Next Article in Journal
Effect of Dietary Supplementation with a Mixture of Natural Antioxidants on Milk Yield, Composition, Oxidation Stability and Udder Health in Dairy Ewes
Next Article in Special Issue
Artifacts Introduced by Sample Handling in Chemiluminescence Assays of Nitric Oxide Metabolites
Previous Article in Journal
A Systematic Review and Meta-Analysis of the Association between Uric Acid and Allantoin and Rheumatoid Arthritis
Previous Article in Special Issue
Hydrogen Sulfide-Releasing Indomethacin-Derivative (ATB-344) Prevents the Development of Oxidative Gastric Mucosal Injuries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Hydrogen Sulfide in Inflammatory Bowel Disease

by
Nathalie Stummer
1,
René G. Feichtinger
1,
Daniel Weghuber
1,
Barbara Kofler
1,2 and
Anna M. Schneider
1,*
1
Department of Pediatrics, University Hospital of the Paracelsus Medical University, 5020 Salzburg, Austria
2
Research Program for Receptor Biochemistry and Tumor Metabolism, Paracelsus Medical University (PMU), 5020 Salzburg, Austria
*
Author to whom correspondence should be addressed.
Antioxidants 2023, 12(8), 1570; https://doi.org/10.3390/antiox12081570
Submission received: 10 July 2023 / Revised: 28 July 2023 / Accepted: 4 August 2023 / Published: 6 August 2023

Abstract

:
Hydrogen sulfide (H2S), originally known as toxic gas, has now attracted attention as one of the gasotransmitters involved in many reactions in the human body. H2S has been assumed to play a role in the pathogenesis of many chronic diseases, of which the exact pathogenesis remains unknown. One of them is inflammatory bowel disease (IBD), a chronic intestinal disease subclassified as Crohn’s disease (CD) and ulcerative colitis (UC). Any change in the amount of H2S seems to be linked to inflammation in this illness. These changes can be brought about by alterations in the microbiota, in the endogenous metabolism of H2S and in the diet. As both too little and too much H2S drive inflammation, a balanced level is needed for intestinal health. The aim of this review is to summarize the available literature published until June 2023 in order to provide an overview of the current knowledge of the connection between H2S and IBD.

Graphical Abstract

1. Introduction

In 1996, the discovery of its neuromodulating effects changed the image of hydrogen sulfide (H2S) from that of a toxic gas to that of an important messenger in cells [1,2]. Since then, many important physiologic functions of H2S have been discovered, which is why it now belongs to the family of gasotransmitters [3]. These physiologic functions range from decreasing blood pressure to controlling the nervous system [4,5,6]. Like the other gasotransmitters, nitric oxide (NO) and carbon monoxide (CO), H2S is able to cross cell membranes independently of transporters or membrane receptors [7,8].
H2S can be the result of endogenous production or a product of microbial metabolism (Figure 1). The intestinal epithelial cells are exposed to both, which is why the intestine must be efficient in regulating the H2S concentration [9]. Endogenous H2S can be produced through the desulfhydration of L-cysteine with or without homocysteine, and from 3-mercaptopyruvate (3MP), produced from cysteine and α-ketoglutarate [10,11]. These reactions are catalyzed by cystathionine-γ-lyase (CSE) and cystathionine-β-synthase (CBS), as well as 3-mercapto-sulfurtransferase (3-MST), respectively [10,12,13]. Normally, the majority of H2S is produced from cysteine and approximately a third from homocysteine [12]. While CBS and CSE seem to be the most important H2S-producing enzymes, 3-MST might be more important in the colon [14,15,16,17,18,19]. Exogenously produced H2S is the result of microbial metabolism degrading proteins into amino acids, with cysteine and other sulfur-containing compounds among them [9,20]. Several bacterial groups, like Fusobacterium, Clostridium, Escherichia, Salmonella, Klebsiella, Streptococcus, Desulfovibri, and Enterobacter, are able to metabolize cysteine and create H2S in the process [21,22,23]. In particular, cysteine desulfhydrase-containing bacteria, which belong to the Clostridium cluster, XIVa, are abundant in the group of high H2S-producing bacteria that can be found in fecal samples [24].
H2S can be removed from the body in three ways: oxidation, methylation and expiration. Methylation and expiration, however, contribute only minimally to H2S disposal [25]. H2S is oxidized by sulfide:quinone oxidoreductase (SQOR), a mitochondrial enzyme on the inner membrane [26]. Then, either ethylmalonic encephalopathy 1 protein (ETHE1, also known as persulfide dioxygenase) or thiosulfate sulfurtransferase (TST, also known as rhodanese), other mitochondrial enzymes, further metabolize H2S to produce sulfate, which can exit the body in the urine via the kidneys [27,28,29,30,31,32].
Along with its physiological functions, which are still being discovered, H2S is gaining more and more attention as an influencing factor in many pathologies [33,34,35,36]. As the intestinal tract is exposed to more H2S than most other organs, many studies are focusing on the role of H2S in intestinal diseases, particularly its role in the not-yet well known pathogenesis of inflammatory bowel disease (IBD) [9,14,15,37,38,39,40].
IBD describes two chronic intestinal diseases: ulcerative colitis (UC) and Crohn’s disease (CD). These differ in clinical, microscopic, macroscopic, and radiological features, and can affect different parts of the gastrointestinal tract [41]. However, if the inflammation pattern does not fit into any of these groups, it is classified as indeterminate colitis (IC, or IBD unclassified (IBD-U)) [42,43]. CD is characterized histopathologically as a patchy transmural inflammatory “skip lesions”, which might appear anywhere along the gastrointestinal tract and show inflamed areas next to uninflamed ones, granulomas and deep-penetrating ulcers. Complications can be fistulas, strictures, and abscesses [44,45,46]. UC is restricted to inflammatory lesions in the colonic mucosa [45]. In CD, the terminal ileum is the most likely affected location, whereas the distal rectum is mainly affected in UC [46,47]. Both CD and UC most commonly have a clinical course of recurrent flares and remission [48].
IBD affects males and females equally, with a peak incidence between 10 and 30 years and then again between 50 and 60 years [49]. The number of people affected is rising globally, particularly in newly industrialized countries. This phenomenon is amplified by the increasing rates of diagnosis and decreasing rates of mortality [50]. About 25% of cases of newly diagnosed IBD occur during childhood and adolescence, where it tends to be more aggressive [51,52]. Many studies have shown an increase in pediatric IBD incidence in the past 20 years [53,54,55,56]. CD seems to occur earlier in the patient’s life than UC [57].
The exact pathogenesis of IBD is still unknown. Currently, it is hypothesized that a combination of genetics, environmental factors, intestinal microbiota, and changes in the immune response are responsible for triggering the onset of the disease [58,59,60,61]. There is an increasing number of genes that are linked to the disease pathogenesis. However, these are only increasing the susceptibility and are only explaining around 25% of heritability [59]. Therefore, it is proposed that interactions among genes and their products are an important factor for the onset of the disease [62]. Additionally, environmental factors, like drugs, diet, smoking, stress and geographical factors, are shown to influence the onset of IBD [63]. Furthermore, changes in the gut microbiota are observed in IBD patients, even though only 20–30% of these bacteria can be cultured [64]. Not only are there changes in the composition of the microbiota, but also in the mucus layer, in which the microbiota is located [65,66]. Lastly, immunologic factors have been suspected to play a role in the pathogenesis of IBD for the longest time. Alterations in the immune response of IBD patients have been observed in innate and adaptive immunity. Interestingly, genetic alterations, environmental factors, and the microbiota are all observed to affect the immune response too [59]. H2S can affect many of these factors, as it has both pro- and anti-inflammatory effects, which is why its influence in IBD is not yet well determined [15]. The following review will focus on the crucial role of H2S in the pathogenesis of IBD.

2. Materials and Methods

This review was conducted in line with the Preferred Reporting Items for Systematic Reviews and Meta-Analysis (PRISMA) guideline [67].

2.1. Data Source/Search Strategy

Research for studies on the connection of H2S and IBD was conducted on PubMed with the use of the following search terms: H2S, hydrogen sulfide, IBD, inflammatory bowel disease, Crohn‘s disease, ulcerative colitis, chronic inflammatory disorder of the bowel, gastrointestinal tract, colonic mucosa, intestinal barrier dysfunction, SRB, sulfate-reducing bacteria, H2S metabolism, cystathionine-γ-lyase, CSE, cystathionine-β-synthase, CBS, 3-mercapto-sulfurtransferase, 3-MST, sulfide:quinone oxidoreductase, SQOR, thiosulfate sulfurtransferase, TST, and ethylmalonic encephalopathy 1 protein ETHE1. The results included studies from July 1967 to June 2023.

2.2. Study Selection

The articles listed on PubMed were first screened based on titles and again after reading the abstract. The list of references in the selected papers was then examined for other studies which were relevant for this review. All studies written in English and listed on PubMed were included in the screening, independently of the study design.
The references in this review were extracted from Endnote software version X9 (Thompson Reuters; New York, NY, USA).
The selection process is visualized in Figure 2.

