Next Article in Journal
[D-Leu1]MC-LR and MC-LR: A Small–Large Difference: Significantly Different Effects on Phaseolus vulgaris L. (Fabaceae) Growth and Phototropic Response after Single Contact during Imbibition with Each of These Microcystin Variants
Next Article in Special Issue
Staphylococcus aureus Infection and Persistence in Chronic Rhinosinusitis: Focus on Leukocidin ED
Previous Article in Journal
Bedside Coagulation Tests in Diagnosing Venom-Induced Consumption Coagulopathy in Snakebite
Previous Article in Special Issue
Consequences of Metabolic Interactions during Staphylococcus aureus Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Staphylococcal Enterotoxin C—An Update on SEC Variants, Their Structure and Properties, and Their Role in Foodborne Intoxications

1
Institute for Food Safety and Hygiene, Vetsuisse Faculty, University of Zürich, 8057 Zürich, Switzerland
2
Laboratory of Food Microbiology, Institute of Food, Nutrition and Health, ETH Zürich, 8092 Zürich, Switzerland
3
Division of Applied Microbiology, Department of Chemistry, Lund University, 22100 Lund, Sweden
*
Author to whom correspondence should be addressed.
Toxins 2020, 12(9), 584; https://doi.org/10.3390/toxins12090584
Submission received: 3 August 2020 / Revised: 7 September 2020 / Accepted: 8 September 2020 / Published: 10 September 2020
(This article belongs to the Special Issue Staphylococcus aureus Toxins: Promoter or Handicap during Infection)

Abstract

:
Staphylococcal enterotoxins are the most common cause of foodborne intoxications (staphylococcal food poisoning) and cause a wide range of diseases. With at least six variants staphylococcal enterotoxin C (SEC) stands out as particularly diverse amongst the 25 known staphylococcal enterotoxins. Some variants present unique and even host-specific features. Here, we review the role of SEC in human and animal health with a particular focus on its role as a causative agent for foodborne intoxications. We highlight structural features unique to SEC and its variants, particularly, the emetic and superantigen activity, as well as the roles of SEC in mastitis and in dairy products. Information about the genetic organization as well as regulatory mechanisms including the accessory gene regulator and food-related stressors are provided.
Key Contribution: We summarize recent findings on SEC with a particular focus on protein structure, emetic activity, and superantigenic properties. Special attention is given to the large number of host-specific variants of this toxin.

1. Introduction

Foodborne illnesses are one of the world’s leading health issues. They are estimated to cause 420,000 deaths every year and generate health costs and economic losses in the range of 110 billion USD worldwide [1]. One of the most common causative agents for food intoxications are staphylococcal enterotoxins (SEs). These exotoxins are preformed by Staphylococcus aureus in food and cause intoxication upon ingestion. In the EU alone, 114 cases of foodborne outbreaks were attributed to these heat-stable SEs in 2018 [2]. Symptoms of staphylococcal food poisoning (SFP) include violent vomiting, diarrhea, fever, and unspecific symptoms like headache and nausea. Due to the generally-quick recovery of patients, the number of cases is likely being underestimated [3,4].
SEs are produced by S. aureus during growth, alongside a number of other virulence factors. The organism can not only cause food intoxications but also various infectious diseases including toxic shock syndrome (TSS) [5,6,7]. Staphylococcal virulence factors include exopolysaccharides, surface-associated protein adhesins, immune modulators, and exoproteins including a variety of toxins [8]. Staphylococcal virulence factors [9,10] and their role in both infectious diseases [6,11] and food intoxications [12,13] have been previously reviewed. The most relevant secreted toxins are SEs, hemolysins, leukotoxins, exfoliative toxins, and toxic shock syndrome toxin (TSST-1) [14]. So far, 25 SEs (SEA–SElZ) have been described, excluding variants and TSST-1 (formerly SEF), but new types are frequently discovered [12]. Non-emetic toxins or not-yet-tested ones are referred to as SE-like (SEl) [15]. These peptides are all classified as pyrogenic toxin superantigens (SAgs) and have the ability to mobilize large proportions of T-cells (20–30%) [16,17,18]. The first enterotoxin SEA was isolated in 1959 [19,20], shortly followed by SEB and SEC [21,22]. All newly-discovered enterotoxins have subsequently been named alphabetically. The genetic location of SEs differs greatly between toxins and can even vary within the same type of SE (Table 1). SEA–SEE are considered "classical enterotoxins", while SEG–SElZ are termed "new enterotoxins". Historically, classical toxins were demonstrated to be emetically active in rhesus monkey feeding assays, while new toxins were either not emetic or had not been tested in monkeys [12,23]. However, this distinction has become obsolete since many new toxins were later recognized as emetically active in rhesus monkey feeding assays [24], and new toxins were found as causative agents for SFP [25,26]. Additionally, smaller mammals such as ferrets or the house musk shrew were proposed as models to simplify testing [27,28]. So far, commercial antibodies can only be obtained for classical SEs.
Originally, SEs were differentiated by immunological methods, whereas today it is recommended to identify the relationship of new toxins via sequence homology (where >90% sequence homology equals a toxin variant and <90% homology determines a new toxin) [15]. SEC takes on a special role in this terminology, since several variants, often even host-specific ones, have been described [29,30]. In particular, the ruminant-adapted toxin variants SECbovine and SECovine show the exceptional ability of S. aureus to adapt to different niches. SEC is by far the most-frequently-isolated toxin in animals suffering from mastitis [31,32,33,34]. Additionally, SEC alone was shown to cause inflammation, proinflammatory cytokine production, and tissue damage in mammary glands. SEC might therefore play an important role in the development of mastitis associated with S. aureus infection. [31]. There are some reports of SEC also being involved in human post-partum mastitis. [35]. Contaminated ruminant milk can also provide an entry-point for S. aureus or its toxins into the food-chain. In addition to TSS, mastitis, and food intoxication SEC seems to play a critical role in the development of infective endocarditis and atopic dermatitis [36,37]. Other diseases that are associated with SEs in general include severe nasal polyposis, perineal erythema, desquamative inflammatory vaginitis, and sudden infant death syndrome [38].
An overview of different SEC variants and their relationships is given in Table 1 and Figure 1. While other Staphylococcus species can produce SEC and variants such as SECcanine [39], this review will focus exclusively on S. aureus SECs. The great variety of SEs and their genomic location further complicates the intricately-intertwined regulatory pathways of S. aureus virulence factors. Previously, SEs in general [4,16,40,41], as well as SEB specifically and its role as a potential bioweapon [42,43] have been reviewed in depth. Information on SEC and its pathogenic role in food intoxication and infection is, however, limited. Potential differences in virulence and toxicity for SEC variants further obscure research efforts. Here, we provide an overview of the role of SEC in foodborne intoxications and clinical manifestations and summarize recent findings in SEC characterization with special regard to the many facets of its variants.

2. SEC in Food Intoxication

A striking feature of SEs is their emetic activity. It provokes vomiting, the key symptom in SFP. Some SEs, but not SEC also cause diarrhea [95]. The exact biological purpose of these properties is still unclear. A function in pathogen spread through emesis or diarrhea seems unlikely since S. aureus survives very poorly in gastric juice [96]. Still, it has been suggested that S. aureus intestinal carriage rate ranges from 8–31% in the healthy population [97]. It is therefore possible that the gastric activity of enterotoxins supports persistence or epithelial barrier invasion of S. aureus.
Most SFP outbreaks are attributed to SEA, although there might be some bias since enterotoxins other than SEA–SEE cannot be detected using commercial kits. Many strains produce multiple toxins and it is often unclear which enterotoxin was the cause for SFP. Synergistic effects of different toxins are also likely. Some outbreaks could be traced back to SEC presumably being the only SE involved [4,98,99,100]. SFP can occur with any food that provides sufficient carbon and amino acid sources for S. aureus growth. The pathogen is usually introduced by food handlers and produces SEs if conditions allow bacterial growth in the food matrix [12]. Most cases of SFP can be prevented by adequate hygiene measures and intact cooling chains [101]. Due to the heat tolerance of SEs, reheating foods may eliminate S. aureus but SEs remain emetically active [3]. SEs are generally very resilient towards external stressors such as heat, acidity, and gastric enzymes [16]. Once SEs reach the small intestine, they enter the lamina propria through mucus-producing goblet cells or epithelial cells [102]. This process may be facilitated in the presence of other S. aureus virulence factors [103]. SEA was shown to stimulate 5-hydroxytryptamine (serotonin) and histamine release from mast cells [104,105]. However, it is generally assumed that this principle applies for all emetic SEs in possession of the disulphide loop (see chapter 4). Serotonin acts on the vagus nerve by evoking an emetic response [106]. The role of T-cell and neutrophil activation is unclear, but may be a contributing factor in gut epithelial invasion of S. aureus [45]. The involvement of the vagus nerve was demonstrated in very early experiments showing that monkeys did not present emetic symptoms upon SE ingestion after a vagotomy [107]. The basic principle of SE-induced emesis is illustrated in Figure 2.
Whether SEC variants have a different emetic potential is unclear. SEC1 [108], SEC2 [21], and SEC3 [60] have been tested in the monkey-feeding assay, but administered amounts varied, and different extraction methods were applied. SEC2 was also tested in the house musk shrew [27]. Therefore, the results cannot be directly compared.

SEC in Milk and Dairy Products

Intoxications caused by SEs are the most common cause of food poisoning after consumption of raw milk or products made thereof. SEC and SEA are the most-commonly-occurring SEs in milk and dairy products [13,109,110,111,112]. In raw milk samples from Sweden, S. aureus was found with a prevalence of up to 71% [113]. This likely stems from the frequent occurrence of SEC in intramammary infections in milk-producing animals [31].
The effect of milk on SE production has been investigated experimentally. When S. aureus was grown in milk the expression of SEC was significantly reduced. Downregulation of the agr system (see chapter 5) likely contributed to the observed reduction, but other factors are also expected to be involved [114]. The effect of heat treatment on the activity of staphylococcal enterotoxins of type A, B, and C in milk was investigated for pasteurization temperatures [115] and higher temperatures [116]. The amount of detectable SEC could be reduced by heat treatment at 100 °C and above. Another study investigated the production of SECbovine in milk and during cheesemaking. S. aureus numbers increased during cheesemaking but did generally not reach the >108 CFU/mL required for SECbovine detection in this food matrix. The influence of the added starter culture was not investigated in detail [117]. Recently, the impact of Weissella paramesenteroides GIR16L4 or Lactobacillus rhamnosus D1 or both together used as starter cultures on the expression of S. aureus SEC was examined. The starter cultures were not able to reduce S. aureus growth, but they influenced toxin expression in some strains [118]. Lactic acid bacteria or their metabolites could impact quorum sensing of S. aureus and therefore influence agr-regulated SEs. In another trial, Staphylococcus vitulinus was used as a starter culture to successfully inhibit SEC-producing S. aureus growth in a barbeque cheese production facility [119].

3. Superantigenic Activity of SEC

All SEs are classified as SAgs. In contrast to normal antigens, they bind to MHCII in a location adjacent to the peptide groove. They thus stimulate T-cell receptors (TCRs) in a non-specific way by cross-bridging them with major histocompatibility complex class II (MHCII) on antigen-presenting cells (APC). The resulting global overstimulation of T-cells interferes with immune system functions that normally counteract bacterial infections [38]. In infectious diseases SAgs therefore contribute substantially to transcytosis and immune system evasion [120]. In addition, SEC activation of T-lymphocytes has been exploited for anti-cancer drug development [121]. In spite of advances in SEC research and related drug development, the role of superantigenic activity in food intoxications remains unknown.
The exact location of MHCII binding varies, depending on the SE. Group II antigens, including SEC, bind to the low-affinity α-chain [94]. The resulting mobilization efficiency is 10–100-fold lower for these SAgs than for group III or V SAgs. This is compensated for by the considerably-higher production of these compounds by S. aureus [38].
TCR binding is specific to the variable region Vβ of the receptor [17,122] (Figure 3). This results in massive Vβ-dependent T-cell proliferation and subsequent release of pro-inflammatory cytokines [38,94]. Which Vβ subfamilies are stimulated, again depends on the SAg and can even vary between different variants [92]. The following human Vβ-specificity was found for variants SEC1–3: C1: 3.2, 6.4, 6.9, 12, 15.1; C2: 12, 13, 14, 15, 17, 20; and C3: 5.1, 12 [123,124,125]. SECbovine has been shown to specifically activate the Vβ repertoire of cattle [126]. Furthermore, when different SEC variants were compared, SECovine strains showed a particularly-strong response in cattle peripheral blood mononuclear cells (PBMCs) [30]. This adaptation of SAgs to mobilize specific T-cell populations contributes to the capacity of S. aureus to adapt to different host species [92]. Even in variants not linked to different hosts, functional differences are apparent. When compared in an MHCII-deficient cell line, SEC1 was able to still induce T-cell proliferation, but SEC2 and SEC3 were not. A few specific amino acids near the NH2-terminus seem to be responsible for this difference [127]. However, many of the specific differences between variants with regard to the interferon response are not fully understood.
Another factor adding to the immunostimulatory properties of SEs is their ability to co-bind CD28 on T-cells [128]. This has only been demonstrated for SEB [129] but the structural resemblance with SEC suggests that these findings may also apply to SEC.
One of the consequences of the overshooting T-cell proliferation is potentially fatal toxic shock syndrome (TSS). TSS was originally attributed to toxic shock syndrome toxin (TSST-1) and the use of high-absorbing tampons [130]. However, it was demonstrated that non-menstrual TSS can also be caused by SEC and other SEs [81,131]. Long-term exposure to SEs and their superantigenic activity has reportedly been associated with auto-immune diseases such as psoriasis, atopic dermatitis, systemic lupus erythematosus, and, potentially, Kawasaki disease [94]. Food intoxication symptoms like fever and malaise can be attributed to superantigenic activity.

4. Physical and Chemical SEC Protein Properties

All SEs share almost identical structural features, although their amino acid sequences vary. In addition, they are structurally related to streptococcal SAgs such as streptococcal pyrogenic exotoxins (SPEs). They are water-soluble peptides with a length of 220–270 aa and have a molecular weight of ~22–29 kDa [4]. Mature SEC, specifically, is 239 aa in length (not including the 27 aa signal peptide at the N-terminus) and has a molecular weight of 27.5–27.6 kDa, depending on the variant.
Sequence similarity of SEs ranges from ~20% to >95%, whereby 15% of amino acids mostly located on the central and C-terminal portions are entirely conserved [16,40]. In regard to its sequence, SEC is most similar to SEB [3]. Within SEC, variants SEC2 and SEC4 are almost identical with only two aa differences, while SEC3 and SECovine differ the most with 18 aa. Sequence similarity amongst SEC variants is at least 93% (Figure 1, Figure S1). Interestingly, the signal peptide of SEC3 is identical to the one of SEB and therefore differs substantially from the other SEC variants. However, to date there is no indication of a different secretion mechanism [132,133,134,135].
As for many other SEs, the crystal structure of some SEC variants has been resolved [136,137,138,139]. The peptide consists of two unequal domains. The larger one contains a β-grasp fold formed by five β-sheets and an α-helix (β6–β12 & α5). The smaller one comprises a Greek key motif of β-sheets, also known as the oligosaccharide/oligonucleotide fold (OB-fold) found in numerous other bacterial toxins and small α-helixes (β1–β5 and α3, α4, and α6) [4]. The binding sites involved in superantigen activity of SEs have been partially identified [140,141]. The shallow cavity between the two domains binds TCR while major histocompatibility complex class II (MHCII) binding is guided by a region in the N-terminal domain (Figure 1) [4]. It is believed that subtype-specific antigenic epitopes are determined by the N-terminus of SECs, while conserved C-terminal regions define antigenic epitopes shared with other pyrogenic toxins [59].
Like some other SEs and SPEs, SECs have two zinc-binding domains close to the MHCII binding site possibly involved in dimerization and MHCII binding mode [137,138,139]. However, SEC can bind to MHC class II molecules outside the groove on the flanking helix from the α chain via the zinc-independent MHCII binding site [16,138,142].
A distinct disulphide loop (also termed cystine loop) present in all emetic SEs and absent in some weakly- or non-emetic types has been implicated in the toxin’s emetic activity (in the N-terminus between β4 and β5). Structural analyses in SEC have confirmed the important role of the disulphide loop in emesis. Substitution of the two cysteines to alanine in SEC1 resulted in loss of emetic activity, however changing the residues to serine did not [143]. This was explained by the ability of serine hydrogen bonds to stabilize the critical loop structure as in the case of the disulphide bond [143]. These findings highlight the importance of the structural conformation, while the chemical makeup itself seems not to be crucial. Therefore, it is the tertiary structure rather than the disulphide bond itself that is responsible for emesis induction.
Mutational analysis in SEA revealed that emetic and superantigenic properties can be attributed to different regions of the peptide, hence separating emesis and superantigenic activity as two different functions of SEs [144,145]. In the case of SEC, it was shown that a lack of the disulphide bond did not affect superantigen activity in an early experiment [143]. However, some articles suggest that even though emetic and superantigen properties are localized in two different regions, they still partly correlate with each other [145]. In particular, superantigen activity seems to depend on multiple domains whereas emetic activity can be allocated more clearly to a specific region [13]. A later study found that regions 21–51 and 81–100 were essential for both emesis and superantigenicity in SEA [146]. To what extent these finding apply to SEC is unclear.

5. Genetic Localization and Regulation of SEC

Genes coding for SEC (sec) are generally located on a S. aureus pathogenicity islands (SaPIs), but plasmids carrying sec have also been described [30,55,57]. SaPIs are mobile genetic elements that harbor many virulence factors and likely contribute to S. aureus evolution via horizontal gene transfer. They derived from prophages and require helper phages for transmission [93]. The sec gene has been described in SaPIbov1, SaPIn1/m1, SaPImw2, SaPIov1, and SePI1. Moreover a BLAST search resulted in SaPINuSaα2, SaPITokyo, and SaPIbov5 (Table 1) [56]. Incidentally, NCBI GenBank contains multiple WGS that include sec in regions containing SaPI-related features although they may not be annotated as such. It is often co-localized with tst-1, the gene coding for toxic shock syndrome toxin-1 (TSST-1) [57].
The expression of SEC is highest at late exponential to stationary phase [147]. It is regulated by the quorum-sensing system of the accessory gene regulator (agr) that uses autoinducing peptides (AIP). It acts on toxin transcription indirectly via RNAIII that represses the repressor of toxins (rot) [148,149,150,151,152]. Additionally, SarA, σB, and SaeRS may play a role in regulation of SEC when environmental factors change (Figure 4) [45,153,154]. Food-related stressors such as glucose or NaCl have been shown to influence SEC expression. In an early study, glucose led to reduced extracellular SEC concentrations. Furthermore, an intact agr system was not required for the effect of glucose on sec expression [149]. In a later study, the expression of sec was shown to be regulated in response to high NaCl concentrations. The results demonstrated that osmoregulation of SEC occurs at the level of mRNA independently of an intact agr allele. Osmoprotective compounds did enhance SEC expression [148]. In an older study on the effect of NaCl and pH on SEC production it was not possible to demonstrate SEC production in broths with 12% NaCl and a pH range of 4.50 to 8.55 [155].
In conclusion, under both glucose and NaCl stress sec mRNA and SEC protein levels were considerably reduced [148,149,155]. It is unclear whether the different genetic locations of SEC variants influence transcriptional regulation. Many external stressors and their influence on SEC expression have not yet been investigated.

6. Challenges and Future Prospects

Many challenges arise from the great variety of SEs. Firstly, new SEs cannot yet be detected by commercial assays. Therefore, reports on the frequency of SEs associated with intoxication and disease are likely biased towards classical SEs. Secondly, the influence of the food matrix and host factors can influence toxin expression in strain-specific, toxin-specific, and even variant-specific ways. Additionally, the genetic location of SEs may play a role in genetic regulation. Nonetheless, many studies do not account for these aspects. Legislative documents limiting S. aureus occurrence by CFU/gram food, regardless of the respective toxin expression, reflect these shortcomings.
For SEC variants in particular, it remains unknown whether genetic location or external stressors result in variant-specific responses in genetic regulation. Furthermore, although agr undoubtedly plays an important role in SEC regulation, additional regulatory elements should be considered in future studies. Whether SEC variants differ in their effects on human and animal health has not been resolved. The different Vβ-specificities of SEC variants likely trigger slightly different superantigenic responses. Hence, some variants might be more relevant in acute diseases while others could be implicated with chronic or persistent progressions. Additionally, variants have never been compared directly with respect to their emetic activity. Therefore, correlations between variants and emetic dosage cannot be provided.
In conclusion, aspects such as toxin variants, strain-specific responses, and influences of external stressors should be considered when investigating SEC expression in the future, and results should not be generalized. Taking all of these factors into consideration will contribute to predicting and maintaining high food safety standards and improving human and animal health.

Supplementary Materials

The following are available online at https://www.mdpi.com/2072-6651/12/9/584/s1, Table S1: Accession numbers; Figure S1: SEC protein sequence alignments.

Author Contributions

Conceptualization, D.E. and S.J.; data curation, D.E.; writing—original draft preparation, D.E.; writing—review and editing, D.E., S.J., J.S., and M.S.; visualization, D.E.; supervision, S.J., J.S., and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Food-Borne Disease Burden Epidemiology Reference Group. WHO Estimates of the Global Burden of Foodborne Diseases; WHO Press: Geneva, Switzerland, 2015. [Google Scholar]
  2. EFSA. The European Union One Health 2018 Zoonoses Report; EFSA Journal: Parma, Italy, 2019; Volume 17. [Google Scholar]
  3. Le Loir, Y.; Baron, F.; Gautier, M. Staphylococcus aureus and food poisoning. Genet. Mol. Res. 2003, 2, 63–76. [Google Scholar] [PubMed]
  4. Argudín, M.Á.; Mendoza, M.C.; Rodicio, M.R. Food poisoning and Staphylococcus aureus enterotoxins. Toxins 2010, 2, 1751–1773. [Google Scholar] [CrossRef] [PubMed]
  5. Tong, S.Y.C.; Davis, J.S.; Eichenberger, E.; Holland, T.L.; Fowler, V.G. Staphylococcus aureus infections: Epidemiology, pathophysiology, clinical manifestations, and management. Clin. Microbiol. Rev. 2015, 28, 603–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Thomer, L.; Schneewind, O.; Missiakas, D. Pathogenesis of Staphylococcus aureus Bloodstream Infections. Annu. Rev. Pathol. Mech. Dis. 2016, 11, 343–364. [Google Scholar] [CrossRef] [Green Version]
  7. Mccormick, J.K.; Yarwood, J.M.; Schlievert, P.M. Toxic shock syndrome and bacterial superantigens: An Update. Ann. Rev. Microbiol. 2001, 55, 77–104. [Google Scholar] [CrossRef]
  8. Junecko, J.M.; Zielinska, A.K.; Mrak, L.N.; Ryan, D.C.; Graham, J.W.; Smeltzer, M.S.; Lee, C.Y. Transcribing virulence in Staphylococcus aureus. World J. Clin. Infect. Dis. 2012, 2, 63. [Google Scholar] [CrossRef]
  9. Kane, T.L.; Carothers, K.E.; Lee, S.W. Virulence factor targeting of the bacterial pathogen Staphylococcus aureus for vaccine and therapeutics. Curr. Drug Targets 2018, 176, 139–148. [Google Scholar] [CrossRef]
  10. Massey, R.C.; Horsburgh, M.J.; Lina, G.; Höök, M.; Recker, M. The evolution and maintenance of virulence in Staphylococcus aureus: A role for host-to-host transmission? Nat. Rev. Microbiol. 2006, 4, 953–958. [Google Scholar] [CrossRef]
  11. Lacey, K.A.; Geoghegan, J.A.; McLoughlin, R.M. The role of Staphylococcus aureus virulence factors in skin infection and their potential as vaccine antigens. Pathogens 2016, 5, 22. [Google Scholar] [CrossRef] [Green Version]
  12. Fetsch, A.; Johler, S. Staphylococcus aureus as a foodborne pathogen. Curr. Clin. Microbiol. 2018, 9, 1–8. [Google Scholar] [CrossRef]
  13. Benkerroum, N. Staphylococcal enterotoxins and enterotoxin-like toxins with special reference to dairy products: An overview. Crit. Rev. Food Sci. Nutr. 2018, 58, 1943–1970. [Google Scholar] [CrossRef]
  14. Kong, C.; Neoh, H.M.; Nathan, S. Targeting Staphylococcus aureus toxins: A potential form of anti-virulence therapy. Toxins 2016, 8, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lina, G.; Bohach, G.A.; Nair, S.P.; Hiramatsu, K.; Jouvin-Marche, E.; Mariuzza, R. Standard nomenclature for the superantigens expressed by Staphylococcus. J. Infect. Dis. 2004, 189, 2334–2336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pinchuk, I.V.; Beswick, E.J.; Reyes, V.E. Staphylococcal enterotoxins. Toxins 2010, 2, 2177–2197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Choi, Y.; Kotzin, B.; Herront, L.; Callahan, J.; Marrack, P.; Kappler, J. Interaction of Staphylococcus aureus toxin “superantigens” with human T cells. Proc. Natl. Acad. Sci. USA 1989, 86, 8941–8945. [Google Scholar] [CrossRef] [Green Version]
  18. Schlievert, P.M.; Bohach, G.A.; Ohlendorf, D.H.; Stauffacher, C.V.; Leung, D.Y.M.; Murray, D.L.; Earhart, C.A.; Jablonski, L.M.; Hoffmann, M.L.; Chi, Y.I. Molecular structure of Staphylococcus and Streptococcus superantigens. J. Clin. Immunol. 1995, 15, 4–10. [Google Scholar] [CrossRef]
  19. Casman, E.P. Further serological studies of staphylococcal enterotoxin. J. Bacteriol. 1960, 79, 849–856. [Google Scholar] [CrossRef] [Green Version]
  20. Bergdoll, M.S.; Sugiyama, H.; Dack, G.M. Staphylococcal enterotoxin. I. Purification. Arch. Biochem. Biophys. 1959, 85, 62–69. [Google Scholar] [CrossRef]
  21. Bergdoll, M.S.; Borja, C.R.; Avena, R.M. Identification of a new enterotoxin as enterotoxin C. J. Bacteriol. 1965, 90, 1481–1485. [Google Scholar] [CrossRef] [Green Version]
  22. Casman, E.P.; Bergdoll, M.S.; Robinson, J. Designation of staphylococcal enterotoxins. J. Bacteriol. 1963, 85, 715–716. [Google Scholar] [CrossRef] [Green Version]
  23. Bergdoll, M.S. Monkey feeding test for staphylococcal enterotoxin. Methods Enzymol. 1988, 165, 324–333. [Google Scholar] [CrossRef] [PubMed]
  24. Omoe, K.; Hu, D.-L.; Ono, H.K.; Shimizu, S.; Takahashi-Omoe, H.; Nakane, A.; Uchiyama, T.; Shinagawa, K.; Imanishi, K. Emetic Potentials of Newly Identified Staphylococcal Enterotoxin-Like Toxins. Infect. Immun. 2013, 81, 3627–3631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ikeda, T.; Tamate, N.; Yamaguchi, K.; Makino, S.I. Mass outbreak of food poisoning disease caused by small amounts of staphylococcal enterotoxins A and H. Appl. Environ. Microbiol. 2005, 71, 2793–2795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Jørgensen, H.J.; Mathisen, T.; Løvseth, A.; Omoe, K.; Qvale, K.S.; Loncarevic, S. An outbreak of staphylococcal food poisoning caused by enterotoxin H in mashed potato made with raw milk. FEMS Microbiol. Lett. 2005, 252, 267–272. [Google Scholar] [CrossRef] [Green Version]
  27. Hu, D.; Omoe, K.; Shimoda, Y.; Nakane, A.; Shinagawa, K. Induction of emetic response to staphylococcal enterotoxins in the house musk shrew (Suncus murinus). Infect. Immun. 2003, 71, 567–570. [Google Scholar] [CrossRef] [Green Version]
  28. Wright, A.; Andrews, P.L.R.; Titball, R.W. Induction of emetic, pyrexic, and behavioral effects of Staphylococcus aureus enterotoxin B in the ferret. Infect. Immun. 2000, 68, 2386–2389. [Google Scholar] [CrossRef] [Green Version]
  29. Johler, S.; Sihto, H.-M.; Macori, G.; Stephan, R. Sequence variability in staphylococcal enterotoxin genes seb, sec, and sed. Toxins 2016, 8, 169. [Google Scholar] [CrossRef] [Green Version]
  30. Marr, J.C.; Lyon, J.D.; Roberson, J.R.; Lupher, M.; Davis, W.C.; Bohach, G.A. Characterization of novel type C staphylococcal enterotoxins: Biological and evolutionary implications. Infect. Immun. 1993, 61, 4254–4262. [Google Scholar] [CrossRef] [Green Version]
  31. Fang, R.; Cui, J.; Cui, T.; Guo, H.; Ono, H.K.; Park, C.H.; Okamura, M.; Nakane, A.; Hu, D.L. Staphylococcal enterotoxin C is an important virulence factor for mastitis. Toxins 2019, 11, 141. [Google Scholar] [CrossRef] [Green Version]
  32. Niskanen, A.; Koranen, L.; Roine, K. Staphylococcal enterotoxin and thermonuclease production during induced bovine mastitis and the clinical reaction of enterotoxin in udders. Infect. Immun. 1978, 19, 493–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Fitzgerald, J.R. Livestock-associated Staphylococcus aureus: Origin, evolution and public health threat. Trends Microbiol. 2012, 20, 192–198. [Google Scholar] [CrossRef] [PubMed]
  34. Orden, J.A.; Goyache, J.; Hernández, J.; Doménech, A.; Suárez, G.; Gómez-Lucia, E. Detection of enterotoxins and TSST-1 secreted by Staphylococcus aureus isolated from ruminant mastitis. Comparison of ELISA and immunoblot. J. Appl. Bacteriol. 1992, 72, 486–489. [Google Scholar] [CrossRef] [PubMed]
  35. Franck, K.T.; Gumpert, H.; Olesen, B.; Larsen, A.R.; Petersen, A.; Bangsborg, J.; Albertsen, P.; Westh, H.; Bartels, M.D. Staphylococcal aureus Enterotoxin C and enterotoxin-like L associated with post-partum mastitis. Front. Microbiol. 2017, 8, 173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kinney, K.J.; Tran, P.M.; Gibson-Corley, K.N.; Forsythe, A.N.; Kulhankova, K.; Salgado-Pabón, W. Staphylococcal enterotoxin C promotes Staphylococcus aureus infective endocarditis independent of superantigen activity. bioRxiv 2019. [Google Scholar] [CrossRef] [Green Version]
  37. Bunikowski, R.; Mielke, M.E.A.; Skarabis, H.; Worm, M.; Anagnostopoulos, I.; Kolde, G.; Wahn, U.; Renz, H. Evidence for a disease-promoting effect of Staphylococcus aureus-derived exotoxins in atopic dermatitis. J. Allergy Clin. Immunol. 2000, 105, 814–819. [Google Scholar] [CrossRef]
  38. Spaulding, A.R.; Salgado-Pabón, W.; Kohler, P.L.; Horswill, A.R.; Leung, D.Y.M.M.; Schlievert, P.M. Staphylococcal and streptococcal superantigen exotoxins. Clin. Microbiol. Rev. 2013, 26, 422–447. [Google Scholar] [CrossRef] [Green Version]
  39. Edwards, V.M.; Deringer, J.R.; Callantine, S.D.; Deobald, C.F.; Berger, P.H.; Kapur, V.; Stauffacher, C.V.; Bohach, G.A. Characterization of the canine type C enterotoxin produced by Staphylococcus intermedius pyoderma isolates. Infect. Immun. 1997, 65, 2346–2352. [Google Scholar] [CrossRef] [Green Version]
  40. Vasconcelos, N.G.; Cunha, M.L.R.S. Staphylococcal enterotoxins: Molecular aspects and detection methods. J. Public Health Epidemiol. 2010, 2, 29–42. [Google Scholar]
  41. Hennekinne, J.A.; De Buyser, M.L.; Dragacci, S. Staphylococcus aureus and its food poisoning toxins: Characterization and outbreak investigation. FEMS Microbiol. Rev. 2012, 36, 815–836. [Google Scholar] [CrossRef] [Green Version]
  42. Fries, B.C.; Varshney, A.K. Bacterial toxins—Staphylococcal enterotoxin B. Microb. Spectr. 2013, 1, 1–21. [Google Scholar] [CrossRef] [Green Version]
  43. Krakauer, T.; Stiles, B.G. The staphylococcal enterotoxin (SE) family: SEB and siblings. Virulence 2013, 4, 759–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Susilo, Y.B. Staphylococcus aureus Toxins—Influence on Food Safety and Animal Health. Ph.D. Thesis, Lund University, Lund, Sweden, 2016. [Google Scholar]
  45. Fisher, E.L.; Otto, M.; Cheung, G.Y.C.C. Basis of virulence in enterotoxin-mediated staphylococcal food poisoning. Front. Microbiol. 2018, 9, 436. [Google Scholar] [CrossRef] [PubMed]
  46. Betley, M.J.; Mekalanos, J.J. Staphylococcal enterotoxin A is encoded by phage. Science 1985, 229, 185–187. [Google Scholar] [CrossRef] [PubMed]
  47. Betley, M.J.; Mekalanos, J.J. Nucleotide sequence of the type A staphylococcal enterotoxin gene. J. Bacteriol. 1988, 170, 34–41. [Google Scholar] [CrossRef] [Green Version]
  48. Kuroda, M.; Ohta, T.; Uchiyama, I.; Baba, T.; Yuzawa, H.; Kobayashi, I.; Cui, L.; Oguchi, A.; Aoki, K.; Nagai, Y.; et al. Whole genome sequencing of meticillin-resistant Staphylococcus aureus. Lancet 2001, 357, 1225–1240. [Google Scholar] [CrossRef]
  49. Borst, D.W.; Betley, M.J. Phage-associated differences in staphylococcal enterotoxin A gene (sea) expression correlate with sea allele class. Infect. Immun. 1994, 62, 113–118. [Google Scholar] [CrossRef] [Green Version]
  50. Shafer, W.M.; Iandolo, J.J. Chromosomal locus for staphylococcal enterotoxin B. Infect. Immun. 1978, 20, 273–278. [Google Scholar] [CrossRef] [Green Version]
  51. Stevens, M.J.A.; Stephan, R.; Johler, S. Complete and assembled genome sequence of Staphylococcus aureus RKI4, a food-poisoning strain exhibiting a novel S. aureus pathogenicity island carrying seb. Genome Announc. 2015, 3, 2015. [Google Scholar] [CrossRef] [Green Version]
  52. Shalita, Z.; Hertman, I.; Sarid, S. Isolation and characterization of a plasmid involved with enterotoxin B production in Staphylococcus aureus. J. Bacteriol. 1977, 129, 317–325. [Google Scholar] [CrossRef] [Green Version]
  53. Altboum, Z.; Hertman, I.; Sarid, S. Penicillinase plasmid-linked genetic determinants for enterotoxins B and C1 production in Staphylococcus aureus. Infect. Immun. 1985, 47, 514–521. [Google Scholar] [CrossRef] [Green Version]
  54. Sato’o, Y.; Omoe, K.; Ono, H.K.; Nakane, A.; Hu, D.-L. A novel comprehensive analysis method for Staphylococcus aureus pathogenicity islands. Microbiol. Immunol. 2013, 57, 91–99. [Google Scholar] [CrossRef] [PubMed]
  55. Betley, M.J.; Bergdoll, M.S. Staphylococcal enterotoxin type C genes not associated with extrachromosomal DNA. In Proceedings of the Annual Meeting ASM, Dallas, TX, USA, 1–6 March 1981; Volume 38, p. 49. [Google Scholar]
  56. Novick, R.P.; Christie, G.E.; Penades, J.R.; Penadés, J.R. The phage-related chromosomal islands of Gram-positive bacteria. Nat. Rev. Microbiol. 2010, 8, 541–551. [Google Scholar] [CrossRef] [PubMed]
  57. Fitzgerald, J.R.; Monday, S.R.; Foster, T.J.; Bohach, G.A.; Hartigan, P.J.; Meaney, W.J.; Smyth, C.J. Characterization of a putative pathogenicity island from bovine Staphylococcus aureus encoding multiple superantigens. J. Bacteriol. 2001, 183, 63–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Bergdoll, M.S.; Robbins, R.N.; Weiss, K.; Borja, C.R.; Huang, Y.; Chu, F.S. The staphylococcal enterotoxins: Similarities. Contrib. Microbiol. Immunol. 1973, 1, 390–396. [Google Scholar]
  59. Carolyn, J.H.; Gregory, A.; Bohach, S.P.H. Nucleotide sequence of the staphlyococcal enterotoxin C3 gene: Sequence comparison of all three Type C staphylococcal enterotoxins. Mol. Gen. Genet. 1990, 3, 329–333. [Google Scholar]
  60. Reiser, R.F.; Robbins, R.N.; Noleto, A.L.; Khoe, G.P.; Bergdoll, M.S. Identification, purification, and some physicochemical properties of staphylococcal enterotoxin C3. Infect. Immun. 1984, 45, 625–630. [Google Scholar] [CrossRef] [Green Version]
  61. Baba, T.; Takeuchi, F.; Kuroda, M.; Yuzawa, H.; Aoki, K.I.; Oguchi, A.; Nagai, Y.; Iwama, N.; Asano, K.; Naimi, T.; et al. Genome and virulence determinants of high virulence community-acquired MRSA. Lancet 2002, 359, 1819–1827. [Google Scholar] [CrossRef]
  62. Viana, D.; Blanco, J.; Tormo-Más, M.Á.; Selva, L.; Guinane, C.M.; Baselga, R.; Corpa, J.M.; Lasa, Í.; Novick, R.P.; Fitzgerald, J.R.; et al. Adaptation of Staphylococcus aureus to ruminant and equine hosts involves SaPI-carried variants of von Willebrand factor-binding protein. Mol. Microbiol. 2010, 77, 1583–1594. [Google Scholar] [CrossRef]
  63. Hajek, V. Identification of enterotoxigenic staphylococci from sheep and sheep cheese. Appl. Environ. Microbiol. 1978, 35, 264–268. [Google Scholar] [CrossRef] [Green Version]
  64. Casman, E.P.; Bennett, R.W.; Dorsey, A.E.; Issa, J.A. Identification of a fourth staphylococcal enterotoxin, enterotoxin D. J. Bacteriol. 1967, 94, 1875–1882. [Google Scholar] [CrossRef] [Green Version]
  65. Bayles, K.W.; Iandolo, J.J. Genetic and molecular analyses of the gene encoding staphylococcal enterotoxin D. J. Bacteriol. 1989, 171, 4799–4806. [Google Scholar] [CrossRef] [Green Version]
  66. Chang, H.C.; Bergdoll, M.S. Purification and some physicochemical properties of staphylococcal enterotoxin D. Biochemistry 1979, 18, 1937–1942. [Google Scholar] [CrossRef] [PubMed]
  67. Bergdoll, M.S.; Borja, C.R.; Robbins, R.N.; Weiss, K.F. Identification of enterotoxin E. Infect. Immun. 1971, 4, 593–595. [Google Scholar] [CrossRef] [Green Version]
  68. Couch, J.L.; Soltis, M.T.; Betley, M.J. Cloning and nucleotide sequence of the type E staphylococcal enterotoxin gene. J. Bacteriol. 1988, 170, 2954–2960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Jarraud, S.; Peyrat, M.A.; Lim, A.; Tristan, A.; Bes, M.; Mougel, C.; Etienne, J.; Vandenesch, F.; Bonneville, M.; Lina, G. egc, a highly prevalent operon of enterotoxin gene, forms a putative nursery of superantigens in Staphylococcus aureus. J. Immunol. 2001, 166, 669–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Jarraud, S.; Peyrat, M.A.; Lim, A.; Tristan, A.; Bes, M.; Mougel, C.; Etienne, J.; Vandenesch, F.; Bonneville, M.; Lina, G. Correction. J. Immunol. 2001, 166, 4260. [Google Scholar]
  71. Sumby, P.; Waldor, M.K. Transcription of the toxin genes present within the staphylococcal phage φSa3ms is intimately linked with the phage’s life cycle. J. Bacteriol. 2003, 185, 6841–6851. [Google Scholar] [CrossRef] [Green Version]
  72. Abe, J.; Ito, Y.; Onimaru, M.; Kohsaka, T.; Takeda, T. Characterization and distribution of a new enterotoxin-related superantigen produced by Staphylococcus aureus. Microbiol. Immunol. 2000, 44, 79–88. [Google Scholar] [CrossRef]
  73. Noto, M.J.; Archer, G.L. A subset of Staphylococcus aureus strains harboring staphylococcal cassette chromosome mec (SCCmec) type IV is deficient in CcrAB-mediated SCCmec excision. Antimicrob. Agents Chemother. 2006, 50, 2782–2788. [Google Scholar] [CrossRef] [Green Version]
  74. Su, Y.C.; Wong, A.C. Identification and purification of a new staphylococcal enterotoxin, H. Appl. Environ. Microbiol. 1995, 61, 1438–1443. [Google Scholar] [CrossRef] [Green Version]
  75. Ren, K.; Bannan, J.D.; Pancholi, V.; Cheung, A.L.; Robbins, J.C.; Fischetti, V.A.; Zabriskie, J.B. Characterization and biological properties of a new staphylococcal exotoxin. J. Exp. Med. 1994, 180, 1675–1683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Munson, S.H.; Tremaine, M.T.; Betley, M.J.; Welch, R.A. Identification and characterization of staphylococcal enterotoxin types G and I from Staphylococcus aureus. Infect. Immun. 1998, 66, 3337–3348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Omoe, K.; Hu, D.L.; Takahashi-Omoe, H.; Nakane, A.; Shinagawa, K. Identification and characterization of a new staphylococcal enterotoxin-related putative toxin encoded by two kinds of plasmids. Infect. Immun. 2003, 71, 6088–6094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Zhang, S.; Iandolo, J.J.; Stewart, G.C. The enterotoxin D plasmid of Staphylococcus aureus encodes a second enterotoxin determinant (sej). FEMS Microbiol. Lett. 1998, 168, 227–233. [Google Scholar] [CrossRef] [Green Version]
  79. Orwin, P.M.; Leung, D.Y.M.; Donahue, H.L.; Novick, R.P.; Schlievert, P.M. Biochemical and biological properties of staphylococcal enterotoxin K. Infect. Immun. 2001, 69, 360–366. [Google Scholar] [CrossRef] [Green Version]
  80. Ono, H.K.; Hirose, S.; Naito, I.; Sato’o, Y.; Asano, K.; Hu, D.-L.; Omoe, K.; Nakane, A. The emetic activity of staphylococcal enterotoxins, SEK, SEL, SEM, SEN and SEO in a small emetic animal model, the house musk shrew. Microbiol. Immunol. 2017, 61, 12–16. [Google Scholar] [CrossRef] [Green Version]
  81. Orwin, P.M.; Fitzgerald, J.R.; Leung, D.Y.M.; Gutierrez, J.A.; Bohach, G.A.; Schlievert, P.M. Characterization of Staphylococcus aureus enterotoxin L. Infect. Immun. 2003, 71, 2916–2919. [Google Scholar] [CrossRef] [Green Version]
  82. Yarwood, J.M.; McCormick, J.K.; Paustian, M.L.; Orwin, P.M.; Kapur, V.; Schlievert, P.M. Characterization and expression analysis of Staphylococcus aureus pathogenicity island 3. Implications for the evolution of staphylococcal pathogenicity islands. J. Biol. Chem. 2002, 277, 13138–13147. [Google Scholar] [CrossRef] [Green Version]
  83. Omoe, K.; Imanishi, K.; Hu, D.-L.L.; Kato, H.; Fugane, Y.; Abe, Y.; Hamaoka, S.; Watanabe, Y.; Nakane, A.; Uchiyama, T.; et al. Characterization of novel staphylococcal enterotoxin-like toxin type P. Infect. Immun. 2005, 73, 5540–5546. [Google Scholar] [CrossRef] [Green Version]
  84. Hu, D.-L.; Ono, H.K.; Isayama, S.; Okada, R.; Okamura, M.; Lei, L.C.; Liu, Z.S.; Zhang, X.-C.; Liu, M.Y.; Cui, J.C.; et al. Biological characteristics of staphylococcal enterotoxin Q and its potential risk for food poisoning. J. Appl. Microbiol. 2017, 122, 1672–1679. [Google Scholar] [CrossRef]
  85. Orwin, P.M.; Leung, D.Y.; Tripp, T.J.; Bohach, G.A.; Earhart, C.A.; Ohlendorf, D.H.; Schlievert, P.M. Characterization of a novel staphylococcal enterotoxin-like superantigen, a member of the group V subfamily of pyrogenic toxins. Biochemistry 2002, 41, 14033–14040. [Google Scholar] [CrossRef] [PubMed]
  86. Ono, H.K.; Omoe, K.; Imanishi, K.; Iwakabe, Y.; Hu, D.L.; Kato, H.; Saito, N.; Nakane, A.; Uchiyama, T.; Shinagawa, K. Identification and characterization of two novel staphylococcal enterotoxins, types S and T. Infect. Immun. 2008, 76, 4999–5005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Letertre, C.; Perelle, S.; Dilasser, F.; Fach, P. Identification of a new putative enterotoxin SEU encoded by the egc cluster of Staphylococcus aureus. J. Appl. Microbiol. 2003, 95, 38–43. [Google Scholar] [CrossRef] [PubMed]
  88. Thomas, D.Y.; Jarraud, S.; Lemercier, B.; Cozon, G.; Echasserieau, K.; Etienne, J.; Gougeon, M.-L.; Lina, G.; Vandenesch, F. Staphylococcal enterotoxin-like toxins U2 and V, two new staphylococcal superantigens arising from recombination within the enterotoxin gene cluster. Infect. Immun. 2006, 74, 4724–4734. [Google Scholar] [CrossRef] [Green Version]
  89. Okumura, K.; Shimomura, Y.; Murayama, S.; Yagi, J.; Ubukata, K.; Kirikae, T.; Miyoshi-Akiyama, T. Evolutionary paths of streptococcal and staphylococcal superantigens. BMC Genom. 2012, 13, 404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Wilson, G.J.; Seo, K.S.; Cartwright, R.A.; Connelley, T.; Chuang-Smith, O.N.; Merriman, J.A.; Guinane, C.M.; Park, J.Y.; Bohach, G.A.; Schlievert, P.M.; et al. A novel core genome-encoded superantigen contributes to lethality of community-associated MRSA necrotizing pneumonia. PLoS Pathog. 2011, 7, e1002271. [Google Scholar] [CrossRef] [Green Version]
  91. Ono, H.K.; Sato’o, Y.; Narita, K.; Naito, I.; Hirose, S.; Hisatsune, J.; Asano, K.; Hu, D.L.; Omoe, K.; Sugai, M.; et al. Identification and characterization of a novel staphylococcal emetic toxin. Appl. Environ. Microbiol. 2015, 81, 7034–7040. [Google Scholar] [CrossRef] [Green Version]
  92. Wilson, G.J.; Tuffs, S.W.; Wee, B.A.; Seo, K.S.; Park, N.; Connelley, T.; Guinane, C.M.; Morrison, W.I.; Fitzgerald, J.R. Bovine Staphylococcus aureus superantigens stimulate the entire T cell repertoire of cattle. Infect. Immun. 2018, 86, 1–16. [Google Scholar] [CrossRef] [Green Version]
  93. Novick, R.P.; Ram, G. Staphylococcal pathogenicity islands—Movers and shakers in the genomic firmament. Curr. Opin. Microbiol. 2017, 38, 197–204. [Google Scholar] [CrossRef]
  94. Tuffs, S.W.; Haeryfar, S.M.M.; McCormick, J.K. Manipulation of innate and adaptive immunity by staphylococcal superantigens. Pathogens 2018, 7, 53. [Google Scholar] [CrossRef] [Green Version]
  95. Maina, E.K.; Hu, D.L.; Tsuji, T.; Omoe, K.; Nakane, A. Staphylococcal enterotoxin A has potent superantigenic and emetic activities but not diarrheagenic activity. Int. J. Med. Microbiol. 2012, 302, 88–95. [Google Scholar] [CrossRef] [PubMed]
  96. Suzuki, S. Studies on the relationship between gastric acidity and the development of MRSA. Especially for the prevention of MRSA enterocolitis. J. Nippon Med. Sch. 1994, 61, 563–571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Acton, D.S.; Tempelmans Plat-Sinnige, M.J.; Van Wamel, W.; De Groot, N.; Van Belkum, A. Intestinal carriage of Staphylococcus aureus: How does its frequency compare with that of nasal carriage and what is its clinical impact? Eur. J. Clin. Microbiol. Infect. Dis. 2009, 28, 115–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Jones, T.F.; Kellum, M.E.; Porter, S.S.; Bell, M.; Schaffner, W. An outbreak of community-acquired foodborne illness caused by methicillin-resistant Staphylococcus aureus. Emerg. Infect. Dis. 2002, 8, 82–84. [Google Scholar] [CrossRef]
  99. Veras, J.F.; do Carmo, L.S.; Tong, L.C.; Shupp, J.W.; Cummings, C.; dos Santos, D.A.; Cerqueira, M.M.O.P.; Cantini, A.; Nicoli, J.R.; Jett, M. A study of the enterotoxigenicity of coagulase-negative and coagulase-positive staphylococcal isolates from food poisoning outbreaks in Minas Gerais, Brazil. Int. J. Infect. Dis. 2008, 12, 410–415. [Google Scholar] [CrossRef] [Green Version]
  100. Wieneke, A.A.; Roberts, D.; Gilbert, R.J. Staphylococcal food poisoning in the United Kingdom, 1969–1990. Epidemiol. Infect. 1993, 110, 519–531. [Google Scholar] [CrossRef] [Green Version]
  101. Schelin, J.; Wallin-Carlquist, N.; Cohn, M.T.; Lindqvist, R.; Barker, G.C.; Rådström, P. The formation of Staphylococcus aureus enterotoxin in food environments and advances in risk assessment. Virulence 2011, 2, 580–592. [Google Scholar] [CrossRef] [Green Version]
  102. Hu, D.L.; Nakane, A. Mechanisms of staphylococcal enterotoxin-induced emesis. Eur. J. Pharmacol. 2014, 722, 95–107. [Google Scholar] [CrossRef]
  103. Edwards, L.A.; O’Neill, C.; Furman, M.A.; Hicks, S.; Torrente, F.; Pérez-Machado, M.; Wellington, E.M.; Phillips, A.D.; Murch, S.H. Enterotoxin-producing staphylococci cause intestinal inflammation by a combination of direct epithelial cytopathy and superantigen-mediated T-cell activation. Inflamm. Bowel Dis. 2012, 18, 624–640. [Google Scholar] [CrossRef]
  104. Ono, H.K.; Nishizawa, M.; Yamamoto, Y.; Hu, D.L.; Nakane, A.; Shinagawa, K.; Omoe, K. Submucosal mast cells in the gastrointestinal tract are a target of staphylococcal enterotoxin type A. FEMS Immunol. Med. Microbiol. 2012, 64, 392–402. [Google Scholar] [CrossRef] [Green Version]
  105. Ono, H.K.; Hirose, S.; Narita, K.; Sugiyama, M.; Asano, K.; Hu, D.L.; Nakane, A. Histamine release from intestinal mast cells induced by staphylococcal enterotoxin a (SEA) evokes vomiting reflex in common marmoset. PLoS Pathog. 2019, 15, 1–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wells, U.; Ravenscroft, M.; Bhandari, P.; Andrews, P.L.R. Serotonin and serotonergic drugs in emesis. In Serotonin: From Cell Biology to Pharmacology and Therapeutics; Vanhoutte, P.M., Saxena, P.R., Paoletti, R., Brunello, N., Jackson, A.S., Eds.; Springer: Dordrecht, The Netherlands, 1993; pp. 179–186. ISBN 978-94-011-1920-7. [Google Scholar]
  107. Sugiyama, H.; Hayama, T. Abdominal viscera as site of emetic action for staphylococcal enterotoxin in the monkey. J. Infect. Dis. 1965, 115, 330–336. [Google Scholar] [CrossRef] [PubMed]
  108. Schlievert, P.M.; Jablonski, L.M.; Roggiani, M.; Sadler, I.; Callantine, S.; Mitchell, D.T.; Ohlendorf, D.H.; Bohach, G.A. Pyrogenic toxin superantigen site specificity in toxic shock syndrome and food poisoning in animals. Infect. Immun. 2000, 68, 3630–3634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Aragon-Alegro, L.C.; Konta, E.M.; Suzuki, K.; Silva, M.G.; Júnior, A.F.; Rall, R.; Rall, V.L.M. Occurrence of coagulase-positive Staphylococcus in various food products commercialized in Botucatu, SP, Brazil and detection of toxins from food and isolated strains. Food Control 2007, 18, 630–634. [Google Scholar] [CrossRef]
  110. Lindqvist, R.; Sylve, S. Quantitative microbial risk assessment exemplified by Staphylococcus aureus in unripened cheese made from raw milk. Int. J. Food Microbiol. 2002, 78, 155–170. [Google Scholar] [CrossRef] [Green Version]
  111. Scherrer, D.; Corti, S.; Muehlherr, J.E.; Zweifel, C.; Stephan, R. Phenotypic and genotypic characteristics of Staphylococcus aureus isolates from raw bulk-tank milk samples of goats and sheep. Vet. Microbiol. 2004, 101, 101–107. [Google Scholar] [CrossRef]
  112. Rosengren, Å.; Fabricius, A.; Guss, B.; Sylvén, S.; Lindqvist, R. Occurrence of foodborne pathogens and characterization of Staphylococcus aureus in cheese produced on farm-dairies. Int. J. Food Microbiol. 2010, 144, 263–269. [Google Scholar] [CrossRef]
  113. Artursson, K.; Schelin, J.; Thisted Lambertz, S.; Hansson, I.; Olsson Engvall, E. Foodborne pathogens in unpasteurized milk in Sweden. Int. J. Food Microbiol. 2018, 284, 120–127. [Google Scholar] [CrossRef]
  114. Valihrach, L.; Alibayov, B.; Demnerova, K. Production of staphylococcal enterotoxin C in milk. Int. Dairy J. 2013, 30, 103–107. [Google Scholar] [CrossRef]
  115. Necidova, L.; Bogdanovicova, K.; Harustiakova, D.; Bartova, K. Short communication: Pasteurization as a means of inactivating staphylococcal enterotoxins A, B, and C in milk. J. Dairy Sci. 2016, 99, 8638–8643. [Google Scholar] [CrossRef] [Green Version]
  116. Necidová, L.; Bursová, Š.; Haruštiaková, D.; Bogdanovičová, K.; Lačanin, I. Effect of heat treatment on activity of staphylococcal enterotoxins of type A, B, and C in milk. J. Dairy Sci. 2019, 102, 3924–3932. [Google Scholar] [CrossRef] [PubMed]
  117. Hunt, K.; Butler, F.; Jordan, K. Modelling production of S. aureus enterotoxin Cbovine in milk, and its production during cheesemaking. Dairy Sci. Technol. 2015, 95, 747–757. [Google Scholar] [CrossRef]
  118. Silva, G.O.; Castro, R.D.; Oliveira, L.G.; Sant’Anna, F.M.; Barbosa, C.D.; Sandes, S.H.C.; Silva, R.S.; Resende, M.F.S.; Lana, A.M.Q.; Nunes, A.C.; et al. Viability of Staphylococcus aureus and expression of its toxins (SEC and TSST-1) in cheeses using Lactobacillus rhamnosus D1 or Weissella paramesenteroides GIR16L4 or both as starter cultures. J. Dairy Sci. 2020, 103, 4100–4108. [Google Scholar] [CrossRef] [PubMed]
  119. Johler, S.; Zurfluh, K.; Stephan, R. Tracing and inhibiting growth of Staphylococcus aureus in barbecue cheese production after product recall. J. Dairy Sci. 2016, 99, 3345–3350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Kadariya, J.; Smith, T.C.; Thapaliya, D. Staphylococcus aureus and staphylococcal food-borne disease: An ongoing challenge in public health. Biomed Res. Int. 2014, 2014, 827965. [Google Scholar] [CrossRef] [Green Version]
  121. Jie, K.G.; Jiang, H.; Sun, L.; Wang, H.R.; Zheng, Y.L.; Li, Y.; Jiang, Y.Q. The pilot study of anti-tumor effects versus immunosuppression of staphylococcal enterotoxin C. Cancer Biol. Ther. 2007, 6, 1584–1591. [Google Scholar] [CrossRef] [Green Version]
  122. Cho, S.; Swaminathan, C.P.; Yang, J.; Kerzic, M.C.; Guan, R.; Kieke, M.C.; Kranz, D.M.; Mariuzza, R.A.; Sundberg, E.J. Structural basis of affinity maturation and intramolecular cooperativity in a protein-protein interaction. Structure 2005, 13, 1775–1787. [Google Scholar] [CrossRef] [Green Version]
  123. Deringer, J.R.; Ely, R.J.; Stauffacher, C.V.; Bohach, G.A. Subtype-specific interactions of type C staphylococcal enterotoxins with the T-cell receptor. Mol. Microbiol. 1996, 22, 523–534. [Google Scholar] [CrossRef]
  124. Langley, R.J.; Fraser, J.D.; Proft, T. Bacterial Superantigens and Superantigen-Like Toxins; Elsevier Ltd.: Amsterdam, The Netherlands, 2015; ISBN 9780128005897. [Google Scholar]
  125. Fields, B.A.; Malchiodi, E.L.; Li, H.; Ysern, X.; Stauffacher, C.V.; Schlievert, P.M.; Karjalainen, K.; Mariuzza, R.A. Crystal structure of a T-cell receptor β-chain complexed with a superantigen. Nature 1996, 384, 188–192. [Google Scholar] [CrossRef]
  126. Deringer, J.R.; Ely, R.J.; Monday, S.R.; Stauffacher, C.V.; Bohach, G.A. Vβ-dependent stimulation of bovine and human T cells by host-specific staphylococcal enterotoxins. Infect. Immun. 1997, 65, 4048–4054. [Google Scholar] [CrossRef] [Green Version]
  127. Lamphear, J.G.; Bohach, G.A.; Rich, R.R. Structural dichotomy of staphylococcal enterotoxin C superantigens leading to MHC class II-independent activation of T lymphocytes. J. Immunol. 1998, 160, 2107–2114. [Google Scholar]
  128. Kaempfer, R.; Arad, G.; Levy, R.; Hillman, D.; Nasie, I.; Rotfogel, Z. CD28: Direct and critical receptor for superantigen toxins. Toxins 2013, 5, 1531–1542. [Google Scholar] [CrossRef] [PubMed]
  129. Arad, G.; Levy, R.; Nasie, I.; Hillman, D.; Rotfogel, Z.; Barash, U.; Supper, E.; Shpilka, T.; Minis, A.; Kaempfer, R. Binding of superantigen toxins into the CD28 homodimer interface is essential for induction of cytokine genes that mediate lethal shock. PLoS Biol. 2011, 9, e1001149. [Google Scholar] [CrossRef] [PubMed]
  130. Shands, K.N.; Schmid, G.P.; Dan, B.B.; Blum, D.; Guidotti, R.J.; Hargrett, N.T.; Anderson, R.L.; Hill, D.L.; Broome, C.V.; Band, J.D.; et al. Toxic-Shock Syndrome in menstruating women. N. Engl. J. Med. 1980, 303, 1436–1442. [Google Scholar] [CrossRef] [PubMed]
  131. Rizkallah, M.F.; Tolaymat, A.; Martinez, J.S.; Schlievert, P.M.; Ayoub, E.M. Toxic shock syndrome caused by a strain of Staphylococcus aureus that produces enterotoxin C but not toxic shock syndrome toxin-1. Am. J. Dis. Child. 1989, 143, 848–849. [Google Scholar] [CrossRef]
  132. Bartlett, A.H.; Hulten, K.G. Staphylococcus aureus pathogenesis: Secretion systems, adhesins, and invasins. Pediatr. Infect. Dis. J. 2010, 29, 860–861. [Google Scholar] [CrossRef]
  133. Christianson, K.K.; Tweten, R.K.; Iandolo, J.J. Transport and processing of staphylococcal enterotoxin A. Appl. Environ. Microbiol. 1985, 50, 696–697. [Google Scholar] [CrossRef] [Green Version]
  134. Tweten, R.K.; Iandolo, J.J. Purification and partial characterization of a putative precursor to staphylococcal enterotoxin B. Infect. Immun. 1981, 34, 900–907. [Google Scholar] [CrossRef] [Green Version]
  135. Tweten, R.K.; Iandolo, J.J. Transport and processing of staphylococcal enterotoxin B. J. Bacteriol. 1983, 153, 297–303. [Google Scholar] [CrossRef] [Green Version]
  136. Papageorgiou, A.C.; Acharya, K.R.; Shapiro, R.; Passalacqua, E.F.; Brehm, R.D.; Tranter, H.S. Crystal structure of the superantigen enterotoxin C2 from Staphylococcus aureus reveals a zinc-binding site. Structure 1995, 3, 769–779. [Google Scholar] [CrossRef] [Green Version]
  137. Baker, M.; Gutman, D.M.; Papageorgiou, A.C.; Collins, C.M.; Acharya, K.R. Structural features of a zinc binding site in the superantigen strepococcal pyrogenic exotoxin A (SpeA1): Implications for MHC class II recognition. Protein Sci. 2001, 10, 1268–1273. [Google Scholar] [CrossRef] [Green Version]
  138. Papageorgiou, A.C.; Baker, M.D.; McLeod, J.D.; Goda, S.K.; Manzotti, C.N.; Sansom, D.M.; Tranter, H.S.; Acharya, K.R. Identification of a secondary zinc-binding site in staphylococcal enterotoxin C2. J. Biol. Chem. 2004, 279, 1297–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Chi, Y.-I.; Sadler, I.; Jablonski, L.M.; Callantine, S.D.; Deobald, C.F.; Stauffacher, C.V.; Bohach, G.A. Zinc-mediated dimerization and its effect on activity and conformation of staphylococcal enterotoxin type C. J. Biol. Chem. 2002, 277, 22839–22846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Herrmann, T.; Accolla, R.S.; MacDonald, H.R. Different staphylococcal enterotoxins bind preferentially to distinct major histocompatibility complex class II isotypes. Eur. J. Immunol. 1989, 19, 2171–2174. [Google Scholar] [CrossRef] [PubMed]
  141. Jardetzky, T.S.; Brown, J.H.; Gorga, J.C.; Stern, L.J.; Urban, R.G.; Chi, Y.; Stauffacher, C.; Strominger, J.L.; Wiley, D.C. Three-dimensional structure of a human class II histocompatibility molecule complexed with superantigen. Nature 1994, 368, 711–718. [Google Scholar] [CrossRef]
  142. Thomas, D.; Chou, S.; Dauwalder, O.; Lina, G. Diversity in Staphylococcus aureus enterotoxins. In Superantigens and Superallergens; Karger: Basel, Switzerland, 2007; Volume 93, pp. 24–41. [Google Scholar]
  143. Hovde, C.J.; Marr, J.C.; Hoffmann, M.L.; Hackett, S.P.; Chi, Y.Y.-I.; Crum, K.K.; Stevens, D.L.; Stauffacher, C.V.; Bohach, G.A. Investigation of the role of the disulphide bond in the activity and structure of staphylococcal enterotoxin C1. Mol. Microbiol. 1994, 13, 897–909. [Google Scholar] [CrossRef]
  144. Hoffman, M.; Tremaine, M.; Mansfield, J.; Betley, M. Biochemical and mutational analysis of the histidine residues of staphylococcal enterotoxin A. Infect. Immun. 1996, 64, 885–890. [Google Scholar] [CrossRef] [Green Version]
  145. Harris, T.O.; Grossman, D.; Kappler, J.W.; Marrack, P.; Rich, R.R.; Betley, M.J. Lack of complete correlation between emetic and T-cell-stimulatory activities of staphylococcal enterotoxins. Infect. Immun. 1993, 61, 3175–3183. [Google Scholar] [CrossRef] [Green Version]
  146. Maina, E.K.; Hu, D.L.; Asano, K.; Nakane, A. Inhibition of emetic and superantigenic activities of staphylococcal enterotoxin A by synthetic peptides. Peptides 2012, 38, 1–7. [Google Scholar] [CrossRef]
  147. Otero, A.; Garcia, M.L.C.; Garcia, M.L.C.; Moreno, B.; Bergdoll, M.S. Production of staphylococcal enterotoxins C1 and C2 and thermonuclease throughout the growth cycle. Appl. Environ. Microbiol. 1990, 56, 555–559. [Google Scholar] [CrossRef] [Green Version]
  148. Regassa, L.B.; Betley, M.J. High sodium chloride concentrations inhibit staphylococcal enterotoxin C gene (sec) expression at the level of sec mRNA. Infect. Immun. 1993, 61, 1581–1585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Regassa, L.B.; Couch, J.L.; Betley, M.J. Steady-state staphylococcal enterotoxin type C mRNA is affected by a product of the accessory gene regulator (agr) and by glucose. Infect. Immun. 1991, 59, 955–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Tseng, C.W.; Zhang, S.; Stewart, G.C. Accessory gene regulator control of staphyloccoccal enterotoxin D gene expression. J. Bacteriol. 2004, 186, 1793–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Hsieh, H.Y.; Ching, W.T.; Stewart, G.C. Regulation of rot expression in Staphylococcus aureus. J. Bacteriol. 2008, 190, 546–554. [Google Scholar] [CrossRef] [Green Version]
  152. Novick, R.P.; Projan, S.J.; Kornblum, J.; Ross, H.F.; Ji, G.; Kreiswirth, B.; Vandenesch, F.; Moghazeh, S.; Novick, R.P. The agr P2 operon: An autocatalytic sensory transduction system in Staphylococcus aureus. MGG Mol. Gen. Genet. 1995, 248, 446–458. [Google Scholar] [CrossRef]
  153. Kusch, K.; Hanke, K.; Holtfreter, S.; Schmudde, M.; Kohler, C.; Erck, C.; Wehland, J.; Hecker, M.; Ohlsen, K.; Bröker, B.; et al. The influence of SaeRS and σB on the expression of superantigens in different Staphylococcus aureus isolates. Int. J. Med. Microbiol. 2011, 301, 488–499. [Google Scholar] [CrossRef]
  154. Liu, Q.; Yeo, W.S.; Bae, T. The SaeRS two-component system of Staphylococcus aureus. Genes 2016, 7, 81. [Google Scholar] [CrossRef] [Green Version]
  155. Genigeorgis, C.; Foda, M.S.; Mantis, A.; Sadler, W.W. Effect of sodium chloride and pH on enterotoxin C production. Appl. Microbiol. 1971, 21, 862–866. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) Similarity tree based on staphylococcal enterotoxins (SE) sequences including signal peptides—close-up of SEC sequences. The tree was constructed with CLC genomics workbench 12. Colors indicate phylogenetic pyrogenic toxin superantigen (Sag) groups according to Wilson et al., 2018 [92]. * SEH has been assigned to group III or separately in group IV. (B) Similarities between SEC variants including signal peptides. The upper right half shows similarity in %. The lower left half shows amino acid differences in absolute numbers. Colors indicate sequence similarity with red = low and blue = high. a Protein sequences are available under the accession numbers indicated in Supplementary Table S1. (C) Ribbon diagram of SEC2. The N terminal domain is colored in light blue, the C-terminal domain in orange. The disulphide bridge and loop are colored in yellow. Binding sites are labelled. The two zinc ions are indicated with black spheres. α-helices and β-sheets are numbered from N to C terminus. PDB file accession number is indicated in Supplementary Table S1.
Figure 1. (A) Similarity tree based on staphylococcal enterotoxins (SE) sequences including signal peptides—close-up of SEC sequences. The tree was constructed with CLC genomics workbench 12. Colors indicate phylogenetic pyrogenic toxin superantigen (Sag) groups according to Wilson et al., 2018 [92]. * SEH has been assigned to group III or separately in group IV. (B) Similarities between SEC variants including signal peptides. The upper right half shows similarity in %. The lower left half shows amino acid differences in absolute numbers. Colors indicate sequence similarity with red = low and blue = high. a Protein sequences are available under the accession numbers indicated in Supplementary Table S1. (C) Ribbon diagram of SEC2. The N terminal domain is colored in light blue, the C-terminal domain in orange. The disulphide bridge and loop are colored in yellow. Binding sites are labelled. The two zinc ions are indicated with black spheres. α-helices and β-sheets are numbered from N to C terminus. PDB file accession number is indicated in Supplementary Table S1.
Toxins 12 00584 g001
Figure 2. Mechanism of SE emetic activity as proposed by Fisher et al., 2018 [45]. Enterotoxins enter through epithelial or mucus producing goblet cells. The activation of mast cells leads to release of serotonin. Serotonin stimulation of the vagus nerve provokes an emetic response. T-cells and neutrophils are activated as well but their role remains unclear. The figure was adapted from the original publication. Schematic illustrations were created using Servier medical art: https://smart.servier.com.
Figure 2. Mechanism of SE emetic activity as proposed by Fisher et al., 2018 [45]. Enterotoxins enter through epithelial or mucus producing goblet cells. The activation of mast cells leads to release of serotonin. Serotonin stimulation of the vagus nerve provokes an emetic response. T-cells and neutrophils are activated as well but their role remains unclear. The figure was adapted from the original publication. Schematic illustrations were created using Servier medical art: https://smart.servier.com.
Toxins 12 00584 g002
Figure 3. (A) Schematic representation of superantigenic activity of SEC. Schematic illustrations are from Servier medical art: https://smart.servier.com. (B) Model of SEC3 bound to MHC II α1 and TCR Vβ. SEC3 is colored in green, TCR in blue, MHC II in cyan, and the normal antigen in red. The model was composed with pyMOL v2.4.0 from two separate models of SEC3 complexed to MHC II, and SEC3 complexed to TCR. PDB accession numbers can be found in Supplementary Table S1.
Figure 3. (A) Schematic representation of superantigenic activity of SEC. Schematic illustrations are from Servier medical art: https://smart.servier.com. (B) Model of SEC3 bound to MHC II α1 and TCR Vβ. SEC3 is colored in green, TCR in blue, MHC II in cyan, and the normal antigen in red. The model was composed with pyMOL v2.4.0 from two separate models of SEC3 complexed to MHC II, and SEC3 complexed to TCR. PDB accession numbers can be found in Supplementary Table S1.
Toxins 12 00584 g003
Figure 4. Regulatory pathways involved in sec transcription. The quorum-sensing Agr system acts on sec transcription indirectly. AgrA induces RNAIII which represses the repressor of toxins (Rot), consequently allowing transcription of the sec gene. SarA, σB, and SaeRS might play an additional role in transcriptional regulation when environmental stress signals act on the cells.
Figure 4. Regulatory pathways involved in sec transcription. The quorum-sensing Agr system acts on sec transcription indirectly. AgrA induces RNAIII which represses the repressor of toxins (Rot), consequently allowing transcription of the sec gene. SarA, σB, and SaeRS might play an additional role in transcriptional regulation when environmental stress signals act on the cells.
Toxins 12 00584 g004
Table 1. Staphylococcal enterotoxin gene location and emetic activity. Adapted from [4,12,44,45]. nd = not determined.
Table 1. Staphylococcal enterotoxin gene location and emetic activity. Adapted from [4,12,44,45]. nd = not determined.
Enterotoxin/Enterotoxin-Like SAgEmetic
Activity
Associated Genetic ElementReferences
SEA+Prophage (φSa3ms, φSa3mw, φ252B, φNM3, φMu50a)[19,46,47,48,49]
SEB+SaPIs (SaPI1, SaPI2, SaPI3, SaPI4, SaPImw2, SaPIrki4)
Plasmid (pZA10)
[50,51,52,53,54]
SEC+SaPIs
Plasmid
[55,56,57]
SEC1+SaPINuSAα2 a, pZA10[53]
SEC2+SaPITokyo a[58]
SEC3+SaPIn1/SaPIm1 b[59,60]
SEC4ndSaPImw2[61]
SECbovinendSaPIbov1[30,62]
SECovinendSaPIbov5 a, SaPIov1[30,62,63]
SED+Plasmid (pIB485-like)[64,65,66]
SEE+Prophage(hypothetical)[67,68]
SEG+egc (egc1–4)
Prophage (φSa3ms)
[61,69,70,71,72]
SEH+Transposon (MGEwm2/mssa476 seh/Δseo)[61,73,74,75]
SEI+egc (egc1–3)[69,70,76]
SElJndPlasmid (pIB485-like, pF5)[27,77,78]
SEK+SaPIs (SaPIbov1, SaPI1, SaPI3, SaPI5)
Prophage (φSa3ms, φSa3mw)
[24,61,71,79,80]
SEL+SaPIs (SaPIbov1, SaPI3, SaPIn1, SaPIm1, SaPImw2)[24,57,80,81,82]
SEM+egc (egc1–2)[24,69,70,80]
SEN+egc (egc1–4)[24,69,70,80]
SEO+egc (egc1–4)
Transposon (MGEwm2/mssa476 sehseo)
[24,69,70,80]
SEP+Prophage (φSa3n, φN315, φMu3A)[24,83]
SEQ+SaPIs (SaPI1, SaPI3, SaPI5)
Prophage (φSa3ms, φSa3mw)
[24,84,85]
SER+Plasmid (pIB485-like, pF5)[77,86]
SES+Plasmid (pF5)[86]
SET+Plasmid (pF5)[86]
SElUndegc (egc2–3)[87,88]
SElU2 cndegc (egc4)[88]
SElVndegc (egc4)[88]
SElWndChromosome[89]
SElXndChromosome[90]
SElYndChromosome[91]
SElZndChromosome[92]
aS. aureus pathogenicity islands (SaPIs) were determined via NCBI nucleotide BLAST, accession numbers can be found in Table S1. b SaPIm1 (MU50) and SaPIn1 (N315) are identical [93]. c SElU2 was suggested to be renamed into SElW but since sequence similarity with SElU is 94% we recommend using SElU2 [94], this also facilitates discrimination from chromosomal SElW [89].

Share and Cite

MDPI and ACS Style

Etter, D.; Schelin, J.; Schuppler, M.; Johler, S. Staphylococcal Enterotoxin C—An Update on SEC Variants, Their Structure and Properties, and Their Role in Foodborne Intoxications. Toxins 2020, 12, 584. https://doi.org/10.3390/toxins12090584

AMA Style

Etter D, Schelin J, Schuppler M, Johler S. Staphylococcal Enterotoxin C—An Update on SEC Variants, Their Structure and Properties, and Their Role in Foodborne Intoxications. Toxins. 2020; 12(9):584. https://doi.org/10.3390/toxins12090584

Chicago/Turabian Style

Etter, Danai, Jenny Schelin, Markus Schuppler, and Sophia Johler. 2020. "Staphylococcal Enterotoxin C—An Update on SEC Variants, Their Structure and Properties, and Their Role in Foodborne Intoxications" Toxins 12, no. 9: 584. https://doi.org/10.3390/toxins12090584

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop