Next Article in Journal
Diffuse Myocardial Fibrosis on Cardiac Magnetic Resonance Imaging Is Related to Galectin-3 and Predicts Outcome in Heart Failure
Next Article in Special Issue
Intestinal Barrier in Post-Campylobacter jejuni Irritable Bowel Syndrome
Previous Article in Journal
Editorial: Special Issue “Protein Modeling and Simulation: Selected Articles from the Computational Structural Bioinformatics Workshop 2021”
Previous Article in Special Issue
Evolution and Role of Proteases in Campylobacter jejuni Lifestyle and Pathogenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Targets in Campylobacter Infections

1
Gastrointestinal Microbiology Research Group, Institute of Microbiology, Infectious Diseases and Immunology, Charité—University Medicine Berlin, Corporate Member of Free University Berlin, Humboldt University Berlin, and Berlin Institute of Health, D-12203 Berlin, Germany
2
Division of Microbiology, Department of Biology, Friedrich-Alexander-Universität Erlangen-Nürnberg, D-91058 Erlangen, Germany
3
Institute of Food Safety and Food Hygiene, School of Veterinary Medicine, Free University Berlin, D-14163 Berlin, Germany
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(3), 409; https://doi.org/10.3390/biom13030409
Submission received: 14 December 2022 / Revised: 8 February 2023 / Accepted: 13 February 2023 / Published: 22 February 2023
(This article belongs to the Special Issue Molecular Targets in Campylobacter Infections)

Abstract

:
Human campylobacteriosis results from foodborne infections with Campylobacter bacteria such as Campylobacter jejuni and Campylobacter coli, and represents a leading cause of bacterial gastroenteritis worldwide. After consumption of contaminated poultry meat, constituting the major source of pathogenic transfer to humans, infected patients develop abdominal pain and diarrhea. Post-infectious disorders following acute enteritis may occur and affect the nervous system, the joints or the intestines. Immunocompromising comorbidities in infected patients favor bacteremia, leading to vascular inflammation and septicemia. Prevention of human infection is achieved by hygiene measures focusing on the reduction of pathogenic food contamination. Molecular targets for the treatment and prevention of campylobacteriosis include bacterial pathogenicity and virulence factors involved in motility, adhesion, invasion, oxygen detoxification, acid resistance and biofilm formation. This repertoire of intervention measures has recently been completed by drugs dampening the pro-inflammatory immune responses induced by the Campylobacter endotoxin lipo-oligosaccharide. Novel pharmaceutical strategies will combine anti-pathogenic and anti-inflammatory effects to reduce the risk of both anti-microbial resistance and post-infectious sequelae of acute enteritis. Novel strategies and actual trends in the combat of Campylobacter infections are presented in this review, alongside molecular targets applied for prevention and treatment strategies.

1. Introduction

Campylobacteriosis is the leading cause of bacterial gastroenteritis worldwide [1,2,3,4,5]. Human campylobacteriosis is mostly caused by Campylobacter jejuni and less frequently by Campylobacter coli, and occurs predominantly after consumption of contaminated chicken meat [6]. According to a recent EFSA report, a total of 31% of chilled broiler carcasses sampled at slaughterhouses were Campylobacter-positive. At the food level, the highest Campylobacter proportions were detected in fresh meat from turkeys (12.9%), followed by fresh meat from broilers (11.5%) [7]. The potent pathogens cause acute and severe enteritis in humans, but are harmless commensals in domestic animals. The asymptomatic colonization in chicken flocks supports the formation of huge pathogen reservoirs [8,9]. Thus, reduction of Campylobacter contamination in chicken breeding and the associated food chain by hygiene interventions is key to the prevention of human campylobacteriosis. Humans develop enteritis even after ingestion of low pathogen doses in the range of hundreds of living bacteria. Campylobacteriosis symptoms vary from mild abdominal pain to severe, inflammatory, bloody diarrhea associated with grievous cramps and fever, which can last for more than a week (as reviewed elsewhere [10,11,12,13,14]). The severity of initial enteritis is significantly associated with the risk of post-infectious sequelae, which may appear weeks or months after the intestinal infection [15]. Post-infectious sequelae include Guillain–Barré-Syndrome (GBS) [16,17], reactive arthritis [18], inflammatory bowel diseases [19,20,21] or irritable bowel syndrome [18]. In rare cases, Campylobacter bacteremia occurring in immunocompromised individuals may lead to meningitis [22] or cardiovascular diseases [23] (for reviews concerning post-infectious complications, see [4,13,24,25,26]). Most importantly, the risk for post-infectious sequelae increases with the severity of the initial enteritis [14]. This supports the need for novel prevention and treatment strategies, given that most patients receive rehydration measures, but no causative treatment of the inflammation [14,17,27]. In consequence of the socioeconomic burden caused by campylobacteriosis, industrialized countries have established control measures by reporting and undertaking statistical analysis of Campylobacter-associated diseases. However, unreported campylobacteriosis cases are expected to exceed by far the number of cases reported in published statistics [28,29,30].

2. From Microbiology and Intestinal Inflammation to Molecular Targeting of Campylobacteriosis

The multifaceted campylobacteriosis symptoms and the severe post-infectious sequelae have stimulated a wealth of research into the pathogenesis of C. jejuni infections. While investigations on the inflammation have long been hampered by the lack of experimental murine infection models [31], there is remarkable progress in the discovery of bacterial pathogenicity and virulence factors. Several microbiological characteristics represent valuable molecular targets for intervention strategies to prevent and treat C. jejuni infections.
Campylobacters belong to the ε-proteobacteria subphylum of Gram-negative bacteria. The spiral C. jejuni bacteria are thermophilic with a growth optimum of 42 °C. The chromosome carries around 1600 genes at high density [32,33,34]. The low gene number restricts the metabolic capabilities of C. jejuni and supports a fastidious and microaerophilic lifestyle with a limited repertoire for environmental adaptation [32,33,34]. Many regulatory systems and carbohydrate utilization pathways present in other Gram-negative pathogens are absent in C. jejuni, as discussed below [30]. This limited adaptation capacity results in sensitivity to environmental stress, favoring the sensitivity of C. jejuni to organic acids as a means of preventive hygiene measures in the poultry industries [35]. In humans, the use of proton pump inhibitors increases the risk of C. jejuni infection, indicating that gastric acid protects from campylobacteriosis [36]. Therefore, maintaining stomach health is a major goal for the prevention of campylobacteriosis and enteric infections in general.
In human hosts, ingested C. jejuni bacteria surviving gastric acid, bile acids and digestive enzymes in the stomach and upper intestines enter the intestinal mucus layer, which is facilitated by their motility. After adhesion to epithelial cells and invasion of the sub-epithelial tissues, the pathogens induce inflammation by activation of the innate immune system. Thus, in the human host, adaptation to acid, bile and oxygen, as well as the structure and functions of flagella, adhesion and invasion proteins, represent major pathogenicity determinants of C. jejuni that are essential for the onset, progression and clinical outcome of campylobacteriosis [37,38,39,40]. In line with the mechanisms mediating bacterial survival in the environment, these factors serve as valid molecular targets for prevention and treatment strategies, as is summarized in the following paragraphs and in Figure 1. In contrast, the use of inflammation as a molecular target for the amelioration of campylobacteriosis awaits further investigation, and corresponding developments are described in the last paragraph of this review.

3. Pathogenesis of Campylobacter Infection and Introduction to the Major Molecular Targets within C. jejuni

Our knowledge of mechanisms involved in the molecular pathogenesis of C. jejuni is based on clinical studies, investigations of the bacteria in cell culture in vitro and in animal models such as chicken and mice [5,41]. After ingestion, the intruding bacteria move very effectively by using their rotating bipolar flagella. C. jejuni controls its motility by a multifaceted chemotaxis machinery that is guided through various chemoreceptors. These chemotaxis signaling pathways allow the pathogen to move toward beneficial chemoattractants and repulse from chemorepellents. The best-characterized C. jejuni chemotaxis cascade comprises the CheA/CheW-CheY proteins, all of which are essential for successful infection of human and animal hosts [42,43,44]. This machinery is used by C. jejuni to cross the mucous layer of the lower gastrointestinal tract, a nutrient-rich niche that is perfectly suited for bacterial colonization and growth [45]. The catabolic versatility of C. jejuni is limited compared to other enterobacteria, as various pathways of the carbohydrate metabolism are absent. However, C. jejuni exhibits numerous chemotactic activities, most notably the sensing of amino acids such as asparagine and aspartate, lactate, formate and intermediates of the citric acid cycle, which are the key energy sources for the bacterium [46,47]. Thus, the metabolism of C. jejuni is highly specialized, a feature that can be used to target this pathogen. A screen of an inhibitor library comprising about 147,000 compounds revealed a number of small molecules that profoundly compromised the in vitro growth of C. jejuni and flagellar gene expression, respectively [48]. Some of these inhibitors were found to display bacteriostatic effects on C. jejuni, while not being harmful to host cells. Oral administration of at least one compound resulted in reduced C. jejuni loads in a chicken model [48]. In addition, several anti-microbial compounds were found to change the morphology of C. jejuni. For instance, meropenem triggered bulking of the bacteria, imipenem induced the conversion of spiral-shaped to coccoid forms, and sitafloxacin resulted in bacterial elongation, all of which were associated with severely compromised C. jejuni motility [49]. These studies are very promising with regard to pinpointing anti-C. jejuni activities, but further investigations are clearly required to validate the specific cellular targets of the compounds, and their possible application in human therapy.
C. jejuni colonizes the intestinal mucus layer as an initial step of the infection [50]. Subsequently, the pathogen adheres to intestinal epithelial cells, transmigrates across the intestinal epithelium and initiates host cell entry and intracellular survival. Interestingly, the addition of purified intestinal chicken mucus, but not mucus of human origin, inhibited both C. jejuni binding and entry of human epithelial cells [51]. Bacterial invasion of the host cells was more strongly repressed than C. jejuni adhesion. Pre-treatment of chicken mucin with sodium-metaperiodate, a chemical compound that oxidizes glycan entities on mucin glycoproteins, restored C. jejuni invasion of human cells [51]. In contrast, exposure of chicken mucin to other enzymes, specifically sialidase or fucosidase, did not abolish C. jejuni cell binding and entry. These observations strongly suggest that glycosylation of chicken mucin mediates the observed suppressive effects on C. jejuni host cell interactions. However, the exact interaction points and specific glycan residues in the mucin are still unknown and should be explored in future studies.
Besides C. jejuni colonization of the mucus layer, the bacteria can also bind to enterocytes, which is facilitated by an array of surface-exposed adhesion proteins. Among several candidates, the two fibronectin-binding proteins CadF and FlpA have been studied in detail [52,53]. These proteins also mediate the invasion of C. jejuni into the intestinal epithelium from the bottom of the cells via a basal mechanism using fibronectin and integrin receptors. C. jejuni invasion is enhanced by so-called Cia (Campylobacter invasion antigens) proteins such as CiaB, CiaC, CiaD and others that are secreted by the flagellar type III secretion system (fT3SS) [54,55,56]. These interactions also trigger the activation of the small Rho GTPases Rac1 and Cdc42, which induce cytoskeletal reorganizations and bacterial uptake [52,53,57,58,59,60]. The ability to survive intracellularly in vacuoles enhances the pathogenicity of C. jejuni. Various natural compounds from plant extracts with so-called GRAS (generally recognized as safe) status have been screened with the aim of identifying candidates targeting binding and invasion of host cells by C. jejuni during infection. For example, glucuronic-acid-enriched polysaccharides prepared from Abelmoschus esculentus (okra fruit) efficiently prevented C. jejuni binding to host cells in vitro [61]. In addition, extracts from various citrus plants (Citrus medica, Citrus limon and Citrus aurantium) inhibited adhesion and invasion of C. jejuni into cultured HeLa cells and reduced the expression of ciaB and cadF genes [62]. Other studies have described that peppermint essential oil, a commonly used substance in the treatment of gastrointestinal diseases, led to decreased expression of various C. jejuni virulence-associated genes such as cheY, cadF and the flagellar genes flhB and flgE, which inhibited bacterial motility [63]. Thus, specific extracts from plants may represent innovative and useful new therapeutic reagents to treat C. jejuni infections in future.
In addition, C. jejuni secretes the serine protease HtrA into the extracellular milieu [64]. This protease disrupts tight and adherens junctions by cleavage of important junctional proteins such as occludin [65], claudin-8 [66] and E-cadherin [67,68]. This leads to the opening of cell–cell junctions and epithelial barrier dysfunction during infection. C. jejuni is thus able to overcome the intestinal epithelial barrier by transmigration via the paracellular route, reaching deeper tissue layers and even migrating to other organs such as liver, spleen or mesenteric lymph nodes, which plays a role in the context of bacteremia and sepsis [40]. It appears that C. jejuni HtrA activity also contributes to in vivo pathogenesis in infant mice [69] and interleukin (IL)-10-/- knockout mice [70,71]. Epithelial barrier dysfunction during C. jejuni infection disturbed the ion homeostasis in the intestinal epithelium, which leads to the loss of water and diarrhea, called leaky flux phenotype [72]. Several infection-induced immune mediators, such as tumor necrosis factor (TNF), IL-6, IL-8, IL-12, IL-23 and IL-1β, intensify the barrier damage as discussed below. They also intensify diarrhea, as characterized by sodium malabsorption through dysregulation of the sodium channel ENaC [73]. Recent studies in cell culture in vitro and in mice in vivo indicated a significant reduction of C. jejuni-triggered epithelial barrier dysfunction by application of the polyphenol curcumin [74] or vitamin D [73]. Vitamin D treatment also decreased C. jejuni transmigration across polarized epithelial cells. Thus, curcumin and vitamin D could be promising compounds for the treatment of C. jejuni infection in animals and humans. Finally, the development of small molecule inhibitors interfering with HtrA protease activity is another promising novel approach in anti-bacterial therapy [75]. For instance, computer-assisted de novo design revealed a small inhibitor compound interfering with protease HtrA from Helicobacter pylori, a gastric pathogen evolutionarily related to C. jejuni [76]. Molecular binding and functional activity studies resulted in the identification and characterization of this presently best-in-class HtrA inhibitor. However, while being effective against H. pylori HtrA, this compound only marginally reduced the number of transmigrated C. jejuni. Nevertheless, these data demonstrated the feasibility of pioneering inhibitor compounds with tailor-made activity, which could be also applied to C. jejuni HtrA in the therapy of corresponding infections.
The above-described pathogenic processes of C. jejuni are further enhanced by the CDT (cytolethal distending toxin), which is secreted by subpopulations of the pathogen [77]. CDT function depends on the presence of three genes, named cdtA, cdtB and cdtC. When expressed, the protein subunits CdtA, CdtB and CdtC bind to each other and form a tripartite complex that is called active holotoxin. The CdtA and CdtC subunits are required for the delivery of CdtB into the cytoplasm of target cells. CdtB represents the catalytically active subunit, which displays DNase activity and is transferred to the host cell nucleus. This activity results in chromosomal DNA cleavage, nuclear and cellular swelling and G(2)/M cell cycle arrest [77,78]. As described above, innovative new tools for better controlling the C. jejuni infection are natural products derived from plants. Among those, extracts of silvery mugwort (Artemisia ludoviciana) and sweet acacia (Acacia farnesiana) inhibited both the adherence of C. jejuni to cultured Vero cells as well as CDT activity [79]. Another report investigated the usefulness of the phytochemicals eugenol, carvacrol and trans-cinnamaldehyde during C. jejuni infection of polarized Caco-2 cells [80]. Most of the phytochemicals reduced the expression of CDT and diminished C. jejuni motility, cellular adherence and invasion, as well as bacterial transmigration [80]. Thus, these plant extracts and phytochemicals represent additional candidates to be investigated for prevention or treatment of C. jejuni contamination in food products, and potential anti-microbial therapy of the infections.

4. Environmental Survival Factors Conduct Campylobacteriosis Prevention

Campylobacter are highly susceptible to different stress conditions that are encountered by these bacteria in the environment or in the food chain. Campylobacter have developed mechanisms to adapt to these conditions, however, enabling them to persist in the environment and the food chain and survive technological stressors applied during food processing. Table 1 summarizes the factors that influence the survival of Campylobacter from the farm to the consumer.
A number of studies used genomic and transcriptomic data to investigate mechanisms of stress response and adaption and additionally linked those data to phenotypic assays [110,111,112,113,114]. While the majority of these studies applied a rather mechanistic approach, a recent review focused specifically on the survival and adaptation of C. jejuni within the poultry production chain [109]. De Vries et al. (2017) performed a genome-wide fitness analysis by transposon mutant library screening of C. jejuni. The authors demonstrated that a large part of the genome is related to fitness and survival, with 486 out of 1424 genes coding for fitness factors [115], for type III and type VI secretion systems [116], for the ferric uptake regulator [114] and furthermore, for components of the flagellar system and cell envelope [117]. Increased aerotolerance may be an important survival mechanism of microaerophilic Campylobacter when encountering extra-intestinal environments [118]. Studies showed that aerotolerant or hyper-aerotolerant strains exist that also contain a higher rate of virulence-related genes if compared to aero-sensitive strains [119]. While investigating an aerotolerant C. jejuni strain under aerobic conditions, Rodrigues et al. (2016) identified overexpression of proteins related to oxidative stress response, to amino acid and to iron uptake [120]. The authors concluded that differential gene expression patterns contribute to aerotolerance in the investigated strain rather than the presence of unique stress response genes. Low temperature conditions have been shown to induce oxidative stress response [115]. Subsequently, genes related to oxidative stress response are needed for survival at low temperatures, such as oxidoreductase (trxC) and regulator of oxidative stress (perR). Furthermore, attenuation of mcp4_2 (involved in chemotaxis), kefB and czcD (antiporters) and fabI (fatty acid metabolism) for survival in different habitats was associated with a response to lower temperatures [115]. Campylobacter strains differ significantly in their ability to survive outside their hosts. It is speculated that Campylobacter strains that potentially adapted to the environment have improved fitness through the evolution of stress resistance mechanisms [121,122].
Formation of biofilms or integration into existing biofilms increases the survival potential of bacteria in the environment. Campylobacter can form biofilms or (rather) integrate into pre-existing biofilms present on contact surfaces or equipment of the food chain, such as slaughter equipment, water supplies or plumbing systems [123]. The ability to form biofilms is lower in Campylobacter compared to other bacteria. While monospecies Campylobacter biofilms have not been demonstrated in vivo, secondary colonization or integration into pre-existing biofilms is possible and of practical concern [124]. A summary of genes involved in biofilm formation was recently provided by Püning et al. (2021) [125]. Based on data published by Tram et al. (2020), approximately 600 genes are differentially expressed in Campylobacter during biofilm formation, with induction of genes related to iron metabolism, cell division and glycan production, among others. Repressed genes are linked to metabolism and parts of the chemotaxis pathway [126]. The ability of flagella-associated attachment influences the degree of biofilm formation [127]. Specifically, biofilm formation depends on flagellar O-linked glycan modification [128]. Other factors influencing the degree of biofilm formation are alkyl hydroperoxide reductase (ahpC) and catalase A (katA), highlighting the role of oxidative stress under these circumstances [129]. The degree of the biofilm formation ability is strain-specific in Campylobacter. For instance, some strains are not able to form biofilms at all, whereas others show biofilm formation at different degrees, suggesting that the genetic composition of the specific strain also plays a role in biofilm formation. Recently, Sung and Khan (2015) and Feng et al. (2016) summarized the molecular mechanisms of C. jejuni biofilm formation [130,131]. Briefly, genes involved in motility, flagellar composition (flaAB), oxidative stress response, LOS structure composition (e.g., waaF, lgtF) or cell structure modification (pgp1) are involved in biofilm formation. By combining genotyping data and biofilm assays, Pascoe et al. (2015) identified hotspots of genetic variation that corresponded to specific biofilm phenotypes [132]. Various environmental factors influence biofilm formation as well [129]. For example, Campylobacter biofilm formation under aerobic conditions is promoted by oxidative stress, presence of organic material or extracellular DNA. On the other hand, biofilm formation is reduced by substances causing osmotic stress. Favorable conditions outside or adverse conditions inside the biofilm can activate biofilm dispersion; for example, the accumulation of signal molecules or increased oxygen levels [133]. Potential biofilm control strategies include modifying the surface of contact structures to reduce attachment and disrupting the biofilm matrix through the application of bacteriophages, bacteriocins, DNase or quorum quenchers (summarized in [134]). Preliminary data are available on the reduction of biofilm mass by application of trans-cinnamaldehyde, eugenol or carvacrol after initial biofilm formation [135]. Zinc oxide nanoparticles, which have a high oxidative potential, can also inhibit biofilm formation [136].
Our knowledge of the role of quorum sensing (QS) mechanisms in Campylobacter is still sparse, even though the presence of LuxS and AI-2 production in C. jejuni was already demonstrated in 2002 [137]. We still lack description of autoinducer receptors in Campylobacter. It is speculated that AI-2 is sensed via a two-component regulatory system [138]. There is no clear picture on the role of luxS in C. jejuni. LuxS mutants showed a reduction in growth, motility, biofilm formation, reaction to oxygen stress, adhesion/invasion, and colonization. However, these phenotypes varied considerably between respective studies. Different substances have already been tested for their potential to disrupt QS mechanisms (quorum quenching). Simunovic et al. (2020) tested 20 plant extracts, almost all of which altered QS-related phenotypes of C. jejuni [139]. For instance, citrus extracts lowered AI-2 activity and influenced motility. As mentioned above, adhesion and invasion in cell cultures, expression of virulence factors (cadF, ciaB) and biofilm development were also impaired [62,140].
As described for other bacteria, Campylobacter is capable of transitioning into a viable but non-culturable (VBNC) state under stress conditions, such as aerobic atmosphere, acid stress, starvation or prolonged cold exposure [83]. Notably, Campylobacter in VBNC state exhibit higher resistance to disinfecting agents, initiate biofilm formation and are still infectious.

5. Anti-Microbial and Anti-Inflammatory Therapy of Campylobacteriosis

The value of molecular targets enabling modulation of inflammation in C. jejuni infections is supported by the fact that antibiotics are not indicated to mitigate enteritis symptoms. The intestinal environment reduces antibiotic efficacy by absorptive removal and diarrheal dilution, both favoring resistance development of the pathogen [141]. In addition, inadequate and improper use of antibiotics has increased C. jejuni resistance to macrolides and fluoroquinolones. Both antibiotics were originally reserved for treatment of particularly severe enteritis with systemic manifestations [142]. In consequence, patients receive symptomatic therapies, including rehydration and electrolyte substitution, without causative measures and need to sustain symptoms of intestinal inflammation, which in turn elevates the risk of post-infectious sequelae of infection as outlined above.
Studies of the mechanisms used by C. jejuni to cause intestinal inflammation revealed that the bacteria lack or at least do not rely on exotoxins that are typically produced by other enteric pathogens. Thus, the majority of C. jejuni strains produce neither cholera-like enterotoxin nor CDT [12,13,143,144]. Instead, endotoxins play a major role in the immunopathogenesis of campylobacteriosis. In gastroenterology, campylobacteriosis is perceived as an inflammatory syndrome caused by innate immune cells, such as neutrophilic granulocytes and macrophages, which are activated by direct contact with C. jejuni (reviewed by [145,146]). In this regard, campylobacteriosis shares common features with the acute purulent Neisseria meningitidis and Neisseria gonorrhoeae infections affecting other body compartments [147,148]. Early studies on the histology of C. jejuni infection proved that neutrophils and macrophages accumulate at intestinal sites of C. jejuni invasion [11,147,149]. Upon activation by bacterial endotoxins, these cells trigger inflammation and tissue destruction by the production of pro-inflammatory mediators and reactive oxygen species (ROS). In this scenario, it is tempting to speculate that ROS plays a key role in induction of intestinal apoptosis. This assumption was confirmed by clinical studies, showing that both severity of campylobacteriosis and post-infectious sequelae are significantly associated with distinct LOS variants produced by the infecting C. jejuni strains [15,150]. In addition, LOS was shown to be a master regulator of pathogenesis, triggering inflammation, apoptosis, tissue destruction and diarrhea via active induction of sodium malabsorption [150]. Structural variability of LOS has been further shown to be responsible for the heterogeneity of disease symptoms. Hence, innate immune activation by C. jejuni endotoxins is key to the treatment of campylobacteriosis and prophylaxis of post-infectious sequelae (reviewed by [145]). Both the LOS and respective immune responses induced via its detection by Toll-like-receptor (TLR) 4 serve as potent molecular targets for dampening inflammation and ameliorating campylobacteriosis [145].
Based on these findings, a combination of anti-microbial and anti-inflammatory drugs would represent an ideal treatment option for human campylobacteriosis. Indeed, this conclusion was further supported by results from novel murine infection models that displayed the symptoms as well as the molecular immunopathogenesis induced by C. jejuni infections in humans. Notably, mice showed colonization resistance to C. jejuni, as well as a high tolerance to LOS, mainly due to the fact that the murine TLR4 responses are approximately 10,000-fold weaker [151,152] compared to humans [153]. Hence, the development of suitable murine infection models required modification of both the intestinal microbiota composition and the LOS responses of mice [145,154,155]. Thus, the major role of C. jejuni LOS in human campylobacteriosis was confirmed by manipulation of the murine immune system. Mice with deficiencies in IL-10 [156], single IgG IL-1 Related Receptor (SIGIRR) [154,157] and mice subjected to zinc depletion [158] developed campylobacteriosis symptoms upon C. jejuni infection. Since IL-10 [159,160,161] and SIGIRR [154] signaling pathways, as well as zinc application [162,163,164], effectively suppress LOS and lipo-polysaccharide (LPS)-mediated inflammation, C. jejuni-induced disease in mice is mainly caused by enhanced pro-inflammatory immune responses to enteropathogenic LOS [145]. Due to its suppressive role in endotoxin signaling, oral zinc supplementation is used to protect children in low-income countries from bacterial diarrheal diseases, including campylobacteriosis [165]. Hence, the development of novel murine infection models represents a breakthrough in Campylobacter research and helped to identify LOS-induced inflammation as an innovative molecular target for amelioration of campylobacteriosis in infected humans [145].
In particular, secondary abiotic (SAB) IL-10-deficient mice have proven useful for the analysis of C. jejuni–host interactions, mainly because the immunopathology characterized by granulocyte and macrophage recruitment, by activation of T and B lymphocytes and by colonic epithelial cell apoptosis mirrors the immune and histopathological responses in C. jejuni-infected humans [11,150,166]. In this model, campylobacteriosis depends on the motility and adhesive properties of the pathogen [167]. Most important was the finding that the immunosuppressive drug rapamycin prevented campylobacteriosis in these mice and supported clearing of the pathogen [166], which provided the first experimental proof that the dampening of inflammation is a valuable target for prevention and treatment of C. jejuni infection. Furthermore, preclinical placebo-controlled intervention studies on SAB IL-10-deficient mice revealed that defined vitamins, including ascorbate [168] and vitamin D [73,169], the short chain fatty acid butyrate [170] and plant-derived compounds that have been used for an extensive period of time in traditional medicine, such as essential oils [171,172,173,174], curcumin [74], resveratrol [175], carvacrol [176], urolithin-A [177] and activated charcoal [178], effectively dampened inflammation in the course of murine campylobacteriosis. Their respective mechanisms of action demonstrate that (i) LOS activity, (ii) pro-inflammatory mediators including oxygen radicals, (iii) intestinal barrier function and (iv) C. jejuni-related pathogenicity factors are well suited as molecular targets for the treatment and prevention of campylobacteriosis [157,179,180]. Finally, based on the key role of iron in ROS formation via the Fenton reaction, it was shown that iron deprivation by desferoxamine ameliorated murine campylobacteriosis [181]. In line with the inactivation of ROS by ascorbate (as mentioned above), these data showed that ROS production by innate immune cells represents another molecular target for treatment of human campylobacteriosis. Thus, progress in murine research provided evidence that traditional medicinal drugs may not only ameliorate human campylobacteriosis, but may also reduce the risk for the appearance of post-infectious sequelae such as GBS or reactive arthritis after LOS-induced inflammation. However, this demands further investigation in murine models of GBS [182,183,184]. Hence, the discovery of the pivotal role of LOS in the induction of acute enteritis and of post-infectious sequelae upon C. jejuni infection has greatly contributed to the identification of defined molecules for the treatment and prophylaxis of human campylobacteriosis.

6. Concluding Remarks

The early formulated “endotoxin concept” of C. jejuni-induced inflammatory diarrhea [148,149] proved valuable, and this has inspired the development of novel murine infection models that are useful for preclinical evaluation of therapeutic and prophylactic strategies to combat human campylobacteriosis. Scientific progress in the understanding of molecular host–pathogen interactions underlying C. jejuni colonization, adhesion, invasion and environmental adaptation, and their potential to induce innate immune activation leading to enteritis, have provided molecular targets for innovative treatment options. The principal concept of bacterial LOS playing a pivotal role in the molecular immunopathogenesis of acute campylobacteriosis and its post-infectious sequelae has successfully been used to overcome LOS/LPS tolerance of mice to solve the puzzle of asymptomatic Campylobacter colonization in poultry versus acute disease in infected humans. Resistance to LOS is 100-fold to 1,000,000-fold higher in chickens and other birds than in mice and humans, respectively [153,185]. Thus, LOS might help to understand why chickens and other poultry do not develop intestinal inflammation upon C. jejuni colonization, as was hypothesized earlier [8].
For the future, it is tempting to speculate that combinations of anti-inflammatory and anti-microbial drugs obtained from traditional and conventional medicine will help to dampen both host intestinal inflammation and anti-bacterial resistance (reviewed by [145]). The innovative discoveries summarized here support preventive measures in farming, as well as clinical studies aiming to improve treatment and prophylaxis of human campylobacteriosis.

Funding

This work is supported by grants from the zoonoses research consortium PAC-Campylobacter of the German Federal Ministry of Education and Research (BMBF) to TA (project IP2/01KI1725A), SBa (project IP9/01KI2007E), MMH, SBe, (project IP7/01KI1725), and from the Federal Ministry for Economic Affairs and Energy following a resolution of the German National Parliament, Deutscher Bundestag to MMH and SBe (ZIM, ZF4117908 AJ8).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Bodo Linz (FAU Erlangen) for critical reading and comments on the manuscript. We acknowledge financial support from the Open Access Publication Fund of Charité—Universitätsmedizin Berlin and the German Research Foundation (DFG).

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; or in the writing of the manuscript.

References

  1. Wassenaar, T.M.; Blaser, M.J. Pathophysiology of Campylobacter jejuni infections of humans. Microbes Infect. 1999, 1, 1023–1033. [Google Scholar] [CrossRef]
  2. Altekruse, S.F. Campylobacter, 3rd edition. Emerg. Infect. Dis. J. 2008, 14, 1977. [Google Scholar] [CrossRef]
  3. Dasti, J.I.; Tareen, A.M.; Lugert, R.; Zautner, A.E.; Gross, U. Campylobacter jejuni: A brief overview on pathogenicity-associated factors and disease-mediating mechanisms. Int. J. Med. Microbiol. 2010, 300, 205–211. [Google Scholar] [CrossRef]
  4. Kaakoush, N.O.; Castano-Rodriguez, N.; Mitchell, H.M.; Man, S.M. Global Epidemiology of Campylobacter Infection. Clin. Microbiol. Rev. 2015, 28, 687–720. [Google Scholar] [CrossRef] [Green Version]
  5. Burnham, P.M.; Hendrixson, D.R. Campylobacter jejuni: Collective components promoting a successful enteric lifestyle. Nat. Rev. Microbiol. 2018, 16, 551–565. [Google Scholar] [CrossRef]
  6. Cody, A.J.; Maiden, M.C.; Strachan, N.J.; McCarthy, N.D. A systematic review of source attribution of human campylobacteriosis using multilocus sequence typing. Eurosurveillance 2019, 24, 1800696. [Google Scholar] [CrossRef] [Green Version]
  7. European Food Safety Authority; European Centre for Disease Prevention and Control. The European Union One Health 2021 Zoonoses Report. EFSA J. 2022, 20, e07666. [Google Scholar] [CrossRef]
  8. Young, K.T.; Davis, L.M.; Dirita, V.J. Campylobacter jejuni: Molecular biology and pathogenesis. Nat. Rev. Microbiol. 2007, 5, 665–679. [Google Scholar] [CrossRef]
  9. Smith, O.M.; Snyder, W.E.; Owen, J.P. Are we overestimating risk of enteric pathogen spillover from wild birds to humans? Biol. Rev. 2020, 95, 652–679. [Google Scholar] [CrossRef] [Green Version]
  10. Skirrow, M. Campylobacter enteritis: A” new” disease. Br. Med. J. 1977, 2, 9–11. [Google Scholar] [CrossRef] [Green Version]
  11. Price, A.; Jewkes, J.; Sanderson, P. Acute diarrhoea: Campylobacter colitis and the role of rectal biopsy. J. Clin. Pathol. 1979, 32, 990–997. [Google Scholar] [CrossRef] [Green Version]
  12. Walker, R.I.; Caldwell, M.B.; Lee, E.C.; Guerry, P.; Trust, T.J.; Ruiz-Palacios, G. Pathophysiology of Campylobacter enteritis. Microbiol. Rev. 1986, 50, 81. [Google Scholar] [CrossRef]
  13. Facciolà, A.; Riso, R.; Avventuroso, E.; Visalli, G.; Delia, S.A.; Laganà, P. Campylobacter: From microbiology to prevention. J. Prev. Med. Hyg. 2017, 58, E79–E92. [Google Scholar]
  14. Kist, M.; Bereswill, S. Campylobacter jejuni. Contrib. Microbiol. 2001, 8, 150–165. [Google Scholar] [CrossRef]
  15. Mortensen, N.P.; Kuijf, M.L.; Ang, C.W.; Schiellerup, P.; Krogfelt, K.A.; Jacobs, B.C.; van Belkum, A.; Endtz, H.P.; Bergman, M.P. Sialylation of Campylobacter jejuni lipo-oligosaccharides is associated with severe gastro-enteritis and reactive arthritis. Microbes Infect. 2009, 11, 988–994. [Google Scholar] [CrossRef]
  16. Finsterer, J. Triggers of Guillain-Barre Syndrome: Campylobacter jejuni Predominates. Int. J. Mol. Sci 2022, 23, 14222. [Google Scholar] [CrossRef]
  17. Allos, B.M. Association between Campylobacter infection and Guillain-Barre syndrome. J. Infect. Dis. 1997, 176 (Suppl. S2), S125–S128. [Google Scholar] [CrossRef] [Green Version]
  18. Esan, O.B.; Pearce, M.; van Hecke, O.; Roberts, N.; Collins, D.R.J.; Violato, M.; McCarthy, N.; Perera, R.; Fanshawe, T.R. Factors Associated with Sequelae of Campylobacter and Non-typhoidal Salmonella Infections: A Systematic Review. EBioMedicine 2017, 15, 100–111. [Google Scholar] [CrossRef] [Green Version]
  19. Newman, A.; Lambert, J.R. Campylobacter jejuni causing flare-up in inflammatory bowel disease. Lancet 1980, 2, 919. [Google Scholar] [CrossRef]
  20. Garcia Rodriguez, L.A.; Ruigomez, A.; Panes, J. Acute gastroenteritis is followed by an increased risk of inflammatory bowel disease. Gastroenterology 2006, 130, 1588–1594. [Google Scholar] [CrossRef]
  21. Gradel, K.O.; Nielsen, H.L.; Schonheyder, H.C.; Ejlertsen, T.; Kristensen, B.; Nielsen, H. Increased short- and long-term risk of inflammatory bowel disease after Salmonella or Campylobacter gastroenteritis. Gastroenterology 2009, 137, 495–501. [Google Scholar] [CrossRef]
  22. Kusulja, M.; Santini, M.; Margetic, K.; Guzvinec, M.; Soprek, S.; Butic, I.; Tambic Andrasevic, A. Meningitis caused by Campylobacter jejuni: A case presentation and literature review. Acta Clin. Belg. 2021, 76, 318–323. [Google Scholar] [CrossRef]
  23. Hessulf, F.; Ljungberg, J.; Johansson, P.A.; Lindgren, M.; Engdahl, J. Campylobacter jejuni-associated perimyocarditis: Two case reports and review of the literature. BMC Infect. Dis. 2016, 16, 289. [Google Scholar] [CrossRef] [Green Version]
  24. Keithlin, J.; Sargeant, J.; Thomas, M.K.; Fazil, A. Systematic review and meta-analysis of the proportion of Campylobacter cases that develop chronic sequelae. BMC Public Health 2014, 14, 1203. [Google Scholar] [CrossRef] [Green Version]
  25. Talukder, R.; Sutradhar, S.; Rahman, K.; Uddin, M.; Akhter, H. Guillian-Barre syndrome. Mymensingh Med. J. MMJ 2011, 20, 748–756. [Google Scholar]
  26. Wakerley, B.R.; Uncini, A.; Yuki, N.; Attarian, S.; Barreira, A.A.; Chan, Y.-C.; Créange, A.; Hsieh, S.-T.; Islam, B.; Kannan, M.A. Guillain–Barré and Miller Fisher syndromes—New diagnostic classification. Nat. Rev. Neurol. 2014, 10, 537. [Google Scholar] [CrossRef]
  27. WHO. World Health Organisation. Campylobacter. Available online: https://www.who.int/news-room/fact-sheets/detail/campylobacter (accessed on 4 June 2020).
  28. EFSA. The European Union Summary Report on Trends and Sources of Zoonoses, Zoonotic Agents and Food-borne Outbreaks in 2009. EFSA J. 2011, 9, 2090. [Google Scholar]
  29. EFSA. EFSA explains zoonotic disease-Campylobacter. Available online: https://doi.org/10.2805/59450 (accessed on 4 June 2020).
  30. Stingl, K.; Knüver, M.-T.; Vogt, P.; Buhler, C.; Krüger, N.-J.; Alt, K.; Tenhagen, B.-A.; Hartung, M.; Schroeter, A.; Ellerbroek, L. Quo vadis?—Monitoring Campylobacter in Germany. Eur. J. Microbiol. Immunol. 2012, 2, 88–96. [Google Scholar] [CrossRef] [Green Version]
  31. Newell, D.G. Animal models of Campylobacter jejuni colonization and disease and the lessons to be learned from similar Helicobacter pylori models. Symp. Ser. (Soc. Appl. Microbiol. ) 2001, 90, 57s–67s. [Google Scholar] [CrossRef]
  32. Parkhill, J.; Wren, B.; Mungall, K.; Ketley, J.; Churcher, C.; Basham, D.; Chillingworth, T.; Davies, R.; Feltwell, T.; Holroyd, S. The genome sequence of the food-borne pathogen Campylobacter jejuni reveals hypervariable sequences. Nature 2000, 403, 665–668. [Google Scholar] [CrossRef] [Green Version]
  33. Fouts, D.E.; Mongodin, E.F.; Mandrell, R.E.; Miller, W.G.; Rasko, D.A.; Ravel, J.; Brinkac, L.M.; DeBoy, R.T.; Parker, C.T.; Daugherty, S.C. Major structural differences and novel potential virulence mechanisms from the genomes of multiple Campylobacter species. PLoS Biol. 2005, 3, e15. [Google Scholar] [CrossRef]
  34. Hofreuter, D.; Tsai, J.; Watson, R.O.; Novik, V.; Altman, B.; Benitez, M.; Clark, C.; Perbost, C.; Jarvie, T.; Du, L. Unique features of a highly pathogenic Campylobacter jejuni strain. Infect. Immun. 2006, 74, 4694–4707. [Google Scholar] [CrossRef] [Green Version]
  35. Peh, E.; Kittler, S.; Reich, F.; Kehrenberg, C. Antimicrobial activity of organic acids against Campylobacter spp. and development of combinations—A synergistic effect? PLoS ONE 2020, 15, e0239312. [Google Scholar] [CrossRef]
  36. Hafiz, R.A.; Wong, C.; Paynter, S.; David, M.; Peeters, G. The risk of community-acquired enteric infection in proton pump inhibitor therapy: Systematic review and meta-analysis. Ann. Pharmacother. 2018, 52, 613–622. [Google Scholar] [CrossRef]
  37. Cróinín, T.; Backert, S. Host epithelial cell invasion by Campylobacter jejuni: Trigger or zipper mechanism? Front. Cell Infect. Microbiol. 2012, 2, 25. [Google Scholar] [CrossRef] [Green Version]
  38. Backert, S.; Hofreuter, D. Molecular methods to investigate adhesion, transmigration, invasion and intracellular survival of the foodborne pathogen Campylobacter jejuni. J. Microbiol. Methods 2013, 95, 8–23. [Google Scholar] [CrossRef] [Green Version]
  39. Backert, S.; Tegtmeyer, N.; Cróinín, T.Ó.; Boehm, M.; Heimesaat, M.M. Chapter 1–Human campylobacteriosis. In Campylobacter; Klein, G., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 1–25. [Google Scholar]
  40. Backert, S.; Boehm, M.; Wessler, S.; Tegtmeyer, N. Transmigration route of Campylobacter jejuni across polarized intestinal epithelial cells: Paracellular, transcellular or both? Cell Commun. Signal. CCS 2013, 11, 72. [Google Scholar] [CrossRef] [Green Version]
  41. Heimesaat, M.M.; Backert, S.; Alter, T.; Bereswill, S. Human Campylobacteriosis-A Serious Infectious Threat in a One Health Perspective. Curr. Top. Microbiol. Immunol. 2021, 431, 1–23. [Google Scholar] [CrossRef]
  42. Lertsethtakarn, P.; Ottemann, K.M.; Hendrixson, D.R. Motility and chemotaxis in Campylobacter and Helicobacter. Annu. Rev. Microbiol. 2011, 65, 389–410. [Google Scholar] [CrossRef]
  43. Cayrou, C.; Barratt, N.A.; Ketley, J.M.; Bayliss, C.D. Phase variation during host colonization and invasion by Campylobacter jejuni and other Campylobacter species. Front. Microbiol. 2021, 12, 705139. [Google Scholar] [CrossRef]
  44. Korolik, V. The role of chemotaxis during Campylobacter jejuni colonisation and pathogenesis. Curr. Opin. Microbiol. 2019, 47, 32–37. [Google Scholar] [CrossRef]
  45. Gao, B.; Vorwerk, H.; Huber, C.; Lara-Tejero, M.; Mohr, J.; Goodman, A.L.; Eisenreich, W.; Galán, J.E.; Hofreuter, D. Metabolic and fitness determinants for in vitro growth and intestinal colonization of the bacterial pathogen Campylobacter jejuni. PLoS Biol. 2017, 15, e2001390. [Google Scholar] [CrossRef]
  46. Hartley-Tassell, L.E.; Shewell, L.K.; Day, C.J.; Wilson, J.C.; Sandhu, R.; Ketley, J.M.; Korolik, V. Identification and characterization of the aspartate chemosensory receptor of Campylobacter jejuni. Mol. Microbiol. 2010, 75, 710–730. [Google Scholar] [CrossRef]
  47. Rahman, H.; King, R.M.; Shewell, L.K.; Semchenko, E.A.; Hartley-Tassell, L.E.; Wilson, J.C.; Day, C.J.; Korolik, V. Characterisation of a multi-ligand binding chemoreceptor CcmL (Tlp3) of Campylobacter jejuni. PLoS Pathog. 2014, 10, e1003822. [Google Scholar] [CrossRef] [Green Version]
  48. Johnson, J.G.; Yuhas, C.; McQuade, T.J.; Larsen, M.J.; DiRita, V.J. Narrow-spectrum inhibitors of Campylobacter jejuni flagellar expression and growth. Antimicrob. Agents Chemother. 2015, 59, 3880–3886. [Google Scholar] [CrossRef] [Green Version]
  49. Yabe, S.; Higuchi, W.; Takano, T.; Razvina, O.; Iwao, Y.; Isobe, H.; Yamamoto, T. In vitro susceptibility to antimicrobial agents and ultrastructural characteristics related to swimming motility and drug action in Campylobacter jejuni and C. coli. J. Infect. Chemother. 2010, 16, 174–185. [Google Scholar] [CrossRef] [PubMed]
  50. Szymanski, C.M.; King, M.; Haardt, M.; Armstrong, G.D. Campylobacter jejuni motility and invasion of Caco-2 cells. Infect. Immun. 1995, 63, 4295–4300. [Google Scholar] [CrossRef] [Green Version]
  51. Alemka, A.; Whelan, S.; Gough, R.; Clyne, M.; Gallagher, M.E.; Carrington, S.D.; Bourke, B. Purified chicken intestinal mucin attenuates Campylobacter jejuni pathogenicity in vitro. J. Med. Microbiol. 2010, 59, 898–903. [Google Scholar] [CrossRef]
  52. Eucker, T.P.; Konkel, M.E. The cooperative action of bacterial fibronectin-binding proteins and secreted proteins promote maximal Campylobacter jejuni invasion of host cells by stimulating membrane ruffling. Cell Microbiol. 2012, 14, 226–238. [Google Scholar] [CrossRef] [Green Version]
  53. Tegtmeyer, N.; Sharafutdinov, I.; Harrer, A.; Esmaeili, D.S.; Linz, B.; Backert, S. Campylobacter Virulence Factors and Molecular Host–Pathogen Interactions. Curr. Top. Microbiol. Immunol. 2021, 431, 169–202. [Google Scholar] [CrossRef] [PubMed]
  54. Young, G.M.; Schmiel, D.H.; Miller, V.L. A new pathway for the secretion of virulence factors by bacteria: The flagellar export apparatus functions as a protein-secretion system. Proc. Natl. Acad. Sci. USA 1999, 96, 6456–6461. [Google Scholar] [CrossRef] [Green Version]
  55. Konkel, M.E.; Klena, J.D.; Rivera-Amill, V.; Monteville, M.R.; Biswas, D.; Raphael, B.; Mickelson, J. Secretion of virulence proteins from Campylobacter jejuni is dependent on a functional flagellar export apparatus. J. Bacteriol. 2004, 186, 3296–3303. [Google Scholar] [CrossRef] [Green Version]
  56. Song, Y.; Jin, S.; Louie, H.; Ng, D.; Lau, R.; Zhang, Y.; Weerasekera, R.; Al Rashid, S.; Ward, L.; Der, S. FlaC, a protein of Campylobacter jejuni TGH9011 (ATCC43431) secreted through the flagellar apparatus, binds epithelial cells and influences cell invasion. Mol. Microbiol. 2004, 53, 541–553. [Google Scholar] [CrossRef]
  57. Krause-Gruszczynska, M.; Rohde, M.; Hartig, R.; Genth, H.; Schmidt, G.; Keo, T.; König, W.; Miller, W.G.; Konkel, M.E.; Backert, S. Role of the small Rho GTPases Rac1 and Cdc42 in host cell invasion of Campylobacter jejuni. Cell. Microbiol. 2007, 9, 2431–2444. [Google Scholar] [CrossRef]
  58. Krause-Gruszczynska, M.; Boehm, M.; Rohde, M.; Tegtmeyer, N.; Takahashi, S.; Buday, L.; Oyarzabal, O.A.; Backert, S. The signaling pathway of Campylobacter jejuni-induced Cdc42 activation: Role of fibronectin, integrin beta1, tyrosine kinases and guanine exchange factor Vav2. Cell Commun. Signal. 2011, 9, 32. [Google Scholar] [CrossRef] [Green Version]
  59. Boehm, M.; Krause-Gruszczynska, M.; Rohde, M.; Tegtmeyer, N.; Takahashi, S.; Oyarzabal, O.A.; Backert, S. Major host factors involved in epithelial cell invasion of Campylobacter jejuni: Role of fibronectin, integrin beta1, FAK, Tiam-1, and DOCK180 in activating Rho GTPase Rac1. Front. Cell. Infect. Microbiol. 2011, 1, 17. [Google Scholar] [CrossRef] [Green Version]
  60. Negretti, N.M.; Gourley, C.R.; Talukdar, P.K.; Clair, G.; Klappenbach, C.M.; Lauritsen, C.J.; Adkins, J.N.; Konkel, M.E. The Campylobacter jejuni CiaD effector co-opts the host cell protein IQGAP1 to promote cell entry. Nat. Commun. 2021, 12, 1339. [Google Scholar] [CrossRef]
  61. Wittschier, N.; Lengsfeld, C.; Vorthems, S.; Stratmann, U.; Ernst, J.; Verspohl, E.; Hensel, A. Large molecules as anti-adhesive compounds against pathogens. J. Pharm. Pharmacol. 2007, 59, 777–786. [Google Scholar] [CrossRef]
  62. Castillo, S.; Dávila-Aviña, J.; Heredia, N.; Garcia, S. Antioxidant activity and influence of Citrus byproduct extracts on adherence and invasion of Campylobacter jejuni and on the relative expression of cadF and ciaB. Food Sci. Biotechnol. 2017, 26, 453–459. [Google Scholar] [CrossRef]
  63. Kovács, J.; Felső, P.; Horváth, G.; Schmidt, J.; Dorn, Á.; Ábrahám, H.; Cox, A.; Márk, L.; Emődy, L.; Kovács, T. Stress response and virulence potential modulating effect of peppermint essential oil in Campylobacter jejuni. BioMed Res. Int. 2019, 2019, 2971741. [Google Scholar] [CrossRef] [Green Version]
  64. Sharafutdinov, I.; Tegtmeyer, N.; Müsken, M.; Backert, S. Campylobacter jejuni Serine Protease HtrA Induces Paracellular Transmigration of Microbiota across Polarized Intestinal Epithelial Cells. Biomolecules 2022, 12, 521. [Google Scholar] [CrossRef] [PubMed]
  65. Harrer, A.; Bücker, R.; Boehm, M.; Zarzecka, U.; Tegtmeyer, N.; Sticht, H.; Schulzke, J.D.; Backert, S. Campylobacter jejuni enters gut epithelial cells and impairs intestinal barrier function through cleavage of occludin by serine protease HtrA. Gut Pathog. 2019, 11, 4. [Google Scholar] [CrossRef]
  66. Sharafutdinov, I.; Esmaeili, D.S.; Harrer, A.; Tegtmeyer, N.; Sticht, H.; Backert, S. Campylobacter jejuni serine protease HtrA cleaves the tight junction component claudin-8. Front. Cell. Infect. Microbiol. 2020, 10, 590186. [Google Scholar] [CrossRef]
  67. Boehm, M.; Hoy, B.; Rohde, M.; Tegtmeyer, N.; Bæk, K.T.; Oyarzabal, O.A.; Brøndsted, L.; Wessler, S.; Backert, S. Rapid paracellular transmigration of Campylobacter jejuni across polarized epithelial cells without affecting TER: Role of proteolytic-active HtrA cleaving E-cadherin but not fibronectin. Gut Pathog. 2012, 4, 3. [Google Scholar] [CrossRef] [Green Version]
  68. Hoy, B.; Geppert, T.; Boehm, M.; Reisen, F.; Plattner, P.; Gadermaier, G.; Sewald, N.; Ferreira, F.; Briza, P.; Schneider, G. Distinct roles of secreted HtrA proteases from gram-negative pathogens in cleaving the junctional protein and tumor suppressor E-cadherin. J. Biol. Chem. 2012, 287, 10115–10120. [Google Scholar] [CrossRef] [Green Version]
  69. Heimesaat, M.M.; Fischer, A.; Alutis, M.; Grundmann, U.; Boehm, M.; Tegtmeyer, N.; Göbel, U.B.; Kühl, A.A.; Bereswill, S.; Backert, S. The impact of serine protease HtrA in apoptosis, intestinal immune responses and extra-intestinal histopathology during Campylobacter jejuni infection of infant mice. Gut Pathog. 2014, 6, 16. [Google Scholar] [CrossRef] [Green Version]
  70. Heimesaat, M.M.; Alutis, M.; Grundmann, U.; Fischer, A.; Tegtmeyer, N.; Bohm, M.; Kuhl, A.A.; Gobel, U.B.; Backert, S.; Bereswill, S. The role of serine protease HtrA in acute ulcerative enterocolitis and extra-intestinal immune responses during Campylobacter jejuni infection of gnotobiotic IL-10 deficient mice. Front. Cell Infect. Microbiol. 2014, 4, 77. [Google Scholar] [CrossRef] [Green Version]
  71. Schmidt, A.M.; Escher, U.; Mousavi, S.; Boehm, M.; Backert, S.; Bereswill, S.; Heimesaat, M.M. Protease Activity of Campylobacter jejuni HtrA Modulates Distinct Intestinal and Systemic Immune Responses in Infected Secondary Abiotic IL-10 Deficient Mice. Front. Cell Infect. Microbiol. 2019, 9, 79. [Google Scholar] [CrossRef]
  72. Lobo de Sá, F.; Schulzke, J.-D.; Bücker, R. Diarrheal mechanisms and the role of intestinal barrier dysfunction in Campylobacter infections. Curr. Top. Microbiol. Immunol. 2021, 431, 203–231. [Google Scholar]
  73. Lobo de Sá, F.D.; Backert, S.; Nattramilarasu, P.K.; Mousavi, S.; Sandle, G.I.; Bereswill, S.; Heimesaat, M.M.; Schulzke, J.-D.; Bücker, R. Vitamin D Reverses Disruption of Gut Epithelial Barrier Function Caused by Campylobacter jejuni. Int. J. Mol. Sci. 2021, 22, 8872. [Google Scholar] [CrossRef]
  74. Lobo de Sa, F.D.; Butkevych, E.; Nattramilarasu, P.K.; Fromm, A.; Mousavi, S.; Moos, V.; Golz, J.C.; Stingl, K.; Kittler, S.; Seinige, D.; et al. Curcumin Mitigates Immune-Induced Epithelial Barrier Dysfunction by Campylobacter jejuni. Int. J. Mol. Sci. 2019, 20, 4830. [Google Scholar] [CrossRef] [Green Version]
  75. Wessler, S.; Schneider, G.; Backert, S. Bacterial serine protease HtrA as a promising new target for antimicrobial therapy? Cell Commun. Signal. 2017, 15, 4. [Google Scholar] [CrossRef] [Green Version]
  76. Perna, A.M.; Rodrigues, T.; Schmidt, T.P.; Böhm, M.; Stutz, K.; Reker, D.; Pfeiffer, B.; Altmann, K.H.; Backert, S.; Wessler, S. Fragment-based de novo design reveals a small-molecule inhibitor of Helicobacter pylori HtrA. Angew. Chem. 2015, 127, 10382–10386. [Google Scholar] [CrossRef]
  77. Lara-Tejero, M.; Galán, J.E. A bacterial toxin that controls cell cycle progression as a deoxyribonuclease I-like protein. Science 2000, 290, 354–357. [Google Scholar] [CrossRef] [Green Version]
  78. Lara-Tejero, M.a.; Galán, J.E. CdtA, CdtB, and CdtC form a tripartite complex that is required for cytolethal distending toxin activity. Infect. Immun. 2001, 69, 4358–4365. [Google Scholar] [CrossRef] [Green Version]
  79. Castillo, S.L.; Heredia, N.; Contreras, J.F.; García, S. Extracts of edible and medicinal plants in inhibition of growth, adherence, and cytotoxin production of Campylobacter jejuni and Campylobacter coli. J. Food Sci. 2011, 76, M421–M426. [Google Scholar] [CrossRef]
  80. Upadhyay, A.; Arsi, K.; Wagle, B.R.; Upadhyaya, I.; Shrestha, S.; Donoghue, A.M.; Donoghue, D.J. Trans-cinnamaldehyde, carvacrol, and eugenol reduce Campylobacter jejuni colonization factors and expression of virulence genes in vitro. Front. Microbiol. 2017, 8, 713. [Google Scholar] [CrossRef] [Green Version]
  81. Skelly, C.; Weinstein, P. Pathogen survival trajectories: An eco-environmental approach to the modeling of human campylobacteriosis ecology. Env. Health Perspect. 2003, 111, 19–28. [Google Scholar] [CrossRef] [Green Version]
  82. Elmonir, W.; Vetchapitak, T.; Amano, T.; Taniguchi, T.; Misawa, N. Survival capability of Campylobacter upsaliensis under environmental stresses. BMC Res. Notes 2022, 15, 47. [Google Scholar] [CrossRef]
  83. Meunier, M.; Guyard-Nicodeme, M.; Dory, D.; Chemaly, M. Control Strategies against Campylobacter at the Poultry Production Level: Biosecurity Measures, Feed Additives and Vaccination. J. Appl. Microbiol. 2016, 120, 1139–1173. [Google Scholar] [CrossRef] [Green Version]
  84. Reichelt, B.; Szott, V.; Epping, L.; Semmler, T.; Merle, R.; Roesler, U.; Friese, A. Transmission pathways of Campylobacter spp. at broiler farms and their environment in Brandenburg, Germany. Front. Microbiol. 2022, 13, 982693. [Google Scholar] [CrossRef]
  85. Meunier, M.; Guyard-Nicodeme, M.; Hirchaud, E.; Parra, A.; Chemaly, M.; Dory, D. Identification of Novel Vaccine Candidates against Campylobacter through Reverse Vaccinology. J. Immunol. Res. 2016, 2016, 5715790. [Google Scholar] [CrossRef] [Green Version]
  86. Neal-McKinney, J.M.; Samuelson, D.R.; Eucker, T.P.; Nissen, M.S.; Crespo, R.; Konkel, M.E. Reducing Campylobacter jejuni colonization of poultry via vaccination. PLoS ONE 2014, 9, e114254. [Google Scholar] [CrossRef] [Green Version]
  87. Wagenaar, J.A.; Bergen, M.A.; Mueller, M.A.; Wassenaar, T.M.; Carlton, R.M. Phage therapy reduces Campylobacter jejuni colonization in broilers. Vet. Microbiol. 2005, 109, 275–283. [Google Scholar] [CrossRef]
  88. Kittler, S.; Steffan, S.; Peh, E.; Plotz, M. Phage Biocontrol of Campylobacter: A One Health Approach. Curr. Top. Microbiol. Immunol. 2021, 431, 127–168. [Google Scholar] [CrossRef]
  89. Szott, V.; Reichelt, B.; Friese, A.; Roesler, U. A Complex Competitive Exclusion Culture Reduces Campylobacter jejuni Colonization in Broiler Chickens at Slaughter Age In Vivo. Vet. Sci. 2022, 9, 181. [Google Scholar] [CrossRef]
  90. Balta, I.; Butucel, E.; Stef, L.; Pet, I.; Gradisteanu-Pircalabioru, G.; Chifiriuc, C.; Gundogdu, O.; McCleery, D.; Corcionivoschi, N. Anti-Campylobacter Probiotics: Latest Mechanistic Insights. Foodborne Pathog. Dis. 2022, 19, 693–703. [Google Scholar] [CrossRef]
  91. Svetoch, E.A.; Stern, N.J. Bacteriocins to control Campylobacter spp. in poultry--A review. Poult. Sci. 2010, 89, 1763–1768. [Google Scholar] [CrossRef]
  92. Szott, V.; Reichelt, B.; Alter, T.; Friese, A.; Roesler, U. In vivo efficacy of carvacrol on Campylobacter jejuni prevalence in broiler chickens during an entire fattening period. Eur. J. Microbiol. Immunol. 2020, 10, 131–138. [Google Scholar] [CrossRef]
  93. Szott, V.; Peh, E.; Friese, A.; Roesler, U.; Kehrenberg, C.; Ploetz, M.; Kittler, S. Antimicrobial effect of a drinking water additive comprising four organic acids on Campylobacter load in broilers and monitoring of bacterial susceptibility. Poult. Sci. 2022, 101, 102209. [Google Scholar] [CrossRef]
  94. Yang, H.; Li, Y.; Johnson, M.G. Survival and death of Salmonella typhimurium and Campylobacter jejuni in processing water and on chicken skin during poultry scalding and chilling. J. Food Prot. 2001, 64, 770–776. [Google Scholar] [CrossRef]
  95. Lehner, Y.; Reich, F.; Klein, G. Influence of process parameter on Campylobacter spp. counts on poultry meat in a slaughterhouse environment. Curr. Microbiol. 2014, 69, 240–244. [Google Scholar] [CrossRef]
  96. Boysen, L.; Rosenquist, H. Reduction of Thermotolerant Campylobacter Species on Broiler Carcasses following Physical Decontamination at Slaughter. J. Food Prot. 2009, 72, 497–502. [Google Scholar] [CrossRef]
  97. Musavian, H.S.; Krebs, N.H.; Nonboe, U.; Corry, J.E.; Purnell, G. Combined steam and ultrasound treatment of broilers at slaughter: A promising intervention to significantly reduce numbers of naturally occurring campylobacters on carcasses. Int. J. Food Microbiol. 2014, 176, 23–28. [Google Scholar] [CrossRef]
  98. Purnell, G.; Mattick, K.; Humphrey, T. The use of ‘hot wash’ treatments to reduce the number of pathogenic and spoilage bacteria on raw retail poultry. J. Food Eng. 2004, 62, 29–36. [Google Scholar] [CrossRef]
  99. Northcutt, J.K.; Smith, D.P.; Musgrove, M.T.; Ingram, K.D.; Hinton, A., Jr. Microbiological impact of spray washing broiler carcasses using different chlorine concentrations and water temperatures. Poult. Sci. 2005, 84, 1648–1652. [Google Scholar] [CrossRef]
  100. Northcutt, J.; Smith, D.; Ingram, K.D.; Hinton, A.; Musgrove, M. Recovery of bacteria from broiler carcasses after spray washing with acidified electrolyzed water or sodium hypochlorite solutions. Poult. Sci. 2007, 86, 2239–2244. [Google Scholar] [CrossRef]
  101. Giombelli, A.; Hammerschmitt, D.; Cerutti, M.F.; Chiarini, E.; Landgraf, M.; Franco, B.D.G.M.; Destro, M.T. High pressure spray with water shows similar efficiency to trimming in controlling microorganisms on poultry carcasses. Poult. Sci. 2015, 94, 2589–2595. [Google Scholar] [CrossRef]
  102. Nagel, G.M.; Bauermeister, L.J.; Bratcher, C.L.; Singh, M.; McKee, S.R. Salmonella and Campylobacter reduction and quality characteristics of poultry carcasses treated with various antimicrobials in a post-chill immersion tank. Int. J. Food Microbiol. 2013, 165, 281–286. [Google Scholar] [CrossRef]
  103. Birk, T.; Gronlund, A.C.; Christensen, B.B.; Knochel, S.; Lohse, K.; Rosenquist, H. Effect of Organic Acids and Marination Ingredients on the Survival of Campylobacter jejuni on Meat. J. Food Prot. 2010, 73, 258–265. [Google Scholar] [CrossRef]
  104. Zakariene, G.; Rokaityte, A.; Ramonaite, S.; Novoslavskij, A.; Mulkyte, K.; Zaborskiene, G.; Malakauskas, M. The antimicrobial effect of spice-based marinades against Campylobacter jejuni on contaminated fresh broiler wings. J. Food Sci. 2015, 80, M627–M634. [Google Scholar] [CrossRef]
  105. Isohanni, P.; Alter, T.; Saris, P.; Lyhs, U. Wines as possible meat marinade ingredients possess antimicrobial potential against Campylobacter. Poult. Sci. 2010, 89, 2704–2710. [Google Scholar] [CrossRef]
  106. Meredith, H.; Valdramidis, V.; Rotabakk, B.T.; Sivertsvik, M.; McDowell, D.; Bolton, D.J. Effect of different modified atmospheric packaging (MAP) gaseous combinations on Campylobacter and the shelf-life of chilled poultry fillets. Food Microbiol. 2014, 44, 196–203. [Google Scholar] [CrossRef]
  107. Murphy, C.; Carroll, C.; Jordan, K.N. Environmental survival mechanisms of the foodborne pathogen Campylobacter jejuni. J. Appl. Microbiol. 2006, 100, 623–632. [Google Scholar] [CrossRef]
  108. Hakeem, M.J.; Feng, J.; Nilghaz, A.; Ma, L.; Seah, H.C.; Konkel, M.E.; Lu, X. Active Packaging of Immobilized Zinc Oxide Nanoparticles Controls Campylobacter jejuni in Raw Chicken Meat. Appl. Environ. Microbiol. 2020, 86, e01195-20. [Google Scholar] [CrossRef]
  109. Hakeem, M.J.; Lu, X. Survival and Control of Campylobacter in Poultry Production Environment. Front. Cell. Infect. Microbiol. 2020, 10, 615049. [Google Scholar] [CrossRef]
  110. Golz, J.C.; Epping, L.; Knuver, M.T.; Borowiak, M.; Hartkopf, F.; Deneke, C.; Malorny, B.; Semmler, T.; Stingl, K. Whole genome sequencing reveals extended natural transformation in Campylobacter impacting diagnostics and the pathogens adaptive potential. Sci Rep. 2020, 10, 3686. [Google Scholar] [CrossRef] [Green Version]
  111. Gundogdu, O.; da Silva, D.T.; Mohammad, B.; Elmi, A.; Wren, B.W.; van Vliet, A.H.; Dorrell, N. The Campylobacter jejuni Oxidative Stress Regulator RrpB Is Associated with a Genomic Hypervariable Region and Altered Oxidative Stress Resistance. Front. Microbiol. 2016, 7, 2117. [Google Scholar] [CrossRef] [Green Version]
  112. Song, H.; Kim, J.; Guk, J.H.; An, J.U.; Lee, S.; Cho, S. Complete genome sequence and comparative genomic analysis of hyper-aerotolerant Campylobacter lari strain SCHS02 isolated from duck for its potential pathogenicity. Microb. Pathog. 2020, 142, 104110. [Google Scholar] [CrossRef]
  113. Hur, J.I.; Kim, J.; Ryu, S.; Jeon, B. Phylogenetic Association and Genetic Factors in Cold Stress Tolerance in Campylobacter jejuni. Microbiol. Spectr. 2022, 10, e0268122. [Google Scholar] [CrossRef]
  114. Askoura, M.; Sarvan, S.; Couture, J.F.; Stintzi, A. The Campylobacter jejuni Ferric Uptake Regulator Promotes Acid Survival and Cross-Protection against Oxidative Stress. Infect. Immun. 2016, 84, 1287–1300. [Google Scholar] [CrossRef] [Green Version]
  115. de Vries, S.P.; Gupta, S.; Baig, A.; Wright, E.; Wedley, A.; Jensen, A.N.; Lora, L.L.; Humphrey, S.; Skovgård, H.; Macleod, K. Genome-wide fitness analyses of the foodborne pathogen Campylobacter jejuni in in vitro and in vivo models. Sci. Rep. 2017, 7, 1251. [Google Scholar] [CrossRef] [Green Version]
  116. Liaw, J.; Hong, G.; Davies, C.; Elmi, A.; Sima, F.; Stratakos, A.; Stef, L.; Pet, I.; Hachani, A.; Corcionivoschi, N.; et al. The Campylobacter jejuni Type VI Secretion System Enhances the Oxidative Stress Response and Host Colonization. Front. Microbiol. 2019, 10, 2864. [Google Scholar] [CrossRef]
  117. Moen, B.; Oust, A.; Langsrud, O.; Dorrell, N.; Marsden, G.L.; Hinds, J.; Kohler, A.; Wren, B.W.; Rudi, K. Explorative multifactor approach for investigating global survival mechanisms of Campylobacter jejuni under environmental conditions. Appl. Environ. Microbiol. 2005, 71, 2086–2094. [Google Scholar] [CrossRef] [Green Version]
  118. O’Kane, P.M.; Connerton, I.F. Characterisation of aerotolerant forms of a robust chicken colonizing Campylobacter coli. Front. Microbiol. 2017, 8, 513. [Google Scholar] [CrossRef] [Green Version]
  119. Oh, E.; McMullen, L.M.; Chui, L.; Jeon, B. Differential survival of hyper-aerotolerant Campylobacter jejuni under different gas conditions. Front. Microbiol. 2017, 8, 954. [Google Scholar] [CrossRef] [Green Version]
  120. Rodrigues, R.C.; Haddad, N.; Chevret, D.; Cappelier, J.-M.; Tresse, O. Comparison of proteomics profiles of Campylobacter jejuni strain Bf under microaerobic and aerobic conditions. Front. Microbiol. 2016, 7, 1596. [Google Scholar] [CrossRef] [Green Version]
  121. Bronowski, C.; James, C.E.; Winstanley, C. Role of environmental survival in transmission of Campylobacter jejuni. FEMS Microbiol. Lett. 2014, 356, 8–19. [Google Scholar] [CrossRef] [Green Version]
  122. Sopwith, W.; Birtles, A.; Matthews, M.; Fox, A.; Gee, S.; Painter, M.; Regan, M.; Syed, Q.; Bolton, E. Identification of potential environmentally adapted Campylobacter jejuni strain, United Kingdom. Emerg. Infect. Dis. 2008, 14, 1769–1773. [Google Scholar] [CrossRef]
  123. Teh, A.H.; Lee, S.M.; Dykes, G.A. Does Campylobacter jejuni form biofilms in food-related environments? Appl. Environ. Microbiol. 2014, 80, 5154–5160. [Google Scholar] [CrossRef] [Green Version]
  124. Hanning, I.; Jarquin, R.; Slavik, M. Campylobacter jejuni as a secondary colonizer of poultry biofilms. J. Appl. Microbiol. 2008, 105, 1199–1208. [Google Scholar] [CrossRef]
  125. Püning, C.; Su, Y.; Lu, X.; Gölz, G. Molecular mechanisms of Campylobacter biofilm formation and quorum sensing. In Fighting Campylobacter Infections: Towards a One Health Approach; Springer: Berlin/Heidelberg, Germany, 2021; pp. 293–319. [Google Scholar]
  126. Tram, G.; Klare, W.P.; Cain, J.A.; Mourad, B.; Cordwell, S.J.; Day, C.J.; Korolik, V. Assigning a role for chemosensory signal transduction in Campylobacter jejuni biofilms using a combined omics approach. Sci. Rep. 2020, 10, 6829. [Google Scholar] [CrossRef] [Green Version]
  127. Joshua, G.W.; Guthrie-Irons, C.; Karlyshev, A.V.; Wren, B.W. Biofilm formation in Campylobacter jejuni. Microbiology 2006, 152, 387–396. [Google Scholar] [CrossRef] [Green Version]
  128. Howard, S.L.; Jagannathan, A.; Soo, E.C.; Hui, J.P.; Aubry, A.J.; Ahmed, I.; Karlyshev, A.; Kelly, J.F.; Jones, M.A.; Stevens, M.P. Campylobacter jejuni glycosylation island important in cell charge, legionaminic acid biosynthesis, and colonization of chickens. Infect. Immun. 2009, 77, 2544–2556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Oh, E.; Jeon, B. Role of alkyl hydroperoxide reductase (AhpC) in the biofilm formation of Campylobacter jejuni. PLoS ONE 2014, 9, e87312. [Google Scholar] [CrossRef] [Green Version]
  130. Sung, K.; Khan, S. Biofilm development by Campylobacter jejuni. In Biofilms in the Food Environment; Wiley Online Library: Hoboken, NJ, USA, 2015; pp. 29–50. [Google Scholar]
  131. Feng, J.; Lamour, G.; Xue, R.; Mirvakliki, M.N.; Hatzikiriakos, S.G.; Xu, J.; Li, H.; Wang, S.; Lu, X. Chemical, physical and morphological properties of bacterial biofilms affect survival of encased Campylobacter jejuni F38011 under aerobic stress. Int. J. Food Microbiol. 2016, 238, 172–182. [Google Scholar] [CrossRef]
  132. Pascoe, B.; Meric, G.; Murray, S.; Yahara, K.; Mageiros, L.; Bowen, R.; Jones, N.H.; Jeeves, R.E.; Lappin-Scott, H.M.; Asakura, H.; et al. Enhanced biofilm formation and multi-host transmission evolve from divergent genetic backgrounds in Campylobacter jejuni. Environ. Microbiol. 2015, 17, 4779–4789. [Google Scholar] [CrossRef] [Green Version]
  133. Petrova, O.E.; Sauer, K. Escaping the biofilm in more than one way: Desorption, detachment or dispersion. Curr. Opin. Microbiol. 2016, 30, 67–78. [Google Scholar] [CrossRef] [Green Version]
  134. Puning, C.; Su, Y.; Lu, X.; Golz, G. Molecular Mechanisms of Campylobacter Biofilm Formation and Quorum Sensing. Curr. Top. Microbiol. Immunol. 2021, 431, 293–319. [Google Scholar] [CrossRef]
  135. Wagle, B.R.; Upadhyay, A.; Upadhyaya, I.; Shrestha, S.; Arsi, K.; Liyanage, R.; Venkitanarayanan, K.; Donoghue, D.J.; Donoghue, A.M. Trans-Cinnamaldehyde, Eugenol and Carvacrol Reduce Campylobacter jejuni Biofilms and Modulate Expression of Select Genes and Proteins. Front. Microbiol. 2019, 10, 1837. [Google Scholar] [CrossRef] [Green Version]
  136. Zhong, X.; Wu, Q.; Zhang, J.; Ma, Z.; Wang, J.; Nie, X.; Ding, Y.; Xue, L.; Chen, M.; Wu, S.; et al. Campylobacter jejuni Biofilm Formation Under Aerobic Conditions and Inhibition by ZnO Nanoparticles. Front. Microbiol. 2020, 11, 207. [Google Scholar] [CrossRef]
  137. Elvers, K.T.; Park, S.F. Quorum sensing in Campylobacter jejuni: Detection of a luxS encoded signalling molecule. Microbiology 2002, 148, 1475–1481. [Google Scholar] [CrossRef] [Green Version]
  138. Adler, L.; Alter, T.; Sharbati, S.; Golz, G. The signalling molecule Autoinducer-2 is not internalised in Campylobacter jejuni. Berl. Munch. Tierarztl. Wochenschr. 2015, 128, 111–116. [Google Scholar]
  139. Simunovic, K.; Ramic, D.; Xu, C.; Smole Mozina, S. Modulation of Campylobacter jejuni Motility, Adhesion to Polystyrene Surfaces, and Invasion of INT407 Cells by Quorum-Sensing Inhibition. Microorganisms 2020, 8, 104. [Google Scholar] [CrossRef] [Green Version]
  140. Castillo, S.; Heredia, N.; Arechiga-Carvajal, E.; García, S. Citrus Extracts as Inhibitors of Quorum Sensing, Biofilm Formation and Motility of Campylobacter jejuni. Food Biotechnol. 2014, 28, 106–122. [Google Scholar] [CrossRef]
  141. Ternhag, A.; Asikainen, T.; Giesecke, J.; Ekdahl, K. A meta-analysis on the effects of antibiotic treatment on duration of symptoms caused by infection with Campylobacter species. Clin. Infect. Dis. 2007, 44, 696–700. [Google Scholar] [CrossRef]
  142. Lübbert, C. Antimicrobial therapy of acute diarrhoea: A clinical review. Expert Rev. Anti-Infect. Ther. 2016, 14, 193–206. [Google Scholar] [CrossRef]
  143. Pickett, C.L.; Whitehouse, C.A. The cytolethal distending toxin family. Trends Microbiol. 1999, 7, 292–297. [Google Scholar] [CrossRef]
  144. Bang, D.D.; Scheutz, F.; Ahrens, P.; Pedersen, K.; Blom, J.; Madsen, M. Prevalence of cytolethal distending toxin (cdt) genes and CDT production in Campylobacter spp. isolated from Danish broilers. J. Med. Microbiol. 2001, 50, 1087–1094. [Google Scholar] [CrossRef] [PubMed]
  145. Mousavi, S.; Bereswill, S.; Heimesaat, M.M. Novel Clinical Campylobacter jejuni Infection Models Based on Sensitization of Mice to Lipooligosaccharide, a Major Bacterial Factor Triggering Innate Immune Responses in Human Campylobacteriosis. Microorganisms 2020, 8, 482. [Google Scholar] [CrossRef] [Green Version]
  146. Phongsisay, V. The immunobiology of Campylobacter jejuni: Innate immunity and autoimmune diseases. Immunobiology 2016, 221, 535–543. [Google Scholar] [CrossRef]
  147. Black, R.E.; Levine, M.M.; Clements, M.L.; Hughes, T.P.; Blaser, M.J. Experimental Campylobacter jejuni infection in humans. J. Infect. Dis. 1988, 157, 472–479. [Google Scholar] [CrossRef]
  148. Moran, A.P.; Prendergast, M.M.; Appelmelk, B.J. Molecular mimicry of host structures by bacterial lipopolysaccharides and its contribution to disease. FEMS Immunol. Med. Microbiol. 1996, 16, 105–115. [Google Scholar] [CrossRef]
  149. Blaser, M.J.; Berkowitz, I.D.; LaForce, F.M.; Cravens, J.; Reller, L.B.; Wang, W.L. Campylobacter enteritis: Clinical and epidemiologic features. Ann. Intern. Med. 1979, 91, 179–185. [Google Scholar] [CrossRef]
  150. Bücker, R.; Krug, S.M.; Moos, V.; Bojarski, C.; Schweiger, M.R.; Kerick, M.; Fromm, A.; Janssen, S.; Fromm, M.; Hering, N.A.; et al. Campylobacter jejuni impairs sodium transport and epithelial barrier function via cytokine release in human colon. Mucosal Immunol. 2018, 11, 575–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Warren, H.S.; Fitting, C.; Hoff, E.; Adib-Conquy, M.; Beasley-Topliffe, L.; Tesini, B.; Liang, X.; Valentine, C.; Hellman, J.; Hayden, D. Resilience to bacterial infection: Difference between species could be due to proteins in serum. J. Infect. Dis. 2010, 201, 223–232. [Google Scholar] [CrossRef] [Green Version]
  152. Munford, R.S. Murine responses to endotoxin: Another dirty little secret? J. Infect. Dis. 2010, 201, 175–177. [Google Scholar] [CrossRef]
  153. Taveira da Silva, A.M.; Kaulbach, H.C.; Chuidian, F.S.; Lambert, D.R.; Suffredini, A.F.; Danner, R.L. Brief report: Shock and multiple-organ dysfunction after self-administration of Salmonella endotoxin. N. Engl. J. Med. 1993, 328, 1457–1460. [Google Scholar] [CrossRef]
  154. Stahl, M.; Ries, J.; Vermeulen, J.; Yang, H.; Sham, H.P.; Crowley, S.M.; Badayeva, Y.; Turvey, S.E.; Gaynor, E.C.; Li, X.; et al. A novel mouse model of Campylobacter jejuni gastroenteritis reveals key pro-inflammatory and tissue protective roles for Toll-like receptor signaling during infection. PLoS Pathog. 2014, 10, e1004264. [Google Scholar] [CrossRef] [Green Version]
  155. Bereswill, S.; Fischer, A.; Plickert, R.; Haag, L.M.; Otto, B.; Kuhl, A.A.; Dasti, J.I.; Zautner, A.E.; Munoz, M.; Loddenkemper, C.; et al. Novel murine infection models provide deep insights into the “menage a trois” of Campylobacter jejuni, microbiota and host innate immunity. PLoS ONE 2011, 6, e20953. [Google Scholar] [CrossRef]
  156. Mansfield, L.S.; Bell, J.A.; Wilson, D.L.; Murphy, A.J.; Elsheikha, H.M.; Rathinam, V.A.; Fierro, B.R.; Linz, J.E.; Young, V.B. C57BL/6 and congenic interleukin-10-deficient mice can serve as models of Campylobacter jejuni colonization and enteritis. Infect. Immun. 2007, 75, 1099–1115. [Google Scholar] [CrossRef] [Green Version]
  157. Stahl, M.; Graef, F.A.; Vallance, B.A. Mouse Models for Campylobacter jejuni Colonization and Infection. In Campylobacter jejuni: Methods and Protocols; Springer: Berlin/Heidelberg, Germany, 2017; pp. 171–188. [Google Scholar]
  158. Giallourou, N.; Medlock, G.L.; Bolick, D.T.; Medeiros, P.H.; Ledwaba, S.E.; Kolling, G.L.; Tung, K.; Guerry, P.; Swann, J.R.; Guerrant, R.L. A novel mouse model of Campylobacter jejuni enteropathy and diarrhea. PLoS Pathog. 2018, 14, e1007083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Emoto, M.; Emoto, Y.; Brinkmann, V.; Miyamoto, M.; Yoshizawa, I.; Staber, M.; van Rooijen, N.; Hamann, A.; Kaufmann, S.H. Increased resistance of LFA-1-deficient mice to lipopolysaccharide-induced shock/liver injury in the presence of TNF-alpha and IL-12 is mediated by IL-10: A novel role for LFA-1 in the regulation of the proinflammatory and anti-inflammatory cytokine balance. J. Immunol. 2003, 171, 584–593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Robertson, S.A.; Care, A.S.; Skinner, R.J. Interleukin 10 regulates inflammatory cytokine synthesis to protect against lipopolysaccharide-induced abortion and fetal growth restriction in mice. Biol. Reprod. 2007, 76, 738–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Robertson, S.A.; Skinner, R.J.; Care, A.S. Essential role for IL-10 in resistance to lipopolysaccharide-induced preterm labor in mice. J. Immunol. 2006, 177, 4888–4896. [Google Scholar] [CrossRef] [Green Version]
  162. Snyder, S.L.; Walker, R.I. Inhibition of lethality in endotoxin-challenged mice treated with zinc chloride. Infect. Immun. 1976, 13, 998–1000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ohata, S.; Moriyama, C.; Yamashita, A.; Nishida, T.; Kusumoto, C.; Mochida, S.; Minami, Y.; Nakada, J.; Shomori, K.; Inagaki, Y.; et al. Polaprezinc Protects Mice against Endotoxin Shock. J. Clin. Biochem. Nutr. 2010, 46, 234–243. [Google Scholar] [CrossRef] [Green Version]
  164. Chen, Y.H.; Zhao, M.; Chen, X.; Zhang, Y.; Wang, H.; Huang, Y.Y.; Wang, Z.; Zhang, Z.H.; Zhang, C.; Xu, D.X. Zinc supplementation during pregnancy protects against lipopolysaccharide-induced fetal growth restriction and demise through its anti-inflammatory effect. J. Immunol. 2012, 189, 454–463. [Google Scholar] [CrossRef] [Green Version]
  165. Lazzerini, M.; Wanzira, H. Oral zinc for treating diarrhoea in children. Cochrane Database Syst. Rev. 2016, 12, CD005436. [Google Scholar] [CrossRef] [Green Version]
  166. Sun, X.; Threadgill, D.; Jobin, C. Campylobacter jejuni induces colitis through activation of mammalian target of rapamycin signaling. Gastroenterology 2012, 142, 86–95.e85. [Google Scholar] [CrossRef] [Green Version]
  167. Schmidt, A.-M.; Escher, U.; Mousavi, S.; Tegtmeyer, N.; Boehm, M.; Backert, S.; Bereswill, S.; Heimesaat, M.M. Immunopathological properties of the Campylobacter jejuni flagellins and the adhesin CadF as assessed in a clinical murine infection model. Gut Pathog. 2019, 11, 24. [Google Scholar] [CrossRef] [Green Version]
  168. Mousavi, S.; Escher, U.; Thunhorst, E.; Kittler, S.; Kehrenberg, C.; Bereswill, S.; Heimesaat, M.M. Vitamin C alleviates acute enterocolitis in Campylobacter jejuni infected mice. Sci. Rep. 2020, 10, 2921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Mousavi, S.; Lobo de Sa, F.D.; Schulzke, J.D.; Bucker, R.; Bereswill, S.; Heimesaat, M.M. Vitamin D in Acute Campylobacteriosis-Results From an Intervention Study Applying a Clinical Campylobacter jejuni Induced Enterocolitis Model. Front. Immunol. 2019, 10, 2094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Du, K.; Foote, M.S.; Mousavi, S.; Buczkowski, A.; Schmidt, S.; Bereswill, S.; Heimesaat, M.M. Less Pronounced Immunopathological Responses Following Oral Butyrate Treatment of Campylobacter jejuni-Infected Mice. Microorganisms 2022, 10, 1953. [Google Scholar] [CrossRef]
  171. Bereswill, S.; Mousavi, S.; Weschka, D.; Buczkowski, A.; Schmidt, S.; Heimesaat, M.M. Peroral Clove Essential Oil Treatment Ameliorates Acute Campylobacteriosis-Results from a Preclinical Murine Intervention Study. Microorganisms 2021, 9, 735. [Google Scholar] [CrossRef]
  172. Heimesaat, M.M.; Mousavi, S.; Weschka, D.; Bereswill, S. Anti-Pathogenic and Immune-Modulatory Effects of Peroral Treatment with Cardamom Essential Oil in Acute Murine Campylobacteriosis. Microorganisms 2021, 9, 169. [Google Scholar] [CrossRef]
  173. Heimesaat, M.M.; Mousavi, S.; Weschka, D.; Bereswill, S. Garlic Essential Oil as Promising Option for the Treatment of Acute Campylobacteriosis—Results from a Preclinical Placebo-Controlled Intervention Study. Microorganisms 2021, 9, 1140. [Google Scholar] [CrossRef]
  174. Mousavi, S.; Weschka, D.; Bereswill, S.; Heimesaat, M.M. Immune-Modulatory Effects upon Oral Application of Cumin-Essential-Oil to Mice Suffering from Acute Campylobacteriosis. Pathogens 2021, 10, 818. [Google Scholar] [CrossRef]
  175. Heimesaat, M.M.; Mousavi, S.; Escher, U.; Lobo de Sá, F.D.; Peh, E.; Schulzke, J.-D.; Kittler, S.; Bücker, R.; Bereswill, S. Resveratrol Alleviates Acute Campylobacter jejuni Induced Enterocolitis in a Preclinical Murine Intervention Study. Microorganisms 2020, 8, 1858. [Google Scholar] [CrossRef]
  176. Mousavi, S.; Schmidt, A.-M.; Escher, U.; Kittler, S.; Kehrenberg, C.; Thunhorst, E.; Bereswill, S.; Heimesaat, M.M. Carvacrol ameliorates acute campylobacteriosis in a clinical murine infection model. Gut Pathog. 2020, 12, 2. [Google Scholar] [CrossRef] [Green Version]
  177. Mousavi, S.; Weschka, D.; Bereswill, S.; Heimesaat, M.M. Preclinical Evaluation of Oral Urolithin-A for the Treatment of Acute Campylobacteriosis in Campylobacter jejuni Infected Microbiota-Depleted IL-10-/-Mice. Pathogens 2021, 10, 7. [Google Scholar] [CrossRef]
  178. Bereswill, S.; Mousavi, S.; Weschka, D.; Heimesaat, M.M. Disease-Alleviating Effects of Peroral Activated Charcoal Treatment in Acute Murine Campylobacteriosis. Microorganisms 2021, 9, 1424. [Google Scholar] [CrossRef]
  179. Masanta, W.O.; Heimesaat, M.M.; Bereswill, S.; Tareen, A.M.; Lugert, R.; Groß, U.; Zautner, A.E. Modification of intestinal microbiota and its consequences for innate immune response in the pathogenesis of campylobacteriosis. Clin. Dev. Immunol. 2013, 2013, 526860. [Google Scholar] [CrossRef] [Green Version]
  180. Heimesaat, M.M.; Bereswill, S. Murine infection models for the investigation of Campylobacter jejuni--host interactions and pathogenicity. Berl. Munch. Tierarztl. Wochenschr. 2015, 128, 98–103. [Google Scholar]
  181. Bereswill, S.; Mousavi, S.; Weschka, D.; Buczkowski, A.; Schmidt, S.; Heimesaat, M.M. Iron Deprivation by Oral Deferoxamine Application Alleviates Acute Campylobacteriosis in a Clinical Murine Campylobacter jejuni Infection Model. Biomolecules 2022, 13, 71. [Google Scholar] [CrossRef]
  182. Goodfellow, J.A.; Willison, H.J. Guillain–Barré syndrome: A century of progress. Nat. Rev. Neurol. 2016, 12, 723–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Brooks, P.T.; Brakel, K.A.; Bell, J.A.; Bejcek, C.E.; Gilpin, T.; Brudvig, J.M.; Mansfield, L.S. Transplanted human fecal microbiota enhanced Guillain Barre syndrome autoantibody responses after Campylobacter jejuni infection in C57BL/6 mice. Microbiome 2017, 5, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Brooks, P.T.; Bell, J.A.; Bejcek, C.E.; Malik, A.; Mansfield, L.S. An antibiotic depleted microbiome drives severe Campylobacter jejuni-mediated Type 1/17 colitis, Type 2 autoimmunity and neurologic sequelae in a mouse model. J. Neuroimmunol. 2019, 337, 577048. [Google Scholar] [CrossRef]
  185. Adler, H.E.; DaMassa, A.J. Toxicity of endotoxin to chicks. Avian Dis. 1979, 23, 174–178. [Google Scholar] [CrossRef]
Figure 1. Overview of the molecular targets in C. jejuni treatment strategies. For more details, see text.
Figure 1. Overview of the molecular targets in C. jejuni treatment strategies. For more details, see text.
Biomolecules 13 00409 g001
Table 1. Factors influencing survival of Campylobacter from the farm to the consumer.
Table 1. Factors influencing survival of Campylobacter from the farm to the consumer.
Stage FactorsReferences
EnvironmentUV light, oxygen concentration, dehydration, temperature[81,82]
FarmBiosecurity[83]
Cleaning and disinfection[84]
Vaccination[85,86]
Bacteriophage application[87,88]
Competitive exclusion[89]
Probiotics[90]
Bacteriocins[91]
Feed supplements (e.g., carvacrol, curcumin, cinnamon oil); drinking water supplements (e.g., organic acids)[83,92,93]
SlaughterScalding temperature[94,95]
Hot steam[96,97]
Washing or rinsing[98]
Chlorinated water[99]
Sodium hypochlorite solutions[100]
High-pressure spray[101]
Cooling[102]
Food processingAdditives (e.g., organic acids, marinades, spices, sugar)[103,104,105]
Modified atmosphere packaging[106]
Chilling or freezing[107]
Bioactive packaging (immobilized zinc oxide nanoparticles, immobilized bacteriophages)[108,109]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Heimesaat, M.M.; Backert, S.; Alter, T.; Bereswill, S. Molecular Targets in Campylobacter Infections. Biomolecules 2023, 13, 409. https://doi.org/10.3390/biom13030409

AMA Style

Heimesaat MM, Backert S, Alter T, Bereswill S. Molecular Targets in Campylobacter Infections. Biomolecules. 2023; 13(3):409. https://doi.org/10.3390/biom13030409

Chicago/Turabian Style

Heimesaat, Markus M., Steffen Backert, Thomas Alter, and Stefan Bereswill. 2023. "Molecular Targets in Campylobacter Infections" Biomolecules 13, no. 3: 409. https://doi.org/10.3390/biom13030409

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