Sanjo H, Takada H, Ogawa T,Takeda K, Akira S. 1999. Differential roles of TLR2 and TLR4 in rec-ognition of Gram-negative and Gram-positive bacterial cell wall com-ponents. Immunity 11:443–451 http://dx.doi.org/10.1016/S1074-7613(00)80119-3.80. Demetri GD, Griffin JD. 1991. Granulocyte colony-stimulating factorand its receptor. Blood 78:2791–2808.

81. Boettcher S, Gerosa RC, Radpour R, Bauer J, Ampenberger F,Heikenwalder M, Kopf M, Manz MG. 2014. Endothelial cells translatepathogen signals into G-CSF-driven emergency granulopoiesis. Blood124:1393–1403 http://dx.doi.org/10.1182/blood-2014-04-570762.

82. Lieschke GJ, Grail D, Hodgson G, Metcalf D, Stanley E, Cheers C,Fowler KJ, Basu S, Zhan YF, Dunn AR. 1994. Mice lacking granulocytecolony-stimulating factor have chronic neutropenia, granulocyte andmacrophage progenitor cell deficiency, and impaired neutrophil mobili-zation. Blood 84:1737–1746.

83. Liu F, Wu HY, Wesselschmidt R, Kornaga T, Link DC. 1996. Im-paired production and increased apoptosis of neutrophils in granulocytecolony-stimulating factor receptor-deficient mice. Immunity 5:491–501http://dx.doi.org/10.1016/S1074-7613(00)80504-X.

84. Bryant AE. 2003. Biology and pathogenesis of thrombosis andprocoagulant activity in invasive infections caused by group A streptococci

and Clostridium perfringens. Clin Microbiol Rev 16:451–462 http://dx.doi.org/10.1128/CMR.16.3.451-462.2003.85. McNee JW, Dunn JS. 1917. The method of spread of gas gangreneinto living tissue. BMJ 1:4–729 http://dx.doi.org/10.1136/bmj.1.2944.726.86. Bryant AE, Chen RY, Nagata Y, Wang Y, Lee CH, Finegold S,Guth PH, Stevens DL. 2000. Clostridial gas gangrene. II. PhospholipaseC-induced activation of platelet gpIIbIIIa mediates vascular occlusionand myonecrosis in Clostridium perfringens gas gangrene. J Infect Dis182:808–815 http://dx.doi.org/10.1086/315757.87. Bryant AE, Chen RY, Nagata Y, Wang Y, Lee CH, Finegold S,Guth PH, Stevens DL. 2000. Clostridial gas gangrene. I. Cellular andmolecular mechanisms of microvascular dysfunction induced by exo-toxins of Clostridium perfringens. J Infect Dis 182:799–807 http://dx.doi.org/10.1086/315756.88. Hickey MJ, Kwan RY, Awad MM, Kennedy CL, Young LF, Hall P,Cordner LM, Lyras D, Emmins JJ, Rood JI. 2008. Molecular and cellu-lar basis of microvascular perfusion deficits induced by Clostridiumperfringens and Clostridium septicum. PLoS Pathog 4:e1000045 http://dx.doi.org/10.1371/journal.ppat.1000045.89. Bryant AE, Bayer CR, Aldape MJ, Wallace RJ, Titball RW, StevensDL. 2006. Clostridium perfringens phospholipase C-induced platelet/leukocyte interactions impede neutrophil diapedesis. J Med Microbiol55:495–504 http://dx.doi.org/10.1099/jmm.0.46390-0.90. Takehara M, Takagishi T, Seike S, Ohtani K, Kobayashi K,Miyamoto K, Shimizu T, Nagahama M. 2016. Clostridium perfringensalpha-toxin impairs innate immunity via inhibition of neutrophil differ-entiation. Sci Rep 6:28192 http://dx.doi.org/10.1038/srep28192.91. Basu S, Hodgson G, Katz M, Dunn AR. 2002. Evaluation of role ofG-CSF in the production, survival, and release of neutrophils from bonemarrow into circulation. Blood 100:854–861 http://dx.doi.org/10.1182/blood.V100.3.854.92. Diaz O, Mébarek-Azzam S, Benzaria A, Dubois M, Lagarde M,Némoz G, Prigent AF. 2005. Disruption of lipid rafts stimulates phos-pholipase D activity in human lymphocytes: implication in the regulationof immune function. J Immunol 175:8077–8086 http://dx.doi.org/10.4049/jimmunol.175.12.8077.93. OstromRS, Insel PA. 2004. The evolving role of lipid rafts and caveolaein G protein-coupled receptor signaling: implications for molecular phar-macology. Br J Pharmacol 143:235–245 http://dx.doi.org/10.1038/sj.bjp.0705930.94. Simons K, Toomre D. 2000. Lipid rafts and signal transduction. NatRev Mol Cell Biol 1:31–39 http://dx.doi.org/10.1038/35036052.95. Kolesnick RN, Krönke M. 1998. Regulation of ceramide productionand apoptosis. Annu Rev Physiol 60:643–665 http://dx.doi.org/10.1146/annurev.physiol.60.1.643.96. Ochi S, Oda M, Nagahama M, Sakurai J. 2003. Clostridiumperfringens alpha-toxin-induced hemolysis of horse erythrocytes is de-pendent on Ca2+ uptake. Biochim Biophys Acta 1613:79–86 http://dx.doi.org/10.1016/S0005-2736(03)00140-8.97. Chang HH, Wang TP, Chen PK, Lin YY, Liao CH, Lin TK,Chiang YW, Lin WB, Chiang CY, Kau JH, Huang HH, Hsu HL, LiaoCY, Sun DS. 2013. Erythropoiesis suppression is associated with anthraxlethal toxin-mediated pathogenic progression. PLoS One 8:e71718 http://dx.doi.org/10.1371/journal.pone.0071718.98. Takagishi T, Takehara M, Seike S, Miyamoto K, Kobayashi K,Nagahama M. 2017. Clostridium perfringens α-toxin impairs erythro-poiesis by inhibition of erythroid differentiation. Sci Rep 7:5217 http://dx.doi.org/10.1038/s41598-017-05567-8.99. Simon TG, Bradley J, Jones A, Carino G. 2014. Massive intravascularhemolysis from Clostridium perfringens septicemia: a review. J IntensiveCare Med 29:327–333 http://dx.doi.org/10.1177/0885066613498043.100. Konstantinidis DG, Pushkaran S, Johnson JF, Cancelas JA,Manganaris S, Harris CE, Williams DA, Zheng Y, Kalfa TA. 2012. Sig-

ASMscience.org/MicrobiolSpectrum 15

Histotoxic Clostridial Infections

Page 16: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

naling and cytoskeletal requirements in erythroblast enucleation. Blood119:6118–6127 http://dx.doi.org/10.1182/blood-2011-09-379263.101. Verherstraeten S, Goossens E, Valgaeren B, Pardon B, TimbermontL, Haesebrouck F, Ducatelle R, Deprez P, Wade KR, Tweten R, VanImmerseel F. 2015. Perfringolysin O: the underrated Clostridium per-fringens toxin? Toxins (Basel) 7:1702–1721 http://dx.doi.org/10.3390/toxins7051702.102. Popoff MR. 2014. Clostridial pore-forming toxins: powerful viru-lence factors. Anaerobe 30:220–238 http://dx.doi.org/10.1016/j.anaerobe.2014.05.014.103. Katayama S, Dupuy B, Garnier T, Cole ST. 1995. Rapid expansionof the physical and genetic map of the chromosome of ClostridiumperfringensCPN50. J Bacteriol 177:5680–5685 http://dx.doi.org/10.1128/jb.177.19.5680-5685.1995.104. Shimizu T, Ohtani K, Hirakawa H, Ohshima K, Yamashita A, ShibaT, Ogasawara N, Hattori M, Kuhara S, Hayashi H. 2002. Completegenome sequence of Clostridium perfringens, an anaerobic flesh-eater.Proc Natl Acad Sci USA 99:996–1001 http://dx.doi.org/10.1073/pnas.022493799.105. Myers GS, Rasko DA, Cheung JK, Ravel J, Seshadri R, DeBoy RT,Ren Q, Varga J, Awad MM, Brinkac LM, Daugherty SC, Haft DH,Dodson RJ,Madupu R, NelsonWC, RosovitzMJ, Sullivan SA, Khouri H,Dimitrov GI, Watkins KL, Mulligan S, Benton J, Radune D, Fisher DJ,Atkins HS, Hiscox T, Jost BH, Billington SJ, Songer JG, McClane BA,Titball RW, Rood JI, Melville SB, Paulsen IT. 2006. Skewed genomicvariability in strains of the toxigenic bacterial pathogen, Clostridium per-fringens.GenomeRes 16:1031–1040 http://dx.doi.org/10.1101/gr.5238106.106. Hassan KA, Elbourne LD, Tetu SG,Melville SB, Rood JI, Paulsen IT.2014. Genomic analyses of Clostridium perfringens isolates from fivetoxinotypes. Res Microbiol 166:255–263.

107. Kiu R, Caim S, Alexander S, Pachori P, Hall LJ. 2017. Probinggenomic aspects of the multi-host pathogen Clostridium perfringensreveals significant pangenome diversity, and a diverse array of virulencefactors. Front Microbiol 8:2485 http://dx.doi.org/10.3389/fmicb.2017.02485.

108. Tweten RK. 1988. Nucleotide sequence of the gene for perfringolysinO (theta-toxin) from Clostridium perfringens: significant homology withthe genes for streptolysin O and pneumolysin. Infect Immun 56:3235–3240.

109. Rossjohn J, Feil SC, McKinstry WJ, Tweten RK, Parker MW. 1997.Structure of a cholesterol-binding, thiol-activated cytolysin and a modelof its membrane form. Cell 89:685–692 http://dx.doi.org/10.1016/S0092-8674(00)80251-2.

110. Shatursky O, Heuck AP, Shepard LA, Rossjohn J, Parker MW,Johnson AE, Tweten RK. 1999. The mechanism of membrane insertionfor a cholesterol-dependent cytolysin: a novel paradigm for pore-formingtoxins. Cell 99:293–299 http://dx.doi.org/10.1016/S0092-8674(00)81660-8.

111. Rosado CJ, Buckle AM, Law RH, Butcher RE, Kan WT, Bird CH,Ung K, Browne KA, Baran K, Bashtannyk-Puhalovich TA, Faux NG,Wong W, Porter CJ, Pike RN, Ellisdon AM, Pearce MC, Bottomley SP,Emsley J, Smith AI, Rossjohn J, Hartland EL, Voskoboinik I, Trapani JA,Bird PI, Dunstone MA, Whisstock JC. 2007. A common fold mediatesvertebrate defense and bacterial attack. Science 317:1548–1551 http://dx.doi.org/10.1126/science.1144706.

112. Tweten RK, Hotze EM, Wade KR. 2015. The unique molecularchoreography of giant pore formation by the cholesterol-dependentcytolysins of Gram-positive bacteria. Annu Rev Microbiol 69:323–340http://dx.doi.org/10.1146/annurev-micro-091014-104233.

113. Farrand AJ, LaChapelle S, Hotze EM, Johnson AE, Tweten RK.2010. Only two amino acids are essential for cytolytic toxin recognition ofcholesterol at the membrane surface. Proc Natl Acad Sci U S A 107:4341–4346 http://dx.doi.org/10.1073/pnas.0911581107.

114. Harris RW, Sims PJ, Tweten RK. 1991. Evidence that Clostridiumperfringens theta-toxin induces colloid-osmotic lysis of erythrocytes. In-fect Immun 59:2499–2501.115. Park JM, Ng VH, Maeda S, Rest RF, Karin M. 2004. Anthrolysin Oand other Gram-positive cytolysins are toll-like receptor 4 agonists. J ExpMed 200:1647–1655 http://dx.doi.org/10.1084/jem.20041215.116. Ribet D, Hamon M, Gouin E, Nahori MA, Impens F, Neyret-KahnH, Gevaert K, Vandekerckhove J, Dejean A, Cossart P. 2010. Listeriamonocytogenes impairs SUMOylation for efficient infection. Nature464:1192–1195 http://dx.doi.org/10.1038/nature08963.117. Asmuth DM, Olson RD, Hackett SP, Bryant AE, Tweten RK,Tso JY, Zollman T, Stevens DL. 1995. Effects of Clostridium perfringensrecombinant and crude phospholipase C and theta-toxin on rabbit he-modynamic parameters. J Infect Dis 172:1317–1323 http://dx.doi.org/10.1093/infdis/172.5.1317.118. Bryant AE, Bergstrom R, Zimmerman GA, Salyer JL, Hill HR,Tweten RK, Sato H, Stevens DL. 1993. Clostridium perfringens inva-siveness is enhanced by effects of theta toxin upon PMNL structure andfunction: the roles of leukocytotoxicity and expression of CD11/CD18adherence glycoprotein. FEMS Immunol Med Microbiol 7:321–336http://dx.doi.org/10.1111/j.1574-695X.1993.tb00414.x.119. Bryant AE, Stevens DL. 1996. Phospholipase C and perfringolysinO from Clostridium perfringens upregulate endothelial cell-leukocyteadherence molecule 1 and intercellular leukocyte adherence molecule 1expression and induce interleukin-8 synthesis in cultured human umbilicalvein endothelial cells. Infect Immun 64:358–362.120. Bunting M, Lorant DE, Bryant AE, Zimmerman GA, McIntyre TM,Stevens DL, Prescott SM. 1997. Alpha toxin fromClostridium perfringensinduces proinflammatory changes in endothelial cells. J Clin Invest 100:565–574 http://dx.doi.org/10.1172/JCI119566.121. Whately R, Zimmerman G, Stevens D, Parker C, McIntyre T,Prescott S. 1989. The regulation of platelet activating factor synthesisin endothelial cells: the role of calcium and protein kinase. J Biol Chem11:6325–6333.122. Bryant AE, Bayer CR, Hayes-Schroer SM, Stevens DL. 2003. Acti-vation of platelet gpIIbIIIa by phospholipase C from Clostridium per-fringens involves store-operated calcium entry. J Infect Dis 187:408–417http://dx.doi.org/10.1086/367964.123. Stevens DL, Mitten J, Henry C. 1987. Effects of α and θ toxins fromClostridium perfringens on human polymorphonuclear leukocytes. J In-fect Dis 156:324–333 http://dx.doi.org/10.1093/infdis/156.2.324.124. Stevens DL. 2000. The pathogenesis of clostridial myonecrosis. Int JMed Microbiol 290:497–502 http://dx.doi.org/10.1016/S1438-4221(00)80074-0.125. O’Brien DK, Melville SB. 2000. The anaerobic pathogen can escapethe phagosome of macrophages under aerobic conditions. Cell Microbiol2:505–519 http://dx.doi.org/10.1046/j.1462-5822.2000.00074.x.126. O’Brien DK, Melville SB. 2004. Effects of Clostridium perfringensalpha-toxin (PLC) and perfringolysin O (PFO) on cytotoxicity to macro-phages, on escape from the phagosomes of macrophages, and on persis-tence of C. perfringens in host tissues. Infect Immun 72:5204–5215 http://dx.doi.org/10.1128/IAI.72.9.5204-5215.2004.127. Stevens DL, Troyer BE, Merrick DT, Mitten JE, Olson RD. 1988.Lethal effects and cardiovascular effects of purified α- and θ-toxins fromClostridium perfringens. J Infect Dis 157:272–279 http://dx.doi.org/10.1093/infdis/157.2.272.128. Stevens DL, Bryant AE. 1997. Pathogenesis of Clostridium per-fringens infection: mechanisms and mediators of shock. Clin Infect Dis 25(Suppl 2):S160–S164 http://dx.doi.org/10.1086/516249.129. Knapp O, Maier E, Mkaddem SB, Benz R, Bens M, Chenal A, GenyB, Vandewalle A, Popoff MR. 2010. Clostridium septicum alpha-toxinforms pores and induces rapid cell necrosis. Toxicon 55:61–72 http://dx.doi.org/10.1016/j.toxicon.2009.06.037.

16 ASMscience.org/MicrobiolSpectrum

Nagahama et al.

Page 17: Histotoxic Clostridial Infections · Typical histotoxic clostridial diseases include human gas gangrene or myonecrosis and blackleg in ... subsequent infection at distal sites (3,

Downloaded from www.asmscience.org by

IP: 130.194.145.38

On: Fri, 06 Sep 2019 02:16:21

130. Dal Peraro M, van der Goot FG. 2016. Pore-forming toxins: ancient,but never really out of fashion.Nat Rev Microbiol 14:77–92 http://dx.doi.org/10.1038/nrmicro.2015.3.

131. Rossjohn J, Buckley JT, Hazes B, Murzin AG, Read RJ, Parker MW.1997. Aerolysin and pertussis toxin share a common receptor-bindingdomain. EMBO J 16:3426–3434 http://dx.doi.org/10.1093/emboj/16.12.3426.

132. Melton JA, Parker MW, Rossjohn J, Buckley JT, Tweten RK. 2004.The identification and structure of the membrane-spanning domain ofthe Clostridium septicum alpha toxin. J Biol Chem 279:14315–14322http://dx.doi.org/10.1074/jbc.M313758200.

133. Ballard J, Sokolov Y, Yuan W-L, Kagan BL, Tweten RK. 1993.Activation and mechanism of Clostridium septicum alpha toxin. MolMicrobiol 10:627–634 http://dx.doi.org/10.1111/j.1365-2958.1993.tb00934.x.

134. Sellman BR, Tweten RK. 1997. The propeptide of Clostridiumsepticum alpha toxin functions as an intramolecular chaperone and is apotent inhibitor of alpha toxin-dependent cytolysis. Mol Microbiol 25:429–440 http://dx.doi.org/10.1046/j.1365-2958.1997.4541820.x.

135. Gordon VM, Nelson KL, Buckley JT, Stevens VL, Tweten RK,Elwood PC, Leppla SH. 1999. Clostridium septicum alpha toxin usesglycosylphosphatidylinositol-anchored protein receptors. J Biol Chem274:27274–27280 http://dx.doi.org/10.1074/jbc.274.38.27274.136. Shin DJ, Lee JJ, Choy HE, Hong Y. 2004. Generation and charac-terization of Clostridium septicum alpha toxin mutants and their use indiagnosing paroxysmal nocturnal hemoglobinuria. Biochem Biophys ResCommun 324:753–760 http://dx.doi.org/10.1016/j.bbrc.2004.09.104.

137. Melton-Witt JA, Bentsen LM, Tweten RK. 2006. Identification offunctional domains of Clostridium septicum alpha toxin. Biochemistry45:14347–14354 http://dx.doi.org/10.1021/bi061334p.

138. Mukamoto M, Kimura R, Hang’ombe MB, Kohda T, Kozaki S.2013. Analysis of tryptophan-rich region in Clostridium septicum alpha-toxin involved with binding to glycosylphosphatidylinositol-anchoredproteins.Microbiol Immunol 57:163–169 http://dx.doi.org/10.1111/1348-0421.12017.

139. Kennedy CL, Lyras D, Cordner LM, Melton-Witt J, Emmins JJ,Tweten RK, Rood JI. 2009. Pore-forming activity of alpha-toxin isessential for Clostridium septicum-mediated myonecrosis. Infect Immun77:943–951 http://dx.doi.org/10.1128/IAI.01267-08.

140. Kennedy CL, Smith DJ, Lyras D, Chakravorty A, Rood JI. 2009.Programmed cellular necrosis mediated by the pore-forming alpha-toxinfrom Clostridium septicum. PLoS Pathog 5:e1000516 http://dx.doi.org/10.1371/journal.ppat.1000516.

141. Hang’ombe MB, Kohda T, Mukamoto M, Kozaki S. 2005. Rela-tionship between Clostridium septicum alpha-toxin activity and bindingto erythrocyte membranes. J Vet Med Sci 67:69–74 http://dx.doi.org/10.1292/jvms.67.69.

142. Hang’ombe MB, Mukamoto M, Kohda T, Sugimoto N, Kozaki S.2004. Cytotoxicity of Clostridium septicum alpha-toxin: its oligomeriza-tion in detergent resistant membranes of mammalian cells.Microb Pathog37:279–286 http://dx.doi.org/10.1016/j.micpath.2004.09.001.

143. Chakravorty A, Awad MM, Cheung JK, Hiscox TJ, Lyras D,Rood JI. 2015. The pore-forming α-toxin from Clostridium septicumactivates the MAPK pathway in a Ras-c-Raf-dependent and independentmanner. Toxins (Basel) 7:516–534 http://dx.doi.org/10.3390/toxins7020516.

144. Stevens DL, Bisno AL, Chambers HF, Dellinger EP, Goldstein EJ,Gorbach SL, Hirschmann JV, Kaplan SL, Montoya JG, Wade JC, Infec-tious Diseases Society of America. 2014. Practice guidelines for the diag-nosis and management of skin and soft tissue infections: 2014 update bythe Infectious Diseases Society of America. Clin Infect Dis 59:e10–e52http://dx.doi.org/10.1093/cid/ciu296.145. Stevens DL, Maier KA, Laine BM, Mitten JE. 1987. Comparisonof clindamycin, rifampin, tetracycline, metronidazole, and penicillin forefficacy in prevention of experimental gas gangrene due to Clostridiumperfringens. J Infect Dis 155:220–228 http://dx.doi.org/10.1093/infdis/155.2.220.146. Stevens DL, Laine BM, Mitten JE. 1987. Comparison of singleand combination antimicrobial agents for prevention of experimentalgas gangrene caused by Clostridium perfringens. Antimicrob AgentsChemother 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,Higashihara M, Nagahama M, Koide N, Yokochi T, Sakurai J. 2008.Effect of erythromycin on biological activities induced by Clostridiumperfringens α-toxin. J Pharmacol Exp Ther 327:934–940 http://dx.doi.org/10.1124/jpet.108.143677.148. Chakravorty A, AwadMM, Hiscox TJ, Cheung JK, Choo JM, LyrasD, Rood JI. 2014. Opioid analgesics stop the development of clostridialgas gangrene. J Infect Dis 210:483–492 http://dx.doi.org/10.1093/infdis/jiu101.149. Stevens DL, Titball RW, Jepson M, Bayer CR, Hayes-Schroer SM,Bryant AE. 2004. Immunization with the C-domain of alpha-toxinprevents lethal infection, localizes tissue injury, and promotes host re-sponse to challenge with Clostridium perfringens. J Infect Dis 190:767–773 http://dx.doi.org/10.1086/422691.150. NagahamaM, OdaM, Kobayashi K, Ochi S, Takagishi T, ShibutaniM, Sakurai J. 2013. A recombinant carboxy-terminal domain of alpha-toxin protects mice against Clostridium perfringens. Microbiol Immunol57:340–345 http://dx.doi.org/10.1111/1348-0421.12036.151. Aronoff DM. 2013. Clostridium novyi, sordellii, and tetani:mecha-nisms of disease. Anaerobe 24:98–101 http://dx.doi.org/10.1016/j.anaerobe.2013.08.009.152. Jóźwiak J, Grzela T, Jankowska-Steifer E, Komar A, Moskalewski S,Martirosian G. 2007. Lethal factor of Clostridium histolyticum kills cellsby apoptosis. FEMS Immunol Med Microbiol 49:296–303 http://dx.doi.org/10.1111/j.1574-695X.2006.00204.x.153. Genth H, Pauillac S, Schelle I, Bouvet P, Bouchier C, Varela-ChavezC, Just I, Popoff MR. 2014. Haemorrhagic toxin and lethal toxin fromClostridium sordellii strain vpi9048: molecular characterization andcomparative analysis of substrate specificity of the large clostridialglucosylating toxins. Cell Microbiol 16:1706–1721 http://dx.doi.org/10.1111/cmi.12321.154. Hausmann R, Albert F, Geissdörfer W, Betz P. 2004. Clostridiumfallax associated with sudden death in a 16-year-old boy. J MedMicrobiol53:581–583 http://dx.doi.org/10.1099/jmm.0.05495-0.155. Biswas R, K D, Sistla S, Chandra Sistla S, Amaranathan A. 2017.Clostridium perfringens: a bacterial pathogen gaining recognition innecrotizing pancreatitis. Anaerobe 47:111–114 http://dx.doi.org/10.1016/j.anaerobe.2017.05.006.156. Rychener L, InAlbon S, Djordjevic SP, Chowdhury PR, Ziech RE,de Vargas AC, Frey J, Falquet L. 2017. Clostridium chauvoei, an evolu-tionary dead-end pathogen. Front Microbiol 8:1054 http://dx.doi.org/10.3389/fmicb.2017.01054.

ASMscience.org/MicrobiolSpectrum 17

Histotoxic Clostridial Infections