3. The Connection between H2S and the Pathogenesis of IBD

3.1. Microbial H2S

In the gastrointestinal tract, the number of bacteria is higher than in any other ecosystem [68]. There they form the microbiota, among which the many functions are the degradation of otherwise undigestible food components, e.g., complex carbohydrates [69]. The result of this metabolic processes are short-chain fatty acids, which are an important energy source for colonocytes [70]. Microbial degradation of sulfur-containing amino acids, on the other hand, produces H2S [71]. Another important function of the microbiota is the defense against pathogenic and opportunistic microorganisms. An imbalance in this microbial system is linked to certain diseases, e.g., IBD, cancer and diabetes [68,72]. These diseases might also be linked to H2S [68,73]. The crucial role of bacteria in the microbiota in the pathogenesis of IBD is also seen in animal models, in which antibiotics or a germ-free environment ameliorate an inflammation of the colon [74]. In bacteria, sulfide can either be the result of degradation of sulfate-containing substances or as-/dissimilatory reduction of sulfate or sulfur [20,71,75]. Since transit time and other factors, like pH, determine which metabolizing pathways can be active, the majority of bacterial sulfide production is located in the large intestine [20]. The substrate for bacterial H2S production can either be dietary or endogenous [76,77]. In the large intestine, the endogenous colonic mucins are much more sulfated than in the small intestine, providing a great supply for H2S production and therefore an optimal environment for H2S-producing bacteria [78].
H2S can be produced by sulfate-reducing bacteria (SRB) in the colonic lumen. A change in the microbiota is seen in patients with CD and UC compared to healthy ones, e.g., a higher level of activity of SRB is suggested in the colon of UC patients [79,80]. Higher levels of H2S and SRB are found in stool samples from patients with UC [38]. Likewise, a decrease in butyrate, acetate, methylamine, and trimethylamine, all products of typical gut bacteria, in fecal extracts of IBD patients reinforces the idea of a change in the microbiota [81]. Higher level of SRB correlated with the symptom severity of IBD. Accordingly, in patients with active disease, more SRB were detected than in patients in remission [82].
In contrast to the increased numbers of SRB, the autochthonous bacterial flora seems to be reduced in UC patients [83]. The higher temperature, created by the inflammatory process, increases the growth of some SRB, which then also increases H2S production [83]. SRB produces acetate, which in turn inhibits the enzyme TST. This leads not only to a decrease in the detoxification of H2S but also to an increased permeability of H2S through the intestinal barrier [84]. This synergy of H2S and acetate intensifies the aggressiveness of H2S in the colonic lumen [83].
The microbiota of children with IBD is altered in contrast to that of healthy ones. The changes differ from the alterations seen in adults. Here, different bacterial groups predominate, like Faecalibacterium prausnitzii, which is found to be decreased in adults with IBD [85,86,87]. These studies used culture-based technology to examine the configuration of the microbiota. As not all bacteria in the microbiota can be cultured, a comprehensive molecular analysis of the pediatric microbiota in IBD patients is needed [64,88]. Nevertheless, the number of H2S-producing bacteria is found to be increased in children as well as in adults with new-onset IBD [39].
In pediatric CD patients, a decrease in butyrate, a short-chain fatty acid, which is not only an important energy source but also has anti-inflammatory properties, is seen [39,89]. Furthermore, butyrate is shown to improve the barrier function [90]. Butyrate seems to stimulate mitochondrial gene expression as well as mitochondrial H2S-detoxification enzymes [39,91]. The decrease in butyrate might be due to a decrease in butyrate-producing bacteria, which is seen in the microbiota of pediatric CD patients [39]. This decrease in butyrate—and consequently in the mitochondrial proteins that are less expressed—seems to decrease the ability of H2S-detoxification [39]. Additionally, butyrate lowers the level of epithelial oxygenation [92,93]. H2S, on the other hand, increases the level of oxygen through the inhibition of β-oxidation [94]. This combination establishes a hostile environment for obligate anaerobe bacteria, which produce short-chain fatty acids [95]. A reduction of butyrate-producing bacteria is also linked to antibiotic exposure, which in turn is suspected to increase the risk of pediatric CD [39,89,96]. The production of H2S seems to be a defense mechanism against antibiotics, which favors SRB growth in the colon and might therefore be partly responsible for the increased risk [97,98]. In addition, the lack of short-chain fatty acids increases the pH, which favors the growth of SRB even more [95].
The decrease in butyrate combined with the ability of H2S to inhibit the cytochrome c oxidase in mitochondria and β-oxidation may result in energy starvation and oxidative stress, which damages colonocytes and subsequently the gut barrier [95,99]. This allows intestinal microbes to get into direct contact with the mucosal immune system, which then gets activated, resulting in inflammation [95]. This creates a vicious cycle: the inflammation further damages the epithelial barrier, driving further activation of the immune system and resulting in more inflammation [95,100].
Another factor that links the decrease of butyrate with H2S is the availability of H2. H2 is crucial for anaerobic respiration in the intestinal lumen. One of the bacterial groups that need H2 for anaerobic respiration are SRB [101]. H2 is the product of intraluminal carbohydrate fermentation by bacteria [102]. In the event of too little production, e.g., low dietary intake of carbohydrate that can be fermented to H2, NADH is consumed as an alternative source of H2, which is used for H2S-production at the cost of butyrate production [95,103].
In stool samples from IBD patients, a higher amount of metabolites containing sulfur compared to stool from healthy individuals is seen, which might indicate a dysfunction in the metabolism of H2S [104,105]. The cause for this is most likely multifactorial, with changes in diet and therefore in the availability of substrate and alterations in the synthesis or metabolism of H2S produced from the cells themselves or bacteria in the microbiota being some of these factors [105]. Part of the increase in these sulfur-containing metabolites might originate from the increase in taurine. Taurine can result from the hydrolyzation of taurine-conjugated bile acids. However, the activity of bile salt hydrolase is seen to be decreased in active IBD [106,107]. One of the most important bile acid hydrolyzers in the gastrointestinal tract is from the Firmicutes phylum, which is reduced in IBD patients [106,108,109]. The highest activity of bile salt hydrolase is seen in Lactobacilli [110]. In patients with UC, reduced numbers of lactic acid bacteria are found [111]. Nevertheless, Kushkevych et al. showed a reduction in lactic acid bacteria levels correlating with increased levels of H2S [112]. Other bile salt hydrolyzers are also found to be decreased in IBD patients [106,113]. The increase in the availability of taurine might therefore be due to the promotion of endogenous taurine production by increased levels of methionine and cysteine from diet or from a dysfunctional cysteine or methionine metabolism as seen in IBD [105]. Taurine, when metabolized by bacteria in the gut, results in increased substrate for H2S production [105]. E. coli, which commonly metabolizes taurine, among others, is found to be increased in IBD [114,115]. The abundance of bile acids in combination with the increased availability of methionine and cysteine in IBD could therefore promote the occurrence of certain bacteria, which consequently increase the level of H2S [105].
A common drug in the treatment of IBD, 5-aminosalicylic acid (mesalamine), suppresses the growth of SRB and consequently decreases the level of produced sulfide, indicating another possible mechanism for its beneficial effects [82,116]. Additionally, mesalamine, like butyrate, stimulates the peroxisome proliferator activated receptor gamma (PPAR-γ) [117,118]. This receptor is responsible for the promotion of β-oxidation and lowers the oxygen levels of the epithelium, which in turn drives the growth of obligate anaerobes. These bacteria produce butyrate and therefore provide energy to the colonocytes [119]. However, an even more efficient decrease in inflammation and especially a decrease in nociception, which is absent with 5-aminosalicylic acid, is seen in a study with ATB-429, a H2S-releasing derivate of mesalamine [120]. Nevertheless, decreasing sulfide levels by removing the substrate for its production from SRB through a reduction in sulfur-containing amino acids ameliorates colitis as well [121,122]. An increased abundance of SRB and consequently H2S is also linked to other inflammatory diseases like periodontitis and pouchitis [98,123,124].
A decrease in the number of SRB might be one possible explanation for the beneficial effect of an appendectomy in UC [125]. In a healthy state, the appendix harbors many gut microbes, one of which is Fusobacterium spp., which is an efficient producer of H2S [126,127]. This genus of bacteria is increased in an inflamed appendix [128]. Consequently, an appendectomy and therefore the removal of a source of H2S production, may be a reason for the preventive effect of UC [95].
Nevertheless, some studies found no change in the luminal H2S of IBD patients and indicate that the bacterially produced H2S might be bound and therefore metabolically inactive [129,130]. However, the accuracy of the estimates of luminal H2S has been questioned [14]. The analysis of fecal samples is limited since it examines only the end product and therefore only the metabolizing processes at the end of the gastrointestinal tract [100]. Additionally, the colonic transition rate of feces influences the fecal H2S levels, resulting in lower accuracy of the estimates of luminal H2S [131].
In colon samples from healthy individuals, the microbiota is organized in biofilms, which form when colonies encapsule themselves in secreted polysaccharides. This supports them in the harsh living conditions they face in the gastrointestinal tract [132,133]. In control colon samples from mice and rats, these biofilms form on the sterile mucus layer. After induction of colitis, they were disorganized with bacterial translocation into the lamina propria [40]. Likewise, in healthy humans, the bacteria from the microbiota have no contact with the intestinal epithelium because of these biofilms [66,134]. In IBD patients, however, this is not the case [135,136,137]. Here, the bacteria actually do have contact because of an even thinner and more discontinuous mucus layer, which results in less epithelium covered than in healthy colons [137,138,139]. In UC patients, the mucus is even seen to be more permeable and less viscous [122,140]. This results in decreased barrier function, which is also suspected to play a significant role in the pathogenesis of IBD [141]. The changes in mucus permeability approach normal levels when IBD patients are in remission [136]. H2S might be able to reduce the linking of disulfide bonds in the mucus layer and allow bacteria to penetrate the remaining mucus layer, distorting it even more [142].
On the contrary, in animal models, it is shown that giving H2S donors does not elicit an inflammatory response but even helps to maintain a normal mucus and microbiota structure [40,143]. H2S is an important stimulant for mucus production in the colon, which promotes the establishment of a microbiota biofilm [144]. During inflammation of the colon, mucus production is reduced, which leads to fragmentation of the otherwise linear biofilms. This effect is reversed by administration of H2S donors [40]. Healing of the tissue damaged by colitis is delayed by inhibiting the synthesis of H2S and accelerated by adding H2S donors [14,19].
The beneficial or destroying effects of H2S on the mucus might therefore be concentration dependent, too, and underline the assumption of H2S having a dome-shaped curve of beneficial concentration levels in the colon, with too high and too low levels being harmful [145].

3.2. Non-Microbial and Endogenous H2S

The contribution of sulfur to the pathogenesis of IBD is supported by many animal models in which dextran sodium sulfate (DSS) damages the colonocytes, which allows for luminal bacteria to enter, consistently resulting in an inflammatory state very similar to that of IBD [146].
In the colon of rats, intraluminal exposure to NaHS, a salt that eventually forms H2S, even for a short period of time, leads to an inflammatory response. The rise in genetic expression interleukin-6 (IL6) is only seen with high concentrations of NaHS; a lower concentration does not seem to elicit an inflammatory response. In those rats as well as in human colonocytes, DNA is not harmed by NaHS. However, it seems to put colonocytes in a hypoxia-like state by inhibiting mitochondrial oxygen consumption. This inhibitory effect is seen at high concentrations of NaHS, whereas low concentrations seem to stimulate oxygen consumption. This leads to the assumption that higher luminal H2S concentrations have a negative effect on colonocytes due to the inhibition of oxygen consumption and, in turn, of the energy metabolism of this highly energy-demanding tissue, as well as due to the elicitation of inflammatory gene expression [147]. However, this seems to be a reversible effect [147,148]. The inhibition of oxygen consumption can be reduced by dietary proanthocyanidin-containing polyphenols in different fruit extracts. It is suggested that these polyphenols can bind H2S [149]. This H2S-binding characteristic is also seen with zinc chloride [150].
CSE, CBS, and 3-MST are endogenous enzymes that produce H2S [10,12,13]. SQOR, ETHE1, and TST are responsible for the detoxification process [26,27,28,29,30,31,32]. Some studies have shown differences in the enzymatic expression and the presence of enzymes involved in the metabolism of H2S in inflamed tissue and healthy controls [14,16,17,39,100,145,151,152] (Table 1). Hirata et al. demonstrated an increase in mRNA levels of CSE and CBS and, consequently, in H2S levels in the colonic mucosa of mice with DSS-induced colitis. In this animal model, the CSE seems to be the dominant producer of H2S, as CSE mRNA levels were more than 200 times higher than those of CBS [151]. De Cicco et al. showed decreased expression of CBS [16]. Wallace et al., however, concluded that the observed increase in the capacity of H2S production in rats with colitis is due to an increase in activity rather than expression, which was decreased overall [14]. In IBD as well as in in vitro inflammatory settings, 3-MST is reduced and in an animal model with complete deficiency of this enzyme, the inflammation, based on the level of inflammatory cytokines, the clinical symptoms and the histological findings, is seen to be more extensive than in the wild-type counterpart. The level of ROS and consequently intestinal epithelial cell apoptosis also correlated with 3-MST expression [153]. Interestingly, providing NaSH as an external source of H2S does not ameliorate the intestinal inflammation in 3-MST-deficient mice. This study suggests that the influence of 3-MST on AKT signaling, a regulator for epithelium inflammation, epithelium apoptosis, cell proliferation and dendritic cell maturation, and not the changes in H2S, is responsible for 3-MST-mediated colitis [153].
The detoxification of H2S also seems to be altered in IBD. Activity and expression of rhodanese are significantly reduced in DSS-induced colitis [17]. De Preter et al. also reported a decrease in TST enzyme activity and gene expression in tissue samples from UC patients [100]. This is in contrast to the study by Picton et al., which showed no difference in the activity of this enzyme in patients with IBD. The samples used, however, were taken from patients who had already been on medication and only rectal biopsies were used [152]. Other H2S-detoxification enzymes, namely ETHE1 and SQOR, are downregulated in the colon of patients with CD. This study, again, showed a repression of TST [39]. Another study, which examined all enzymes except CBS, showed a decrease in all enzymes in intestinal epithelial cells of adult IBD patients when compared to healthy adults by immunohistochemical examination of intestinal samples. Interestingly, in children with IBD, there is not such a pronounced difference in enzyme expression in epithelial cells [145].
In UC and CD patients, the gene expression, as well as the enzyme activity of the detoxification enzyme TST are reduced [100,154]. In CD patients, the reduction is less extensive after therapy with anti-tumor necrosis factor α-therapy with infliximab [154]. In these studies, a connection between inflammation and the expression of these enzymes is postulated. The authors of these studies therefore describe the changes in gene expression as a result of inflammation rather than a cause [100,154].
The reason for the contradictory data regarding the difference in enzyme expression has not been identified yet. Most studies discovered a lower expression, while only a few reported an increase in the enzyme expression [14,16,17,39,100,145,151,153].
None of the studies that showed increased expression were conducted on human samples and Wallace et al. reported in the immunohistochemical staining an increase in CSE expression in the mucosa and submucosa while the epithelial cells remained unstained [14,151]. To fully comprehend the changes in the endogenous metabolism further and more extensive studies are needed.
The crucial role of a healthy detoxification ability of the colon is confirmed in an animal study. Here, it is shown that the colonic mucosa of healthy rats absorbs up to 95% of the produced H2S. The amount of absorbed H2S would be lethal if it were to reach systemic circulation. Due to the efficient conversion of H2S to thiosulfate, the colonic tissue is not damaged by the high luminal H2S levels and the toxic amount of H2S is prevented from entering the systemic circulation [131]. The colonic mucosa is especially equipped for this task, as it detoxifies H2S much faster than other parts of the gastrointestinal tract [155].
CBS and CSE, two endogenous synthesizers of H2S, both require vitamin B6 as a cofactor [156]. This vitamin is deficient in nearly a third of IBD patients due to inadequate intake or absorption secondary to the inflammation [157]. Flannigan et al. showed that an induced vitamin B6 deficiency resulted in an impairment of H2S synthesis in the colon and worsened colitis in animal models. Consequently, a H2S donor ameliorated these effects. They even observed decreased colonic CSE expression in rats with vitamin B6 deficiency [156].
On the other hand, H2S provides a source of energy to the epithelial cells in the gastrointestinal tract as long as the concentration does not exceed certain levels [158]. H2S itself seems to have anti-inflammatory effects in low concentrations: H2S prevents leucocyte adherence in the vasculature [159] and inflammation-triggered plasma exudation [160]. H2S can also decrease pain sensation in the colon [161], and promotes healing [162] as well as resolution of colitis in preclinical animal trials [14,19,151,163]. Moreover, H2S protects mitochondria and their function in situations in which oxygen is low by upregulating the nuclear factor erythroid 2-related factor 2 (Nrf2) stress response pathway, which increases detoxifying proteins and antioxidants [164,165,166,167]. Endogenous H2S is also able to reduce inflammation by decreasing the production of pro-inflammatory cytokines and by modulating the frequency and number of granulocyte-like myeloid-derived suppressor cells [16]. Hirata et al. showed that endogenously derived H2S also acts as an antioxidant [151]. Furthermore, H2S suppresses the activation of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), an important regulator for the immune response, and even promotes the healing of gastric ulcers [168,169]. Giving H2S donors significantly alleviated colitis in rat models, whereas inhibition of H2S synthesis in those rats and healthy rats induced or worsened colitis. The mechanism behind this might be a decrease in cyclooxygenase-2 messenger RNA expression and the resulting decrease in prostaglandin synthesis [14]. In chronic inflammation, this is important for the resolution of the inflammatory response [144]. The exacerbation of inflammation after blocking the synthesis of H2S is also seen in cells with DSS-induced colitis [170].
Flannigan et al. confirmed a connection between IL-10, an anti-inflammatory cytokine, and H2S [156]. The initially observed reduction in H2S synthesis in IL-10-deficient mice was reestablished to normal levels with recombinant IL-10 [156]. H2S has a stimulatory effect on IL-10 production while decreasing pro-inflammatory cytokines such as IL-1β, IL-6, IL-8, IL-18, tumor necrosis factor-alpha (TNF-α) and interferon-gamma (IFN-γ) [120,171,172,173,174]. A link between IL-10 and IBD has already been shown in many studies, which confirm that impaired IL-10 secretion intensifies the inflammation in IBD [175,176,177]. IL-10 is crucial for mucosal homeostasis and regulation of the immune response in the colon [178,179]. Indeed, a mutation in IL-10 signaling, as seen in nucleotide-binding oligomerization domain 2 (NOD2) mutations, is linked to an increased susceptibility to IBD [175,180]. H2S thereby decreases inflammation by stimulation of IL-10. This is also observed in other organs, such as the brain, liver or lung [181,182,183]. Additionally, H2S protects the mucosa by increasing blood flow through vasodilation. This is important for decreasing mucosal damage through all erosive substances, like bile, acids and digestive enzymes, it faces and for fastening tissue repair after damage occurs [167,184,185,186].
This controversy in pro- and anti-inflammatory effect of H2S shows, again, that unbalanced levels of H2S—whether they are too high or too low—drive inflammation. Therefore, a certain range of H2S concentration level might be needed for intestinal health.

3.3. Dietary H2S

Dietary factors also seem to influence the inflammatory activity in IBD. Sulfur-containing amino acids, which can be transformed to H2S, can also be derived through diet [20,187]. Sulfate can be absorbed in the small intestine very efficiently; however, this mechanism has a saturation point [188]. Once this level is reached, the amount of sulfate reaching the colon increases with dietary intake. Nevertheless, other factors, like food preparation or meal consumption habits, can also influence the number of sulfates reaching the colon, making dietary factors difficult to study regarding their role on H2S and IBD even harder to study [95,189,190,191]. A western diet seems to be especially rich in inorganic sulfate and protein-derived sulfate [192]. A link between IBD and diet seems likely, considering the rise in IBD numbers in westernized nations. Until a decade ago, mostly Caucasian people were affected by IBD. Nevertheless, the incidence has been rising recently in the Asian and Hispanic populations, especially in those who immigrated into high-prevalence countries and particularly in their children [193]. Additionally, residents of urban centers are more at risk of developing IBD than those in rural settings [49]. Studies on the link between H2S and IBD are limited. Most of them focus on lowering sulfur intake by transitioning to a more plant-based diet, which seemed to ameliorate the disease activity. However, these studies are either case reports or have relatively low patient numbers [121,194,195,196,197].
Sulfate can be found in high amounts in many food additives, dried fruit, nuts, some vegetables, wheat bread, sausages, milk products, beer, canned and pickled products [68,198]. Dietary habits also have an influence on the microbiota. A diet high in protein is linked to increased numbers of SRB, which increases H2S production and decreases the number of butyrate producers and therefore the amount of butyrate in the colon [199,200]. These changes are also observed in patients with IBD [200,201,202,203]. A diet providing a high amount of fermentable carbohydrates leads to increased butyrate production, which subsequently lowers the pH [204,205]. A more acidic environment favors the growth of other bacterial groups, like butyrate producers, over SRB [95,204,206], while H2S production works best in an alkaline environment [207]. In vegans and vegetarians, a lower stool pH is found in comparison to omnivores, consistent with a more even distribution of short-chain fatty acid up to the distal end of the colon [95,208]. When feeding mice a diet with a high fat content, a reduced production of short-chain fatty acid and an increase in H2S production is observed [209]. In IL-10 knockout mice, which show a similar colitis to that of IBD, a diet containing a lot of saturated fat increases inflammation severely and leads, in combination with a mutagen, to the formation of adenomas [95,209,210]. All in all, a typical western animal-based diet, consisting of a high fat and protein intake, might create the perfect environment for H2S production at the expense of butyrate production, which subsequently drives inflammation [95]. Accordingly, many IBD patients believe that their symptomatic relief relies more on their diet than their medication [211].
This assumption is partly supported by the success of exclusive enteral nutrition (EEN) in the treatment of CD. EEN describes a strict liquid diet consisting of formula for a duration of 4 to 12 weeks [212,213]. Especially with pediatric IBD patients, EEN is shown to be as efficient as therapies with corticosteroids and to achieve remission in nearly 80% [212,214,215]. The reason for the reduction in inflammation with EEN is still mostly unknown [216]. One of its mechanisms of action could be alterations in the microbiota [217,218]. It was reported that EEN reduced the number of a very potent H2S producer, Atopobium parvulum [39,219]. Unfortunately, after cessation of EEN, the changes in the composition of the microbiota return to their prior state, concurrently with rising fecal sulfide levels [199,220,221].

4. Conclusions

The pressing need for a deeper understanding of the pathogenesis of IBD is growing with the rising numbers of people affected by this chronic disease [222]. As suspected in many other diseases, H2S is believed to play a crucial role in IBD [33,34,35,36]. As the intestinal tract is exposed to microbial, endogenous and dietary H2S and therefore, to more than most other organ systems, a change in the level of H2S is hypothesized to influence the health of the intestinal system [9,14,15,37,38,39,40]. Changes in the microbiota as seen in IBD, e.g., an increase in SRB and consequently a decrease in butyrate-producing bacteria, the decrease of H2S-detoxifying enzymes in the intestinal epithelial cells and the increasing intake of dietary sulfate by eating a typical western animal-based diet, indicate that too much H2S has pro-inflammatory effects in IBD. However, the restorative effects of H2S on the mucus barrier and microbiota biofilms, its decreasing effects on pro-inflammatory cytokines and its overall healing-promoting characteristics also suggest that too little H2S can be pro-inflammatory. Therefore, it is likely that there is a dome-shaped curve of beneficial concentration levels in the intestinal tract. Even though it would be essential to measure intraluminal intestinal H2S levels directly to determine if H2S could be a possible target in the treatment of IBD, all these studies definitely underline the connection between H2S and the pathogenesis of IBD.

Author Contributions

Conceptualization, N.S., A.M.S. and D.W.; methodology N.S. and A.M.S., investigation, N.S.; resources, N.S. and D.W.; data curation, N.S.; writing—original draft preparation, N.S.; writing—review and editing, N.S., A.M.S., B.K., R.G.F. and D.W; visualization, N.S.; supervision, A.M.S., B.K., R.G.F. and D.W.; project administration, N.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. Evans, C.L. The toxicity of hydrogen sulphide and other sulphides. Q. J. Exp. Physiol. Cogn. Med. Sci. 1967, 52, 231–248. [Google Scholar] [CrossRef] [PubMed]
  2. Abe, K.; Kimura, H. The possible role of hydrogen sulfide as an endogenous neuromodulator. J. Neurosci. 1996, 16, 1066–1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wang, R. The gasotransmitter role of hydrogen sulfide. Antioxid. Redox Signal 2003, 5, 493–501. [Google Scholar] [CrossRef]
  4. Yang, G.; Wu, L.; Jiang, B.; Yang, W.; Qi, J.; Cao, K.; Meng, Q.; Mustafa, A.K.; Mu, W.; Zhang, S.; et al. H2S as a physiologic vasorelaxant: Hypertension in mice with deletion of cystathionine gamma-lyase. Science 2008, 322, 587–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Paul, B.D.; Sbodio, J.I.; Xu, R.; Vandiver, M.S.; Cha, J.Y.; Snowman, A.M.; Snyder, S.H. Cystathionine γ-lyase deficiency mediates neurodegeneration in Huntington’s disease. Nature 2014, 509, 96–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hine, C.; Mitchell, J.R. Calorie restriction and methionine restriction in control of endogenous hydrogen sulfide production by the transsulfuration pathway. Exp. Gerontol. 2015, 68, 26–32. [Google Scholar] [CrossRef] [Green Version]
  7. Wang, R. Two’s company, three’s a crowd: Can H2S be the third endogenous gaseous transmitter? FASEB J. 2002, 16, 1792–1798. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, R. Gasotransmitters: Growing pains and joys. Trends Biochem. Sci. 2014, 39, 227–232. [Google Scholar] [CrossRef]
  9. Blachier, F.; Beaumont, M.; Kim, E. Cysteine-derived hydrogen sulfide and gut health: A matter of endogenous or bacterial origin. Curr. Opin. Clin. Nutr. Metab. Care 2019, 22, 68–75. [Google Scholar] [CrossRef]
  10. Kabil, O.; Banerjee, R. Redox biochemistry of hydrogen sulfide. J. Biol. Chem. 2010, 285, 21903–21907. [Google Scholar] [CrossRef] [Green Version]
  11. Kimura, H. Signaling molecules: Hydrogen sulfide and polysulfide. Antioxid. Redox Signal 2015, 22, 362–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chiku, T.; Padovani, D.; Zhu, W.; Singh, S.; Vitvitsky, V.; Banerjee, R. H2S biogenesis by human cystathionine gamma-lyase leads to the novel sulfur metabolites lanthionine and homolanthionine and is responsive to the grade of hyperhomocysteinemia. J. Biol. Chem. 2009, 284, 11601–11612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Singh, S.; Padovani, D.; Leslie, R.A.; Chiku, T.; Banerjee, R. Relative contributions of cystathionine beta-synthase and gamma-cystathionase to H2S biogenesis via alternative trans-sulfuration reactions. J. Biol. Chem. 2009, 284, 22457–22466. [Google Scholar] [CrossRef] [Green Version]
  14. Wallace, J.L.; Vong, L.; McKnight, W.; Dicay, M.; Martin, G.R. Endogenous and exogenous hydrogen sulfide promotes resolution of colitis in rats. Gastroenterology 2009, 137, 569–578, 578.e1. [Google Scholar] [CrossRef] [Green Version]
  15. Guo, F.F.; Yu, T.C.; Hong, J.; Fang, J.Y. Emerging Roles of Hydrogen Sulfide in Inflammatory and Neoplastic Colonic Diseases. Front. Physiol. 2016, 7, 156. [Google Scholar] [CrossRef] [Green Version]
  16. De Cicco, P.; Sanders, T.; Cirino, G.; Maloy, K.J.; Ianaro, A. Hydrogen Sulfide Reduces Myeloid-Derived Suppressor Cell-Mediated Inflammatory Response in a Model of Helicobacter hepaticus-Induced Colitis. Front. Immunol. 2018, 9, 499. [Google Scholar] [CrossRef] [Green Version]
  17. Taniguchi, E.; Matsunami, M.; Kimura, T.; Yonezawa, D.; Ishiki, T.; Sekiguchi, F.; Nishikawa, H.; Maeda, Y.; Ishikura, H.; Kawabata, A. Rhodanese, but not cystathionine-gamma-lyase, is associated with dextran sulfate sodium-evoked colitis in mice: A sign of impaired colonic sulfide detoxification? Toxicology 2009, 264, 96–103. [Google Scholar] [CrossRef] [PubMed]
  18. Linden, D.R.; Sha, L.; Mazzone, A.; Stoltz, G.J.; Bernard, C.E.; Furne, J.K.; Levitt, M.D.; Farrugia, G.; Szurszewski, J.H. Production of the gaseous signal molecule hydrogen sulfide in mouse tissues. J. Neurochem. 2008, 106, 1577–1585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Flannigan, K.L.; Ferraz, J.G.; Wang, R.; Wallace, J.L. Enhanced synthesis and diminished degradation of hydrogen sulfide in experimental colitis: A site-specific, pro-resolution mechanism. PLoS ONE 2013, 8, e71962. [Google Scholar] [CrossRef] [Green Version]
  20. Blachier, F.; Davila, A.M.; Mimoun, S.; Benetti, P.H.; Atanasiu, C.; Andriamihaja, M.; Benamouzig, R.; Bouillaud, F.; Tomé, D. Luminal sulfide and large intestine mucosa: Friend or foe? Amino Acids 2010, 39, 335–347. [Google Scholar] [CrossRef]
  21. Barton, L.L.; Ritz, N.L.; Fauque, G.D.; Lin, H.C. Sulfur Cycling and the Intestinal Microbiome. Dig. Dis. Sci. 2017, 62, 2241–2257. [Google Scholar] [CrossRef]
  22. Hale, V.L.; Jeraldo, P.; Mundy, M.; Yao, J.; Keeney, G.; Scott, N.; Cheek, E.H.; Davidson, J.; Greene, M.; Martinez, C.; et al. Synthesis of multi-omic data and community metabolic models reveals insights into the role of hydrogen sulfide in colon cancer. Methods 2018, 149, 59–68. [Google Scholar] [CrossRef]
  23. Tomasova, L.; Konopelski, P.; Ufnal, M. Gut Bacteria and Hydrogen Sulfide: The New Old Players in Circulatory System Homeostasis. Molecules 2016, 21, 1558. [Google Scholar] [CrossRef] [Green Version]
  24. Feng, Y.; Stams, A.J.M.; de Vos, W.M.; Sánchez-Andrea, I. Enrichment of sulfidogenic bacteria from the human intestinal tract. FEMS Microbiol. Lett. 2017, 364, fnx028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Cao, X.; Ding, L.; Xie, Z.Z.; Yang, Y.; Whiteman, M.; Moore, P.K.; Bian, J.S. A Review of Hydrogen Sulfide Synthesis, Metabolism, and Measurement: Is Modulation of Hydrogen Sulfide a Novel Therapeutic for Cancer? Antioxid. Redox Signal 2019, 31, 1–38. [Google Scholar] [CrossRef] [Green Version]
  26. Augustyn, K.D.; Jackson, M.R.; Jorns, M.S. Use of Tissue Metabolite Analysis and Enzyme Kinetics to Discriminate between Alternate Pathways for Hydrogen Sulfide Metabolism. Biochemistry 2017, 56, 986–996. [Google Scholar] [CrossRef]
  27. Libiad, M.; Sriraman, A.; Banerjee, R. Polymorphic Variants of Human Rhodanese Exhibit Differences in Thermal Stability and Sulfur Transfer Kinetics. J. Biol. Chem. 2015, 290, 23579–23588. [Google Scholar] [CrossRef] [Green Version]
  28. Jackson, M.R.; Melideo, S.L.; Jorns, M.S. Human sulfide: Quinone oxidoreductase catalyzes the first step in hydrogen sulfide metabolism and produces a sulfane sulfur metabolite. Biochemistry 2012, 51, 6804–6815. [Google Scholar] [CrossRef]
  29. Jackson, M.R.; Melideo, S.L.; Jorns, M.S. Role of human sulfide: Quinone oxidoreductase in H2S metabolism. Methods Enzymol. 2015, 554, 255–270. [Google Scholar] [CrossRef]
  30. Kabil, O.; Motl, N.; Strack, M.; Seravalli, J.; Metzler-Nolte, N.; Banerjee, R. Mechanism-based inhibition of human persulfide dioxygenase by γ-glutamyl-homocysteinyl-glycine. J. Biol. Chem. 2018, 293, 12429–12439. [Google Scholar] [CrossRef] [Green Version]
  31. Kabil, O.; Banerjee, R. Characterization of patient mutations in human persulfide dioxygenase (ETHE1) involved in H2S catabolism. J. Biol. Chem. 2012, 287, 44561–44567. [Google Scholar] [CrossRef] [Green Version]
  32. Pettinati, I.; Brem, J.; McDonough, M.A.; Schofield, C.J. Crystal structure of human persulfide dioxygenase: Structural basis of ethylmalonic encephalopathy. Hum. Mol. Genet. 2015, 24, 2458–2469. [Google Scholar] [CrossRef] [Green Version]
  33. Zhen, Y.; Pan, W.; Hu, F.; Wu, H.; Feng, J.; Zhang, Y.; Chen, J. Exogenous hydrogen sulfide exerts proliferation/anti-apoptosis/angiogenesis/migration effects via amplifying the activation of NF-κB pathway in PLC/PRF/5 hepatoma cells. Int. J. Oncol. 2015, 46, 2194–2204. [Google Scholar] [CrossRef] [Green Version]
  34. Zhen, Y.; Zhang, W.; Liu, C.; He, J.; Lu, Y.; Guo, R.; Feng, J.; Zhang, Y.; Chen, J. Exogenous hydrogen sulfide promotes C6 glioma cell growth through activation of the p38 MAPK/ERK1/2-COX-2 pathways. Oncol. Rep. 2015, 34, 2413–2422. [Google Scholar] [CrossRef] [Green Version]
  35. Lei, Y.; Zhen, Y.; Zhang, W.; Sun, X.; Lin, X.; Feng, J.; Luo, H.; Chen, Z.; Su, C.; Zeng, B.; et al. Exogenous hydrogen sulfide exerts proliferation, anti-apoptosis, angiopoiesis and migration effects via activating HSP90 pathway in EC109 cells. Oncol. Rep. 2016, 35, 3714–3720. [Google Scholar] [CrossRef] [Green Version]
  36. Yang, N.; Liu, Y.; Li, T.; Tuo, Q. Role of Hydrogen Sulfide in Chronic Diseases. DNA Cell Biol. 2020, 39, 187–196. [Google Scholar] [CrossRef]
  37. Blachier, F.; Andriamihaja, M.; Larraufie, P.; Ahn, E.; Lan, A.; Kim, E. Production of hydrogen sulfide by the intestinal microbiota and epithelial cells and consequences for the colonic and rectal mucosa. Am. J. Physiol. Gastrointest. Liver Physiol. 2021, 320, G125–G135. [Google Scholar] [CrossRef]
  38. Medani, M.; Collins, D.; Docherty, N.G.; Baird, A.W.; O’Connell, P.R.; Winter, D.C. Emerging role of hydrogen sulfide in colonic physiology and pathophysiology. Inflamm. Bowel Dis. 2011, 17, 1620–1625. [Google Scholar] [CrossRef]
  39. Mottawea, W.; Chiang, C.K.; Mühlbauer, M.; Starr, A.E.; Butcher, J.; Abujamel, T.; Deeke, S.A.; Brandel, A.; Zhou, H.; Shokralla, S.; et al. Altered intestinal microbiota-host mitochondria crosstalk in new onset Crohn’s disease. Nat. Commun. 2016, 7, 13419. [Google Scholar] [CrossRef] [Green Version]
  40. Motta, J.P.; Flannigan, K.L.; Agbor, T.A.; Beatty, J.K.; Blackler, R.W.; Workentine, M.L.; Da Silva, G.J.; Wang, R.; Buret, A.G.; Wallace, J.L. Hydrogen sulfide protects from colitis and restores intestinal microbiota biofilm and mucus production. Inflamm. Bowel Dis. 2015, 21, 1006–1017. [Google Scholar] [CrossRef]
  41. Podolsky, D.K. Inflammatory bowel disease. N. Engl. J. Med. 2002, 347, 417–429. [Google Scholar] [CrossRef]
  42. Geboes, K.; Van Eyken, P. Inflammatory bowel disease unclassified and indeterminate colitis: The role of the pathologist. J. Clin. Pathol. 2009, 62, 201–205. [Google Scholar] [CrossRef]
  43. Tremaine, W.J. Diagnosis and treatment of indeterminate colitis. Gastroenterol. Hepatol. 2011, 7, 826–828. [Google Scholar]
  44. Baumgart, D.C.; Sandborn, W.J. Crohn’s disease. Lancet 2012, 380, 1590–1605. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, T.C.; Stappenbeck, T.S. Genetics and Pathogenesis of Inflammatory Bowel Disease. Annu. Rev. Pathol. 2016, 11, 127–148. [Google Scholar] [CrossRef] [Green Version]
  46. Torres, J.; Mehandru, S.; Colombel, J.F.; Peyrin-Biroulet, L. Crohn’s disease. Lancet 2017, 389, 1741–1755. [Google Scholar] [CrossRef]
  47. Ordás, I.; Eckmann, L.; Talamini, M.; Baumgart, D.C.; Sandborn, W.J. Ulcerative colitis. Lancet 2012, 380, 1606–1619. [Google Scholar] [CrossRef] [Green Version]
  48. Sairenji, T.; Collins, K.L.; Evans, D.V. An Update on Inflammatory Bowel Disease. Prim. Care 2017, 44, 673–692. [Google Scholar] [CrossRef]
  49. Molodecky, N.A.; Soon, I.S.; Rabi, D.M.; Ghali, W.A.; Ferris, M.; Chernoff, G.; Benchimol, E.I.; Panaccione, R.; Ghosh, S.; Barkema, H.W.; et al. Increasing incidence and prevalence of the inflammatory bowel diseases with time, based on systematic review. Gastroenterology 2012, 142, 46–54.e42; quiz e30. [Google Scholar] [CrossRef] [Green Version]
  50. Chaparro, M.; Garre, A.; Núñez Ortiz, A.; Diz-Lois Palomares, M.T.; Rodríguez, C.; Riestra, S.; Vela, M.; Benítez, J.M.; Fernández Salgado, E.; Sánchez Rodríguez, E.; et al. Incidence, Clinical Characteristics and Management of Inflammatory Bowel Disease in Spain: Large-Scale Epidemiological Study. J. Clin. Med. 2021, 10, 2885. [Google Scholar] [CrossRef]
  51. Sauer, C.G.; Kugathasan, S. Pediatric inflammatory bowel disease: Highlighting pediatric differences in IBD. Med. Clin. N. Am. 2010, 94, 35–52. [Google Scholar] [CrossRef] [PubMed]
  52. Van Limbergen, J.; Russell, R.K.; Drummond, H.E.; Aldhous, M.C.; Round, N.K.; Nimmo, E.R.; Smith, L.; Gillett, P.M.; McGrogan, P.; Weaver, L.T.; et al. Definition of phenotypic characteristics of childhood-onset inflammatory bowel disease. Gastroenterology 2008, 135, 1114–1122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Benchimol, E.I.; Fortinsky, K.J.; Gozdyra, P.; Van den Heuvel, M.; Van Limbergen, J.; Griffiths, A.M. Epidemiology of pediatric inflammatory bowel disease: A systematic review of international trends. Inflamm. Bowel Dis. 2011, 17, 423–439. [Google Scholar] [CrossRef]
  54. Sýkora, J.; Pomahačová, R.; Kreslová, M.; Cvalínová, D.; Štych, P.; Schwarz, J. Current global trends in the incidence of pediatric-onset inflammatory bowel disease. World J. Gastroenterol. 2018, 24, 2741–2763. [Google Scholar] [CrossRef]
  55. Ghione, S.; Sarter, H.; Fumery, M.; Armengol-Debeir, L.; Savoye, G.; Ley, D.; Spyckerelle, C.; Pariente, B.; Peyrin-Biroulet, L.; Turck, D.; et al. Dramatic Increase in Incidence of Ulcerative Colitis and Crohn’s Disease (1988–2011): A Population-Based Study of French Adolescents. Am. J. Gastroenterol. 2018, 113, 265–272. [Google Scholar] [CrossRef]
  56. Jakobsen, C.; Paerregaard, A.; Munkholm, P.; Faerk, J.; Lange, A.; Andersen, J.; Jakobsen, M.; Kramer, I.; Czernia-Mazurkiewicz, J.; Wewer, V. Pediatric inflammatory bowel disease: Increasing incidence, decreasing surgery rate, and compromised nutritional status: A prospective population-based cohort study 2007–2009. Inflamm. Bowel Dis. 2011, 17, 2541–2550. [Google Scholar] [CrossRef]
  57. Vatn, M.H.; Sandvik, A.K. Inflammatory bowel disease. Scand. J. Gastroenterol. 2015, 50, 748–762. [Google Scholar] [CrossRef]
  58. Guan, Q. A Comprehensive Review and Update on the Pathogenesis of Inflammatory Bowel Disease. J. Immunol. Res. 2019, 2019, 7247238. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, Y.Z.; Li, Y.Y. Inflammatory bowel disease: Pathogenesis. World J. Gastroenterol. 2014, 20, 91–99. [Google Scholar] [CrossRef]
  60. Dixon, L.J.; Kabi, A.; Nickerson, K.P.; McDonald, C. Combinatorial effects of diet and genetics on inflammatory bowel disease pathogenesis. Inflamm. Bowel Dis. 2015, 21, 912–922. [Google Scholar] [CrossRef]
  61. Leone, V.; Chang, E.B.; Devkota, S. Diet, microbes, and host genetics: The perfect storm in inflammatory bowel diseases. J. Gastroenterol. 2013, 48, 315–321. [Google Scholar] [CrossRef] [Green Version]
  62. Zuk, O.; Hechter, E.; Sunyaev, S.R.; Lander, E.S. The mystery of missing heritability: Genetic interactions create phantom heritability. Proc. Natl. Acad. Sci. USA 2012, 109, 1193–1198. [Google Scholar] [CrossRef] [PubMed]
  63. Loftus, E.V. Clinical epidemiology of inflammatory bowel disease: Incidence, prevalence, and environmental influences. Gastroenterology 2004, 126, 1504–1517. [Google Scholar] [CrossRef] [PubMed]
  64. Eckburg, P.B.; Bik, E.M.; Bernstein, C.N.; Purdom, E.; Dethlefsen, L.; Sargent, M.; Gill, S.R.; Nelson, K.E.; Relman, D.A. Diversity of the human intestinal microbial flora. Science 2005, 308, 1635–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ott, S.J.; Musfeldt, M.; Wenderoth, D.F.; Hampe, J.; Brant, O.; Fölsch, U.R.; Timmis, K.N.; Schreiber, S. Reduction in diversity of the colonic mucosa associated bacterial microflora in patients with active inflammatory bowel disease. Gut 2004, 53, 685–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Johansson, M.E.; Phillipson, M.; Petersson, J.; Velcich, A.; Holm, L.; Hansson, G.C. The inner of the two Muc2 mucin-dependent mucus layers in colon is devoid of bacteria. Proc. Natl. Acad. Sci. USA 2008, 105, 15064–15069. [Google Scholar] [CrossRef] [PubMed]
  67. Liberati, A.; Altman, D.G.; Tetzlaff, J.; Mulrow, C.; Gotzsche, P.C.; Ioannidis, J.P.; Clarke, M.; Devereaux, P.J.; Kleijnen, J.; Moher, D. The PRISMA statement for reporting systematic reviews and meta-analyses of studies that evaluate healthcare interventions: Explanation and elaboration. BMJ 2009, 339, b2700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Dordević, D.; Jančíková, S.; Vítězová, M.; Kushkevych, I. Hydrogen sulfide toxicity in the gut environment: Meta-analysis of sulfate-reducing and lactic acid bacteria in inflammatory processes. J. Adv. Res. 2021, 27, 55–69. [Google Scholar] [CrossRef]
  69. Tremaroli, V.; Bäckhed, F. Functional interactions between the gut microbiota and host metabolism. Nature 2012, 489, 242–249. [Google Scholar] [CrossRef] [PubMed]
  70. Donohoe, D.R.; Garge, N.; Zhang, X.; Sun, W.; O’Connell, T.M.; Bunger, M.K.; Bultman, S.J. The microbiome and butyrate regulate energy metabolism and autophagy in the mammalian colon. Cell Metab. 2011, 13, 517–526. [Google Scholar] [CrossRef] [Green Version]
  71. Awano, N.; Wada, M.; Mori, H.; Nakamori, S.; Takagi, H. Identification and functional analysis of Escherichia coli cysteine desulfhydrases. