ABSTRACT
Cysteine proteases are hydrolases that share a common catalytic mechanism involving a nucleophilic cysteine thiol in a catalytic dyad or triad. Here, we review the current clans that make up the cysteine proteases in the MEROPS database as of March 2025. We also discuss cysteine proteases made by C. difficile, with a particular focus on recent analysis of the sporulation-specific protease, YabG, that supports its reclassification into clan CD of the MEROPS protease database. YabG is a highly conserved sporulation-specific protease that, until more recently, has been mostly studied in B. subtilis, where YabG is important for processing coat proteins. In C. difficile, YabG processes proteins required for spore germination and is important in coat/exosporium protein expression.
KEYWORDS: spore, protease, Bacillus subtilis, Clostridioides difficile, MEROPS
INTRODUCTION
The MEROPS database is a database of peptidase classification, where proteases are grouped into one of 66 clans. After clan classification, the proteases remain unassigned or are further classified into one of 281 families (1). Proteases classified into families are defined as having an evolutionary relationship based on similar tertiary structure, primary sequence order of catalytic residues if the structure is not available, or common sequence motifs near the catalytic residues. Based on catalytic mechanisms, proteases are classified into six distinct classes: aspartic, glutamic, metallo, cysteine, serine, and threonine proteases.
In the MEROPS database, there are 1,071 cysteine proteases that remain unclassified or are classified into one of 11 clans, where C denotes it is a cysteine protease; CA, CD, CE, CF, CL, CM, CN, CO, CP, CQ, and CR (1). Among all the cysteine protease clans, there are 101 families. Some families are further divided into subfamilies due to evidence of divergence within the family.
Cysteine proteases are found in all domains of life, and many are synthesized as inactive precursors, or zymogens, where a propeptide or prodomain blocks their catalytic site and prevents unwanted protein degradation until the zymogen is activated. This prodomain can also have other functions, such as facilitating protein folding or intracellular trafficking of the peptide (2, 3). Activation of the zymogen requires removal of the prodomain, which can be accomplished by autocatalytic processing, trans-processing, or changes in pH (4, 5). Many cysteine proteases contain a Cys-His-Asn/Asp catalytic triad, while others contain a Cys-His catalytic dyad. For catalysis, the histidine residue serves as a proton donor to the cysteine. This allows the thiol of the active site cysteine to act as a nucleophile, which then attacks the carbon present in the carboxyl group next to the peptide bond in the target protein (4, 6, 7). Cysteine proteases that are in the same clan contain the same sequence motif of catalytic residues. Founding members of the clan are the basis for which other proteases are compared. As new evidence arises (such as crystal structures and predictive modeling), proteases can be moved to new families within the clans or new clans generated that are more representative of their characteristics.
CLAN CA
Clan CA contains 46 families of cysteine proteases and is the largest of the cysteine clans (1). All of these enzymes have a Cys, His, Asn (or Asp) catalytic motif and have structural similarity to papain (8). Papain is a protease extracted from the papaya plant, Carica papaya, and was one of the first enzymes characterized and structure determined (8). Papain-like proteases contain a Cys and His catalytic dyad and two other important residues: the Gln that precedes the catalytic Cys and the Asn/Asp that follows the catalytic His residue. Papain has broad substrate specificity, although it has a preference for arginine at the P1 position (the amino acid immediately N-terminal to the scissile bond) and Val/Phe/Thr at the P2 position (the amino acid residue two positions upstream of the scissile bond) (9). Members of this clan mostly consist of endopeptidases, though there are several exopeptidases found in a variety of organisms. In C. difficile, the Cwp84 cysteine protease is a member of the clan CA and the C1 family (10–12).
CLAN CD
The protease clan CD is comp0sed of eight families, where the families resemble the founding member, caspase-1. These include C11 (clostripain), C123 (RARP2), C13 (legumain), C14 (caspase-1), C25 (gingipain RgpA), C50 (separase), C80 (RTX self-cleaving toxin), and C84 (prtH peptidase) (1).
Members of clan CD are site-specific endopeptidases whose substrate specificity is driven by the identity of the P1 residue. Family C11 (clostripain) specifically cleaves after Arg residues; family C13 (legumain) specifically cleaves after Asp or Asn residue; family C14 (caspases) cleaves specifically after aspartate, although the metacaspases cleave after basic residues (Arg or Lys); and family 25 (gingipain) also cleaves after Arg or Lys (1, 13–20).
Caspase-1, a member of the C14 family, has a His-Cys catalytic dyad. This protease family specificity is dependent upon the specific caspase. For example, members of the subfamily C14A (caspases) have strict specificity for aspartyl bonds, whereas members of the C14B (metacaspases and paracaspases) subfamily have strict specificity for arginyl or lysyl bonds (21–23). Caspases are commonly held inactive as a zymogen and are activated upon apoptosis by cleavage on the carboxyl side of the specific aspartic acid residues, resulting in the alpha and beta subunits (24). Because many have functions in homeostasis, regulating inflammation, or cell death, their activation is tightly controlled (25, 26).
Clostripain was first identified in the endospore-forming Hathewaya histolyticum (formerly Clostridium histolyticum [27]), and a homolog is also found in Clostridium perfringens (28–30). Members of the C11 family have strict specificity for hydrolysis of arginyl bonds, with a conserved active site motif -His-Gly-(Xaa)n-Ala-Cys (31). Though clostripain is a major extracellular protease for H. histolyticum, it is non-essential for disease. The C. perfringens clostripain-like protease is reported to induce uptake of apoptotic neutrophils by macrophage phagocytosis (28). This protease is typically used in the laboratory as an effective arginine-specific peptidase for protein sequencing techniques like mass spectrophotometry and human islet isolation (32).
The RTX-self-cleaving toxin from Vibrio cholerae is a member of the C80 family that contains bacterial toxins that self-cleave via a cysteine peptidase mechanism and classified as a multifunctional autoprocessing repeats-in-toxin (MARTX) (33, 34). MARTX toxins are characterized by having glycine-rich repeat regions I the N- and C-termini and are secreted by the type I secretion system (T1SS), followed by a cysteine protease domain (CPD), where His and Cys make up the catalytic dyad (34–38). The glycine-rich repeats are hypothesized to form a pore within the host cell membrane (36, 39). The cysteine protease domain (CPD) is activated by the binding of host inositol hexakisphosphate (InsP6), resulting in autoprocessing after Leu2447, Leu3099, and Leu3441 releasing the Rho inactivation domain (RID) and α/β hydrolase (ABH) from the large holotoxin (34). The actin-crosslinking domain (ACD) and CPD remain attached to the N- and C-terminal membrane-bound regions (34, 39–43). The activation of ACD results in actin crosslinking leading to the depolarization of cells (44–46). Release of RID inactivates small Rho protein GTPases, blocking interaction with downstream effectors (47). In 2008, the tertiary structure of the V. cholerae RTX toxin was determined and, because of the structural similarities to caspase-1 (C14 family), was included in the clan CD (42).
Other members of this family are the Clostridial toxins TcdA and TcdB that are similar to MARTX family toxins. These large glucosylating toxins are the drive for the primary symptoms associated with C. difficile infection (CDI), severe diarrhea (48, 49). TcdA and TcdB contain a catalytic dyad composed of Asp, His, and Cys that are discussed more in depth later in this review (39, 50, 51).
CLAN CE
Clan CE proteases are adenain-like proteases. Proteases in this clan are endopeptidases, and their catalysis is thought to be similar to papain (containing a His-Cys catalytic dyad). There are seven families in this clan: C5 (adenain), C122 (SdeA), C48 (Ulp1 peptidase), C55 (YopJ), C57 (vaccinia virus I7L processing peptidase), C63 (African swine fever virus processing peptidase), and C79 (ElaD peptidase) (1). Clan CE members are placed within this clan based upon similarity to adenain, an adenovirus endopeptidase, that is required for virion uncoating, maturation, and release from the infected cell (52–54). Members of this family possess active site residues in the order of His, Asp / Asn, Gln, and Cys (1).
CLAN CF
Clan CF only has one family, C15, that contains the Bacillus amyloliquefaciens pyroglutamyl-peptidase I. This family contains omega peptidases (peptidases that cleave non-standard peptide bonds) that remove terminal residue that are substituted, cyclized, or linked to isopeptide bonds (55). Peptides carrying information, such as hormones, evade aminopeptidase degradation through the presence of an N-terminal pyroglutamyl (pGlu) residue (e.g., the human hormone, thyrotropin-releasing hormone [TRH], which is produced in the hypothalamus and stimulates the pituitary gland) (56, 57). Members of the C15 family release an N-terminal pyroglutamate residue and have a catalytic triad of Glu, Cys, His (57).
CLAN CL
Clan CL contains the C60 and C82 families, which contain sortase A and L,D-transpeptidase, respectively. Members of this clan are involved in bacterial cell wall hydrolysis and possess a His, Cys catalytic dyad. Sortases are transpeptidases found in the cell envelope of many Gram-positive bacteria, where they anchor surface proteins to the peptidoglycan cross bridge of the cell wall. Other sortases are involved in pilus assembly (58–60). L,D-transpeptidases catalyze the formation of 3-3 peptidoglycan cross-links in Gram-positive bacteria (61, 62).
CLAN CM
Clan CM is composed of the C18 family, where the founding member is the hepatitis C virus peptidase 2 (also known as NS2/3). NS2/3 is a viral polyprotein endopeptidase and is an active homodimer. Each monomer contains catalytic residues His, Glu, and Cys. His and Glu from one monomer interact with Cys from the other monomer, and vice versa. The NS2/3 protease cleaves the HCV polyprotein between non-structural proteins NS2 and NS3, generating mature viral proteins (63, 64).
CLAN CN
Clan CN only has one family, C9, that is made up of the Sindbis virus-type nsP2 peptidase. This peptidase has a Cys His catalytic dyad and is responsible for cleaving the non-structural proteins of the Sindbis virus (65). Originally believed to be a papain-like protease due to the similar catalytic dyad, the tertiary structure is unrelated and was thus the founding member of clan CN (66).
CLAN CO
Clan CO is composed of the C40 family of dipeptidyl-peptidases. Members of this family are involved in cell-wall modification, e.g., dipeptidyl-peptidase VI (DPP VI), a cytoplasmic peptidase of Bacillus sphaericus. During sporulation of B. sphaericus, DPP VI hydrolyzes γ-D-Glu-DAP(Lys) linkages in peptides that have a free N-terminal L-alanine (67). In the closely related B. subtilis, the cell wall lytic enzymes LytF, LytE, and CwlS are involved in cell wall separation (68–70). Members of this family have conserved cysteine and histidine residues that form a catalytic triad of Cys, His, His (71).
CLAN CP
Clan CP is made up of the C97 family, deSUMOylating isopeptidase 1. SUMO (small ubiquitin-like modifier) deconjugating enzyme is a post-translational protein modification pathway in eukaryotes (72). The reversible attachment of SUMO is essential for correct activity, localization, and/or interactions with targets. Members possess a His, Cys catalytic dyad, the reverse of what is observed in papain-like proteases.
CLAN CQ
The only member of CQ is the C53 family made up of pestivirus Npro peptidase. Family members are endopeptidases that process the Pestivirus p20 polyprotein and contain a His, Cys catalytic dyad (73). This autoprotease is located at the N-terminus of the polyprotein and releases the core protein using autolysis (74). Cleavage is observed between a Cys and Ser but can also follow Ala or Ser at the P1 site (74). The release of Npro blocks host interferon response by inducing degradation of interferon regulatory factor-3 (75, 76).
CLAN CR
Clan CR is composed of the C108 family, the Prp peptidase identified in Staphylococcus aureus. Possessing a His, Cys catalytic dyad, this protease is important for processing the N-terminal extension of ribosomal protein L27 and is essential in S. aureus (77, 78). The Streptococcus phage CP-1 is also included in this family. CP-1 is an endopeptidase that cleaves the first 48 amino acids of the major head protein (79).
C. DIFFICILE CYSTEINE PROTEASES
Clostridioides difficile is the causative agent of C. difficile infection (CDI) (80, 81). The severity of CDI can range from mild diarrhea to more severe pseudomembranous colitis and toxic megacolon (82). In 2019, the Centers for Disease Control and Prevention (CDC) estimated that there were 223,900 hospitalized cases and approximately 13,000 deaths caused by CDI (83). Moreover, the annual treatment-associated costs to the United States healthcare system are estimated to ~$5 billion (84, 85). Though CDI is a common occurrence in the healthcare setting, the incidence of community-acquired CDI has increased, particularly in immunocompromised patients (81, 84, 86, 87).
The infectious form of C. difficile is the spore because its vegetative cells are strictly anaerobic (88, 89). Spores are metabolically dormant, survive in the aerobic environment outside of the host, and permit host-to-host transmission (89–93). Upon ingestion, spores persist in the gut environment and germinate in response to host-derived bile acid germinants and amino acid co-germinants (88, 93–95).
As of the 2025 MEROPS database, many C. difficile strains encode approximately 192 known or putative proteases. One of the most well-studied is Cwp84, a member of clan CA. Cwp84 is a cell wall-localized cysteine protease and is responsible for processing the S-layer protein SlpA (96). The two C. difficile toxins TcdA and TcdB have a cysteine protease domain adjacent to the glucosyltransferase domain (GTD) and are members of clan CD (97). The loss of catalytic activity results in less cytotoxicity to the cell (50, 97–99).
The vegetative cells secrete the TcdA and TcdB toxins that are necessary and sufficient for disease symptoms (100–102). TcdA is a 308 kDa enterotoxin, and TcdB is a 270 kDa cytotoxin (100, 103, 104). Both toxins are classified as ABCD toxins, where there are four functional domains: A stands for biological activity; B, binding; C, cutting; and D, delivery (105). TcdA and TcdB are composed of a glucosyltransferase domain (GTD), a Clan CD cysteine protease domain, also known as the autoprocessing domain (APD), a delivery domain, and a receptor binding domain (50, 104, 106). The His and Cys catalytic residues of the TcdA/TcdB APD bind to Zn2+ and are essential for its autoprocessing activity (51, 107). These toxins bind to the target cell membrane through their C-terminal receptor binding domains, resulting in internalization of the toxin. After binding, the toxins are endocytosed to the endosomal compartment. A conformational change occurs that allows the toxin to interact with the endosomal membrane and generation of the pore. Autoprocessing occurs in the presence of host-derived InsP6 that binds to the adjacent domain of the monoglucosyltransferase that releases the N-terminal glucosyltransferase domain (GTD) into the host cytosol and targets Rho-family GTPases domain (50, 99, 105, 108, 109). The domain adjacent to the GTD of TcdA/TcdB shares sequence homology to MARTX toxins and is activated in the presence of InsP6 in the host cytosol (40, 110). GTPases are molecular switches that regulate cellular processes. Rho GTPases are involved in signaling processes, such as actin cytoskeleton regulation, phagocytosis, cell polarity, and more (111). The released GTD is then free to glucosylate Rho proteins, thus inactivating them. The result of this inactivation is cytopathic effects characterized by loss of actin stress fibers, disruption of intercellular tight junctions, increased cell barrier permeability, and eventually, cell death (100, 112–117). Interestingly, work done by Chumbler et al. (2012) revealed TcdB autoprocessing resulting in release of the GTD is not required for necrosis in epithelial cells (118). In addition to epithelial cell death, TcdA and TcdB cause inflammation in the epithelial cells induced by chemokines that promote neutrophil recruitment. The recruitment of neutrophils has been demonstrated to cause increased fluid secretion and enhanced mucosal inflammation, leading to diarrhea (119–122).
YabG: AN UNCLASSIFIED CYSTEINE PROTEASE
YabG is a sporulation-specific cysteine protease found in all endospore-forming bacteria (123–127). In C. difficile, YabG processes two proteins that are critical for spore germination, CspBA and preproSleC, but is likely to have other unidentified targets in the C. difficile spore (127–130). Though much work has been done to study this protease, it remains unclassified in the MEROPS database (123–131).
In B. subtilis, YabG is responsible for proper processing of spore coat proteins. In C. difficile, YabG has been shown to both process proteins important for spore germination and regulate the incorporation of coat and exosporium proteins (123, 124, 126, 128–130). B. subtilis YabG has Cys218 and His172 catalytic residues required for autoprocessing (126).
Currently, YabG remains unclassified in the MEROPS database as U57. YabG possesses His-Cys catalytic residues with arginine specificity in the P1 position that are conserved across many endospore-forming bacteria (126, 129). In C. difficile, YabG has conserved His161, Asp162, and Cys207 residues that are required for autoprocessing (129). B. subtilis YabG possesses an Asp at position 173 that corresponds to C. difficile Asp162; however, the involvement in autoprocessing has not been tested (126). There are currently four cysteine protease clans that possess a His-Cys catalytic dyad: clans CD, CL, CP, and CQ (1). It was suggested by both Marini et al. and Yamazawa et al. that YabG be classified as a clan CD protease. Like other proteases within the clan CD, YabG does not share sequence homology with other members of this clan; however, the His-Cys catalytic dyad and Arg specificity make it a strong candidate for the CD clan (Fig. 1) (126, 128, 129). Members of the MEROPS database are not classified based on sequence similarity; rather, they are classified based on similar structure features and functions. The clan CD contains proteases with a highly conserved His-Cys or His-Cys-Asp/Asn catalytic dyad or triad. Other clans of cysteine proteases contain a His-Cys catalytic dyad, but they do not share the same function as those in clan CD. Members of the clan CD possess a Gly residue preceding the catalytic His residue. However, in YabG, the Gly residue following the His residue and the active Cys occurs in an Ala-Cys motif (129). This could result in a new family present within clan CD that contains enzymes with the same features as YabG.
Fig 1.
Sequence comparison of YabG to clan CD proteases. Alignment of catalytic residue amino acid sequences. B. subtilis 168 YabG (B. sub) (P37548); C. difficile R20291 (C. diff) (A0A9R0BNR8https://www.ncbi.nlm.nih.gov/protein/P09870/); Clostripain from H. histolytica (P09870); Legumain from mouse (O89017); Caspase-1 from human (P29466); Gingipain from P. gingivalis (P28784); Separase from S. cerevisiae (Q03018); RTX from V. cholerae (Q9KS12); and PrtH from T. forsythia (O24742). The conserved His-Gly and Cys residues are highlighted in black. Residues highlighted in gray are 80% conserved. Red * indicates the YabG catalytic His, and red ** indicates the YabG catalytic Cys residues relative to other cysteine proteases.
CONCLUDING REMARKS
Here, we present the current classification of cysteine protease clans and some of their families. With advances in protein structure and modeling, such as AlphaFold, better assignment of proteases to clans will be established. Much of the work done to understand the role and function of the sporulation-specific protease, YabG, has been performed in B. subtilis, with more recent efforts focused on characterizing the function of YabG in C. difficile (123–126, 128–132). Yamazawa et al. have shown that B. subtilis YabG possesses a catalytic dyad of His172 and Cys218 that allows for auto-processing and processing of substrates. Recently, it was shown that C. difficile YabG functions similarly, with a catalytic dyad of His161 and Cys207, and processes germination proteins CspBA and preproSleC along with regulating transcription of cotA and cdeM (128, 129). Based upon these data that identified the conserved catalytic dyad, we propose that YabG is an ideal candidate for clan CD in the MEROPS (Fig. 1).
ACKNOWLEDGMENTS
Thanks to the members of the Sorg laboratory for their helpful comments during the preparation of this manuscript. This project was supported by awards 5R01AI116895 and 5R01AI172043 to J.A.S. from the National Institute of Allergy and Infectious Diseases. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIAID. The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.
Contributor Information
Joseph A. Sorg, Email: jsorg@bio.tamu.edu.
Tina M. Henkin, The Ohio State University, Columbus, Ohio, USA
REFERENCES
- 1. Rawlings ND, Waller M, Barrett AJ, Bateman A. 2014. MEROPS: the database of proteolytic enzymes, their substrates and inhibitors. Nucl Acids Res 42:D503–D509. doi: 10.1093/nar/gkt953 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Subramanian S, Sijwali PS, Rosenthal PJ. 2007. Falcipain cysteine proteases require bipartite motifs for trafficking to the Plasmodium falciparum food vacuole. J Biol Chem 282:24961–24969. doi: 10.1074/jbc.M703316200 [DOI] [PubMed] [Google Scholar]
- 3. Tao K, Stearns NA, Dong JM, Wu QI, Sahagian GG. 1994. The proregion of cathepsin L is required for proper folding, stability, and ER exit. Arch Biochem Biophys 311:19–27. doi: 10.1006/abbi.1994.1203 [DOI] [PubMed] [Google Scholar]
- 4. Verma S, Dixit R, Pandey KC. 2016. Cysteine proteases: modes of activation and future prospects as pharmacological targets. Front Pharmacol 7:107. doi: 10.3389/fphar.2016.00107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Mason RW, Massey SD. 1992. Surface activation of pro-cathepsin L. Biochem Biophys Res Commun 189:1659–1666. doi: 10.1016/0006-291x(92)90268-p [DOI] [PubMed] [Google Scholar]
- 6. Yang N, Matthew MA, Yao C. 2023. Roles of cysteine proteases in biology and pathogenesis of parasites. Microorganisms 11:1397. doi: 10.3390/microorganisms11061397 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Rao MB, Tanksale AM, Ghatge MS, Deshpande VV. 1998. Molecular and biotechnological aspects of microbial proteases. Microbiol Mol Biol Rev 62:597–635. doi: 10.1128/MMBR.62.3.597-635.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Drenth J, Jansonius JN, Koekoek R, Swen HM, Wolthers BG. 1968. Structure of papain. Nature 218:929–932. doi: 10.1038/218929a0 [DOI] [PubMed] [Google Scholar]
- 9. Harris JL, Backes BJ, Leonetti F, Mahrus S, Ellman JA, Craik CS. 2000. Rapid and general profiling of protease specificity by using combinatorial fluorogenic substrate libraries. Proc Natl Acad Sci USA 97:7754–7759. doi: 10.1073/pnas.140132697 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. de la Riva L, Willing SE, Tate EW, Fairweather NF. 2011. Roles of cysteine proteases Cwp84 and Cwp13 in biogenesis of the cell wall of Clostridium difficile. J Bacteriol 193:3276–3285. doi: 10.1128/JB.00248-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Kirby JM, Ahern H, Roberts AK, Kumar V, Freeman Z, Acharya KR, Shone CC. 2009. Cwp84, a surface-associated cysteine protease, plays a role in the maturation of the surface layer of Clostridium difficile. J Biol Chem 284:34666–34673. doi: 10.1074/jbc.M109.051177 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Janoir C, Péchiné S, Grosdidier C, Collignon A. 2007. Cwp84, a surface-associated protein of Clostridium difficile, is a cysteine protease with degrading activity on extracellular matrix proteins. J Bacteriol 189:7174–7180. doi: 10.1128/JB.00578-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Soh WT, Demir F, Dall E, Perrar A, Dahms SO, Kuppusamy M, Brandstetter H, Huesgen PF. 2020. ExteNDing proteome coverage with legumain as a highly specific digestion protease. Anal Chem 92:2961–2971. doi: 10.1021/acs.analchem.9b03604 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Exconde PM, Bourne CM, Kulkarni M, Discher BM, Taabazuing CY. 2024. Inflammatory caspase substrate specificities. mBio 15:e0297523. doi: 10.1128/mbio.02975-23 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Ullmann D, Jakubke H-D. 1994. Kinetic characterization of affinity chromatography purified clostripain. Biol Chem Hoppe Seyler 375:89–92. doi: 10.1515/bchm3.1994.375.2.89 [DOI] [PubMed] [Google Scholar]
- 16. Manabe S, Nariya H, Miyata S, Tanaka H, Minami J, Suzuki M, Taniguchi Y, Okabe A. 2010. Purification and characterization of a clostripain-like protease from a recombinant Clostridium perfringens culture. Microbiology 156:561–569. doi: 10.1099/mic.0.031609-0 [DOI] [PubMed] [Google Scholar]
- 17. Tanaka H, Nariya H, Suzuki M, Houchi H, Tamai E, Miyata S, Okabe A. 2011. High-level production and purification of clostripain expressed in a virulence-attenuated strain of Clostridium perfringens. Protein Expr Purif 76:83–89. doi: 10.1016/j.pep.2010.10.002 [DOI] [PubMed] [Google Scholar]
- 18. Potempa J, Pavloff N, Travis J. 1995. Porphyromonas gingivalis: a proteinase/gene accounting audit. Trends Microbiol 3:430–434. doi: 10.1016/s0966-842x(00)88996-9 [DOI] [PubMed] [Google Scholar]
- 19. Buonomo SBC, Clyne RK, Fuchs J, Loidl J, Uhlmann F, Nasmyth K. 2000. Disjunction of homologous chromosomes in meiosis I depends on proteolytic cleavage of the meiotic cohesin rec8 by separin. Cell 103:387–398. doi: 10.1016/S0092-8674(00)00131-8 [DOI] [PubMed] [Google Scholar]
- 20. Uhlmann F, Lottspeich F, Nasmyth K. 1999. Sister-chromatid separation at anaphase onset is promoted by cleavage of the cohesin subunit Scc1. Nature 400:37–42. doi: 10.1038/21831 [DOI] [PubMed] [Google Scholar]
- 21. McLuskey K, Mottram JC. 2015. Comparative structural analysis of the caspase family with other clan CD cysteine peptidases. Biochem J 466:219–232. doi: 10.1042/BJ20141324 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Choi CJ, Berges JA. 2013. New types of metacaspases in phytoplankton reveal diverse origins of cell death proteases. Cell Death Dis 4:e490. doi: 10.1038/cddis.2013.21 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Tsiatsiani L, Van Breusegem F, Gallois P, Zavialov A, Lam E, Bozhkov PV. 2011. Metacaspases. Cell Death Differ 18:1279–1288. doi: 10.1038/cdd.2011.66 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Pop C, Salvesen GS. 2009. Human caspases: activation, specificity, and regulation. J Biol Chem 284:21777–21781. doi: 10.1074/jbc.R800084200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Shi Y. 2004. Caspase activation, inhibition, and reactivation: a mechanistic view. Protein Sci 13:1979–1987. doi: 10.1110/ps.04789804 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Seaman JE, Julien O, Lee PS, Rettenmaier TJ, Thomsen ND, Wells JA. 2016. Cacidases: caspases can cleave after aspartate, glutamate and phosphoserine residues. Cell Death Differ 23:1717–1726. doi: 10.1038/cdd.2016.62 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Lawson PA, Rainey FA. 2016. Proposal to restrict the genus Clostridium prazmowski to Clostridium butyricum and related species. Int J Syst Evol Microbiol 66:1009–1016. doi: 10.1099/ijsem.0.000824 [DOI] [PubMed] [Google Scholar]
- 28. Guzik K, Bzowska M, Smagur J, Krupa O, Sieprawska M, Travis J, Potempa J. 2007. A new insight into phagocytosis of apoptotic cells: proteolytic enzymes divert the recognition and clearance of polymorphonuclear leukocytes by macrophages. Cell Death Differ 14:171–182. doi: 10.1038/sj.cdd.4401927 [DOI] [PubMed] [Google Scholar]
- 29. Labrou NE, Rigden DJ. 2004. The structure-function relationship in the clostripain family of peptidases. Eur J Biochem 271:983–992. doi: 10.1111/j.1432-1033.2004.04000.x [DOI] [PubMed] [Google Scholar]
- 30. Tanaka H, Tamai E, Miyata S, Taniguchi Y, Nariya H, Hatano N, Houchi H, Okabe A. 2008. Construction and characterization of a clostripain-like protease-deficient mutant of Clostridium perfringens as a strain for clostridial gene expression. Appl Microbiol Biotechnol 77:1063–1071. doi: 10.1007/s00253-007-1245-9 [DOI] [PubMed] [Google Scholar]
- 31. Chen JM, Rawlings ND, Stevens RA, Barrett AJ. 1998. Identification of the active site of legumain links it to caspases, clostripain and gingipains in a new clan of cysteine endopeptidases. FEBS Lett 441:361–365. doi: 10.1016/s0014-5793(98)01574-9 [DOI] [PubMed] [Google Scholar]
- 32. Ståhle M, Foss A, Gustafsson B, Lempinen M, Lundgren T, Rafael E, Tufveson G, Korsgren O, Friberg A. 2015. Clostripain, the missing link in the enzyme blend for efficient human islet isolation. Transplant Direct 1:e19. doi: 10.1097/TXD.0000000000000528 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Montero DA, Vidal RM, Velasco J, George S, Lucero Y, Gómez LA, Carreño LJ, García-Betancourt R, O’Ryan M. 2023. Vibrio cholerae, classification, pathogenesis, immune response, and trends in vaccine development. Front Med 10:1155751. doi: 10.3389/fmed.2023.1155751 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Shen A, Lupardus PJ, Albrow VE, Guzzetta A, Powers JC, Garcia KC, Bogyo M. 2009. Mechanistic and structural insights into the proteolytic activation of Vibrio cholerae MARTX toxin. Nat Chem Biol 5:469–478. doi: 10.1038/nchembio.178 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Linhartová I, Bumba L, Mašín J, Basler M, Osička R, Kamanová J, Procházková K, Adkins I, Hejnová-Holubová J, Sadílková L, Morová J, Sebo P. 2010. RTX proteins: a highly diverse family secreted by a common mechanism. FEMS Microbiol Rev 34:1076–1112. doi: 10.1111/j.1574-6976.2010.00231.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Satchell KJF. 2007. MARTX, multifunctional autoprocessing repeats-in-toxin toxins. Infect Immun 75:5079–5084. doi: 10.1128/IAI.00525-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Lin W, Fullner KJ, Clayton R, Sexton JA, Rogers MB, Calia KE, Calderwood SB, Fraser C, Mekalanos JJ. 1999. Identification of a Vibrio cholerae RTX toxin gene cluster that is tightly linked to the cholera toxin prophage. Proc Natl Acad Sci USA 96:1071–1076. doi: 10.1073/pnas.96.3.1071 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Boardman BK, Satchell KJF. 2004. Vibrio cholerae strains with mutations in an atypical type I secretion system accumulate RTX toxin intracellularly. J Bacteriol 186:8137–8143. doi: 10.1128/JB.186.23.8137-8143.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Sheahan KL, Cordero CL, Satchell KJF. 2007. Autoprocessing of the Vibrio cholerae RTX toxin by the cysteine protease domain. EMBO J 26:2552–2561. doi: 10.1038/sj.emboj.7601700 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Egerer M, Satchell KJF. 2010. Inositol hexakisphosphate-induced autoprocessing of large bacterial protein toxins. PLoS Pathog 6:e1000942. doi: 10.1371/journal.ppat.1000942 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Prochazkova K, Shuvalova LA, Minasov G, Voburka Z, Anderson WF, Satchell KJF. 2009. Structural and molecular mechanism for autoprocessing of MARTX toxin of Vibrio cholerae at multiple sites. J Biol Chem 284:26557–26568. doi: 10.1074/jbc.M109.025510 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Lupardus PJ, Shen A, Bogyo M, Garcia KC. 2008. Small molecule-induced allosteric activation of the Vibrio cholerae RTX cysteine protease domain. Science 322:265–268. doi: 10.1126/science.1162403 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Prochazkova K, Satchell KJF. 2008. Structure-function analysis of inositol hexakisphosphate-induced autoprocessing of the Vibrio cholerae multifunctional autoprocessing RTX toxin. J Biol Chem 283:23656–23664. doi: 10.1074/jbc.M803334200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Fullner KJ, Mekalanos JJ. 2000. In vivo covalent cross-linking of cellular actin by the Vibrio cholerae RTX toxin. EMBO J 19:5315–5323. doi: 10.1093/emboj/19.20.5315 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Sheahan KL, Cordero CL, Satchell KJF. 2004. Identification of a domain within the multifunctional Vibrio cholerae RTX toxin that covalently cross-links actin. Proc Natl Acad Sci USA 101:9798–9803. doi: 10.1073/pnas.0401104101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Kudryashov DS, Cordero CL, Reisler E, Satchell KJF. 2008. Characterization of the enzymatic activity of the actin cross-linking domain from the Vibrio cholerae MARTX Vc toxin. J Biol Chem 283:445–452. doi: 10.1074/jbc.M703910200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Zhou Y, Huang C, Yin L, Wan M, Wang X, Li L, Liu Y, Wang Z, Fu P, Zhang N, Chen S, Liu X, Shao F, Zhu Y. 2017. N(ε)-fatty acylation of Rho GTPases by a MARTX toxin effector. Science 358:528–531. doi: 10.1126/science.aam8659 [DOI] [PubMed] [Google Scholar]
- 48. Rupnik M, Wilcox MH, Gerding DN. 2009. Clostridium difficile infection: new developments in epidemiology and pathogenesis. Nat Rev Microbiol 7:526–536. doi: 10.1038/nrmicro2164 [DOI] [PubMed] [Google Scholar]
- 49. Chandrasekaran R, Lacy DB. 2017. The role of toxins in Clostridium difficile infection. FEMS Microbiol Rev 41:723–750. doi: 10.1093/femsre/fux048 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Egerer M, Giesemann T, Jank T, Satchell KJF, Aktories K. 2007. Auto-catalytic cleavage of Clostridium difficile toxins A and B depends on cysteine protease activity. J Biol Chem 282:25314–25321. doi: 10.1074/jbc.M703062200 [DOI] [PubMed] [Google Scholar]
- 51. Chumbler NM, Rutherford SA, Zhang Z, Farrow MA, Lisher JP, Farquhar E, Giedroc DP, Spiller BW, Melnyk RA, Lacy DB. 2016. Crystal structure of Clostridium difficile toxin A. Nat Microbiol 1:15002. doi: 10.1038/nmicrobiol.2015.2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Greber UF, Webster P, Weber J, Helenius A. 1996. The role of the adenovirus protease on virus entry into cells. EMBO J 15:1766–1777. doi: 10.1002/j.1460-2075.1996.tb00525.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Cotten M, Weber JM. 1995. The adenovirus protease is required for virus entry into host cells. Virology 213:494–502. doi: 10.1006/viro.1995.0022 [DOI] [PubMed] [Google Scholar]
- 54. Greber UF. 1998. Virus assembly and disassembly: the adenovirus cysteine protease as a trigger factor. Rev Med Virol 8:213–222. doi: [DOI] [PubMed] [Google Scholar]
- 55. Mentlein R. 2004. Cell-surface peptidases. Int Rev Cytol 235:165–213. doi: 10.1016/S0074-7696(04)35004-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Doolittle RF, Armentrout RW. 1968. Pyrrolidonyl peptidase. an enzyme for selective removal of pyrrolidonecarboxylic acid residues from polypeptides. Biochemistry 7:516–521. doi: 10.1021/bi00842a005 [DOI] [PubMed] [Google Scholar]
- 57. Odagaki Y, Hayashi A, Okada K, Hirotsu K, Kabashima T, Ito K, Yoshimoto T, Tsuru D, Sato M, Clardy J. 1999. The crystal structure of pyroglutamyl peptidase I from Bacillus amyloliquefaciens reveals a new structure for a cysteine protease. Structure 7:399–411. doi: 10.1016/s0969-2126(99)80053-7 [DOI] [PubMed] [Google Scholar]
- 58. Hendrickx APA, Budzik JM, Oh S-Y, Schneewind O. 2011. Architects at the bacterial surface - sortases and the assembly of pili with isopeptide bonds. Nat Rev Microbiol 9:166–176. doi: 10.1038/nrmicro2520 [DOI] [PubMed] [Google Scholar]
- 59. Mazmanian SK, Liu G, Ton-That H, Schneewind O. 1999. Staphylococcus aureus sortase, an enzyme that anchors surface proteins to the cell wall. Science 285:760–763. doi: 10.1126/science.285.5428.760 [DOI] [PubMed] [Google Scholar]
- 60. Das S, Pawale VS, Dadireddy V, Singh AK, Ramakumar S, Roy RP. 2017. Structure and specificity of a new class of Ca2+-independent housekeeping sortase from Streptomyces avermitilis provide insights into its non-canonical substrate preference. J Biol Chem 292:7244–7257. doi: 10.1074/jbc.M117.782037 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Magnet S, Arbeloa A, Mainardi JL, Hugonnet JE, Fourgeaud M, Dubost L, Marie A, Delfosse V, Mayer C, Rice LB, Arthur M. 2007. Specificity of L,D-transpeptidases from gram-positive bacteria producing different peptidoglycan chemotypes. J Biol Chem 282:13151–13159. doi: 10.1074/jbc.M610911200 [DOI] [PubMed] [Google Scholar]
- 62. Mainardi JL, Fourgeaud M, Hugonnet JE, Dubost L, Brouard JP, Ouazzani J, Rice LB, Gutmann L, Arthur M. 2005. A novel peptidoglycan cross-linking enzyme for a beta-lactam-resistant transpeptidation pathway. J Biol Chem 280:38146–38152. doi: 10.1074/jbc.M507384200 [DOI] [PubMed] [Google Scholar]
- 63. Lorenz IC, Marcotrigiano J, Dentzer TG, Rice CM. 2006. Structure of the catalytic domain of the hepatitis C virus NS2-3 protease. Nature 442:831–835. doi: 10.1038/nature04975 [DOI] [PubMed] [Google Scholar]
- 64. Welbourn S, Pause A. 2007. The hepatitis C virus NS2/3 protease. Curr Issues Mol Biol 9:63–69. doi: 10.21775/cimb.009.063 [DOI] [PubMed] [Google Scholar]
- 65. Strauss EG, De Groot RJ, Levinson R, Strauss JH. 1992. Identification of the active site residues in the nsP2 proteinase of Sindbis virus. Virology 191:932–940. doi: 10.1016/0042-6822(92)90268-t [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Russo AT, White MA, Watowich SJ. 2006. The crystal structure of the Venezuelan equine encephalitis alphavirus nsP2 protease. Structure 14:1449–1458. doi: 10.1016/j.str.2006.07.010 [DOI] [PubMed] [Google Scholar]
- 67. Bourgogne T, Vacheron MJ, Guinand M, Michel G. 1992. Purification and partial characterization of the gamma-D-glutamyl-L-di-amino acid endopeptidase II from Bacillus sphaericus. Int J Biochem 24:471–476. doi: 10.1016/0020-711x(92)90041-x [DOI] [PubMed] [Google Scholar]
- 68. Fukushima T, Afkham A, Kurosawa S-I, Tanabe T, Yamamoto H, Sekiguchi J. 2006. A new D,L-endopeptidase gene product, YojL (renamed CwlS), plays a role in cell separation with LytE and LytF in Bacillus subtilis. J Bacteriol 188:5541–5550. doi: 10.1128/JB.00188-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Ishikawa S, Hara Y, Ohnishi R, Sekiguchi J. 1998. Regulation of a new cell wall hydrolase gene, cwlF, which affects cell separation in Bacillus subtilis. J Bacteriol 180:2549–2555. doi: 10.1128/JB.180.9.2549-2555.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Ohnishi R, Ishikawa S, Sekiguchi J. 1999. Peptidoglycan hydrolase LytF plays a role in cell separation with CwlF during vegetative growth of Bacillus subtilis. J Bacteriol 181:3178–3184. doi: 10.1128/JB.181.10.3178-3184.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Aramini JM, Rossi P, Huang YJ, Zhao L, Jiang M, Maglaqui M, Xiao R, Locke J, Nair R, Rost B, Acton TB, Inouye M, Montelione GT. 2008. Solution NMR structure of the NlpC/P60 domain of lipoprotein Spr from Escherichia coli: structural evidence for a novel cysteine peptidase catalytic triad. Biochemistry 47:9715–9717. doi: 10.1021/bi8010779 [DOI] [PubMed] [Google Scholar]
- 72. Nayak A, Müller S. 2014. SUMO-specific proteases/isopeptidases: SENPs and beyond. Genome Biol 15:422. doi: 10.1186/s13059-014-0422-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Rümenapf T, Stark R, Heimann M, Thiel HJ. 1998. N-terminal protease of pestiviruses: identification of putative catalytic residues by site-directed mutagenesis. J Virol 72:2544–2547. doi: 10.1128/JVI.72.3.2544-2547.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Gottipati K, Acholi S, Ruggli N, Choi KH. 2014. Autocatalytic activity and substrate specificity of the pestivirus N-terminal protease Npro. Virology 452–453:303–309. doi: 10.1016/j.virol.2014.01.026 [DOI] [PubMed] [Google Scholar]
- 75. Bauhofer O, Summerfield A, Sakoda Y, Tratschin J-D, Hofmann MA, Ruggli N. 2007. Classical swine fever virus Npro interacts with interferon regulatory factor 3 and induces its proteasomal degradation. J Virol 81:3087–3096. doi: 10.1128/JVI.02032-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Chen Z, Rijnbrand R, Jangra RK, Devaraj SG, Qu L, Ma Y, Lemon SM, Li K. 2007. Ubiquitination and proteasomal degradation of interferon regulatory factor-3 induced by Npro from a cytopathic bovine viral diarrhea virus. Virology 366:277–292. doi: 10.1016/j.virol.2007.04.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Wall EA, Caufield JH, Lyons CE, Manning KA, Dokland T, Christie GE. 2015. Specific N-terminal cleavage of ribosomal protein L27 in Staphylococcus aureus and related bacteria. Mol Microbiol 95:258–269. doi: 10.1111/mmi.12862 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Hotinger JA, Pendergrass HA, Peterson D, Wright HT, May AE. 2022. Phage-related ribosomal protease (Prp) of Staphylococcus aureus: in vitro michaelis-menten kinetics, screening for inhibitors, and crystal structure of a covalent inhibition product complex. Biochemistry 61:1323–1336. doi: 10.1021/acs.biochem.2c00010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Martín AC, López R, García P. 1998. Pneumococcal bacteriophage Cp-1 encodes its own protease essential for phage maturation. J Virol 72:3491–3494. doi: 10.1128/JVI.72.4.3491-3494.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Bartlett JG, Moon N, Chang TW, Taylor N, Onderdonk AB. 1978. Role of Clostridium difficile in antibiotic-associated pseudomembranous colitis. Gastroenterology 75:778–782. doi: 10.1016/0016-5085(78)90457-2 [DOI] [PubMed] [Google Scholar]
- 81. Di Bella S, Sanson G, Monticelli J, Zerbato V, Principe L, Giuffre M, Pipitone G, Luzzati R. 2024. Clostridioides difficile infection: history, epidemiology, risk factors, prevention, clinical manifestations, treatment, and future options. Clin Microbiol Rev 37:e0013523. doi: 10.1128/cmr.00135-23:e0013523 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Sayedy L, Kothari D, Richards RJ. 2010. Toxic megacolon associated Clostridium difficile colitis. World J Gastrointest Endosc 2:293–297. doi: 10.4253/wjge.v2.i8.293 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Anonymous . 2019. CDC. Antibiotic resistance threats in the United States, 2019
- 84. Feuerstadt P, Theriault N, Tillotson G. 2023. The burden of CDI in the United States: a multifactorial challenge. BMC Infect Dis 23:132. doi: 10.1186/s12879-023-08096-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Dubberke ER, Olsen MA. 2012. Burden of Clostridium difficile on the healthcare system. Clin Infect Dis 55 Suppl 2:S88–92. doi: 10.1093/cid/cis335 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Theriot CM, Koenigsknecht MJ, Carlson PE, Jr, Hatton GE, Nelson AM, Li B, Huffnagle GB, Li JZ, Young VB. 2014. Antibiotic-induced shifts in the mouse gut microbiome and metabolome increase susceptibility to Clostridium difficile infection. Nat Commun 5:3114. doi: 10.1038/ncomms4114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Buffie CG, Bucci V, Stein RR, McKenney PT, Ling L, Gobourne A, No D, Liu H, Kinnebrew M, Viale A, Littmann E, van den Brink MRM, Jenq RR, Taur Y, Sander C, Cross JR, Toussaint NC, Xavier JB, Pamer EG. 2015. Precision microbiome reconstitution restores bile acid mediated resistance to Clostridium difficile. Nature 517:205–208. doi: 10.1038/nature13828 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Smits WK, Lyras D, Lacy DB, Wilcox MH, Kuijper EJ. 2016. Clostridium difficile infection. Nat Rev Dis Primers 2:16020. doi: 10.1038/nrdp.2016.20 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Lawley TD, Clare S, Walker AW, Goulding D, Stabler RA, Croucher N, Mastroeni P, Scott P, Raisen C, Mottram L, Fairweather NF, Wren BW, Parkhill J, Dougan G. 2009. Antibiotic treatment of Clostridium difficile carrier mice triggers a supershedder state, spore-mediated transmission, and severe disease in immunocompromised hosts. Infect Immun 77:3661–3669. doi: 10.1128/IAI.00558-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Deakin LJ, Clare S, Fagan RP, Dawson LF, Pickard DJ, West MR, Wren BW, Fairweather NF, Dougan G, Lawley TD. 2012. The Clostridium difficile spo0A gene is a persistence and transmission factor. Infect Immun 80:2704–2711. doi: 10.1128/IAI.00147-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Paredes-Sabja D, Shen A, Sorg JA. 2014. Clostridium difficile spore biology: sporulation, germination, and spore structural proteins. Trends Microbiol 22:406–416. doi: 10.1016/j.tim.2014.04.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Buddle JE, Fagan RP. 2023. Pathogenicity and virulence of Clostridioides difficile Virulence 14:2150452. doi: 10.1080/21505594.2022.2150452 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Seekatz AM, Young VB. 2014. Clostridium difficile and the microbiota. J Clin Invest 124:4182–4189. doi: 10.1172/JCI72336 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Sorg JA, Sonenshein AL. 2008. Bile salts and glycine as cogerminants for Clostridium difficile spores. J Bacteriol 190:2505–2512. doi: 10.1128/JB.01765-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Sorg JA. 2009. Chenodeoxycholate is an inhibitor of Clostridium difficile spore germination. J Bacteriol 191:1115–1117. doi: 10.1128/JB.01260-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Pantaléon V, Soavelomandroso AP, Bouttier S, Briandet R, Roxas B, Chu M, Collignon A, Janoir C, Vedantam G, Candela T. 2015. The Clostridium difficile protease Cwp84 modulates both biofilm formation and cell-surface properties. PLoS One 10:e0124971. doi: 10.1371/journal.pone.0124971 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Rupnik M, Pabst S, Rupnik M, von Eichel-Streiber C, Urlaub H, Söling HD. 2005. Characterization of the cleavage site and function of resulting cleavage fragments after limited proteolysis of Clostridium difficile toxin B (TcdB) by host cells. Microbiology 151:199–208. doi: 10.1099/mic.0.27474-0 [DOI] [PubMed] [Google Scholar]
- 98. Li S, Shi L, Yang Z, Feng H. 2013. Cytotoxicity of Clostridium difficile toxin B does not require cysteine protease-mediated autocleavage and release of the glucosyltransferase domain into the host cell cytosol. Pathog Dis 67:11–18. doi: 10.1111/2049-632X.12016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Reineke J, Tenzer S, Rupnik M, Koschinski A, Hasselmayer O, Schrattenholz A, Schild H, von Eichel-Streiber C. 2007. Autocatalytic cleavage of Clostridium difficile toxin B. Nature 446:415–419. doi: 10.1038/nature05622 [DOI] [PubMed] [Google Scholar]
- 100. Voth DE, Ballard JD. 2005. Clostridium difficile toxins: mechanism of action and role in disease. Clin Microbiol Rev 18:247–263. doi: 10.1128/CMR.18.2.247-263.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Shen A. 2012. Clostridium difficile toxins: mediators of inflammation. J Innate Immun 4:149–158. doi: 10.1159/000332946 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Janoir C. 2016. Virulence factors of Clostridium difficile and their role during infection. Anaerobe 37:13–24. doi: 10.1016/j.anaerobe.2015.10.009 [DOI] [PubMed] [Google Scholar]
- 103. Lyras D, O’Connor JR, Howarth PM, Sambol SP, Carter GP, Phumoonna T, Poon R, Adams V, Vedantam G, Johnson S, Gerding DN, Rood JI. 2009. Toxin B is essential for virulence of Clostridium difficile. Nature 458:1176–1179. doi: 10.1038/nature07822 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Pruitt RN, Chambers MG, Ng KK-S, Ohi MD, Lacy DB. 2010. Structural organization of the functional domains of Clostridium difficile toxins A and B. Proc Natl Acad Sci USA 107:13467–13472. doi: 10.1073/pnas.1002199107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Jank T, Aktories K. 2008. Structure and mode of action of clostridial glucosylating toxins: the ABCD model. Trends Microbiol 16:222–229. doi: 10.1016/j.tim.2008.01.011 [DOI] [PubMed] [Google Scholar]
- 106. Pruitt RN, Chagot B, Cover M, Chazin WJ, Spiller B, Lacy DB. 2009. Structure-function analysis of inositol hexakisphosphate-induced autoprocessing in Clostridium difficile toxin A. J Biol Chem 284:21934–21940. doi: 10.1074/jbc.M109.018929 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Cowardin CA, Jackman BM, Noor Z, Burgess SL, Feig AL, Petri WA, Jr. 2016. Glucosylation drives the innate inflammatory response to Clostridium difficile toxin A. Infect Immun 84:2317–2323. doi: 10.1128/IAI.00327-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Just I, Selzer J, Wilm M, von Eichel-Streiber C, Mann M, Aktories K. 1995. Glucosylation of Rho proteins by Clostridium difficile toxin B. Nature 375:500–503. doi: 10.1038/375500a0 [DOI] [PubMed] [Google Scholar]
- 109. Just I, Wilm M, Selzer J, Rex G, von Eichel-Streiber C, Mann M, Aktories K. 1995. The enterotoxin from Clostridium difficile (ToxA) monoglucosylates the Rho proteins. J Biol Chem 270:13932–13936. doi: 10.1074/jbc.270.23.13932 [DOI] [PubMed] [Google Scholar]
- 110. Shen A. 2010. Allosteric regulation of protease activity by small molecules. Mol Biosyst 6:1431–1443. doi: 10.1039/c003913f [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Etienne-Manneville S, Hall A. 2002. Rho GTPases in cell biology. Nature 420:629–635. doi: 10.1038/nature01148 [DOI] [PubMed] [Google Scholar]
- 112. Donta ST, Sullivan N, Wilkins TD. 1982. Differential effects of Clostridium difficile toxins on tissue-cultured cells. J Clin Microbiol 15:1157–1158. doi: 10.1128/jcm.15.6.1157-1158.1982 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Chaves-Olarte E, Weidmann M, Eichel-Streiber C, Thelestam M. 1997. Toxins A and B from Clostridium difficile differ with respect to enzymatic potencies, cellular substrate specificities, and surface binding to cultured cells. J Clin Invest 100:1734–1741. doi: 10.1172/JCI119698 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Chumbler NM, Farrow MA, Lapierre LA, Franklin JL, Lacy DB. 2016. Clostridium difficile toxins TcdA and TcdB cause colonic tissue damage by distinct mechanisms. Infect Immun 84:2871–2877. doi: 10.1128/IAI.00583-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Just I, Gerhard R. 2004. Large clostridial cytotoxins. Rev Physiol Biochem Pharmacol 152:23–47. doi: 10.1007/s10254-004-0033-5 [DOI] [PubMed] [Google Scholar]
- 116. Hecht G, Pothoulakis C, LaMont JT, Madara JL. 1988. Clostridium difficile toxin A perturbs cytoskeletal structure and tight junction permeability of cultured human intestinal epithelial monolayers. J Clin Invest 82:1516–1524. doi: 10.1172/JCI113760 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Hecht G, Koutsouris A, Pothoulakis C, LaMont JT, Madara JL. 1992. Clostridium difficile toxin B disrupts the barrier function of T84 monolayers. Gastroenterology 102:416–423. doi: 10.1016/0016-5085(92)90085-D [DOI] [PubMed] [Google Scholar]
- 118. Chumbler NM, Farrow MA, Lapierre LA, Franklin JL, Haslam D, Goldenring JR, Lacy DB. 2012. Clostridium difficile toxin B causes epithelial cell necrosis through an autoprocessing-independent mechanism. PLoS Pathog 8:e1003072. doi: 10.1371/journal.ppat.1003072 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Kelly CP, Becker S, Linevsky JK, Joshi MA, O’Keane JC, Dickey BF, LaMont JT, Pothoulakis C. 1994. Neutrophil recruitment in Clostridium difficile toxin A enteritis in the rabbit. J Clin Invest 93:1257–1265. doi: 10.1172/JCI117080 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120. Kim JM, Kim JS, Jung HC, Oh Y, Song IS, Kim CY. 2002. Differential expression and polarized secretion of CXC and CC chemokines by human intestinal epithelial cancer cell lines in response to Clostridium difficile toxin A . Microbiol Immunol 46:333–342. doi: 10.1111/j.1348-0421.2002.tb02704.x [DOI] [PubMed] [Google Scholar]
- 121. Savidge TC, Pan W-H, Newman P, O’brien M, Anton PM, Pothoulakis C. 2003. Clostridium difficile toxin B is an inflammatory enterotoxin in human intestine. Gastroenterology 125:413–420. doi: 10.1016/s0016-5085(03)00902-8 [DOI] [PubMed] [Google Scholar]
- 122. Kelly CP, Kyne L. 2011. The host immune response to Clostridium difficile. J Med Microbiol 60:1070–1079. doi: 10.1099/jmm.0.030015-0 [DOI] [PubMed] [Google Scholar]
- 123. Takamatsu H, Imamura A, Kodama T, Asai K, Ogasawara N, Watabe K. 2000. The yabG gene of Bacillus subtilis encodes a sporulation specific protease which is involved in the processing of several spore coat proteins. FEMS Microbiol Lett 192:33–38. doi: 10.1111/j.1574-6968.2000.tb09355.x [DOI] [PubMed] [Google Scholar]
- 124. Takamatsu H, Kodama T, Imamura A, Asai K, Kobayashi K, Nakayama T, Ogasawara N, Watabe K. 2000. The Bacillus subtilis yabG gene is transcribed by SigK RNA polymerase during sporulation, and yabG mutant spores have altered coat protein composition. J Bacteriol 182:1883–1888. doi: 10.1128/JB.182.7.1883-1888.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Kuwana R, Okuda N, Takamatsu H, Watabe K. 2006. Modification of GerQ reveals a functional relationship between Tgl and YabG in the coat of Bacillus subtilis spores. J Biochem 139:887–901. doi: 10.1093/jb/mvj096 [DOI] [PubMed] [Google Scholar]
- 126. Yamazawa R, Kuwana R, Takeuchi K, Takamatsu H, Nakajima Y, Ito K. 2022. Identification of the active site and characterization of a novel sporulation-specific cysteine protease YabG from Bacillus subtilis. J Biochem 171:315–324. doi: 10.1093/jb/mvab135 [DOI] [PubMed] [Google Scholar]
- 127. Kevorkian Y, Shirley DJ, Shen A. 2016. Regulation of Clostridium difficile spore germination by the CspA pseudoprotease domain. Biochimie 122:243–254. doi: 10.1016/j.biochi.2015.07.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Osborne MS, Brehm JN, Olivença C, Cochran AM, Serrano M, Henriques AO, Sorg JA. 2024. The Impact of YabG mutations on Clostridioides difficile spore germination and processing of spore substrates. Mol Microbiol 122:534–548. doi: 10.1111/mmi.15316 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Marini E, Olivença C, Ramalhete S, Aguirre AM, Ingle P, Melo MN, Antunes W, Minton NP, Hernandez G, Cordeiro TN, Sorg JA, Serrano M, Henriques AO. 2023. A sporulation signature protease is required for assembly of the spore surface layers, germination and host colonization in Clostridioides difficile. PLoS Pathog 19:e1011741. doi: 10.1371/journal.ppat.1011741 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Shrestha R, Cochran AM, Sorg JA. 2019. The requirement for co-germinants during Clostridium difficile spore germination is influenced by mutations in yabG and cspA. PLoS Pathog 15:e1007681. doi: 10.1371/journal.ppat.1007681 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Kuwana R, Takamatsu H, Watabe K. 2007. Expression, localization and modification of YxeE spore coat protein in Bacillus subtilis. J Biochem 142:681–689. doi: 10.1093/jb/mvm179 [DOI] [PubMed] [Google Scholar]
- 132. Takamatsu H, Imamura D, Kuwana R, Watabe K. 2009. Expression of yeeK during Bacillus subtilis sporulation and localization of YeeK to the inner spore coat using fluorescence microscopy. J Bacteriol 191:1220–1229. doi: 10.1128/JB.01269-08 [DOI] [PMC free article] [PubMed] [Google Scholar]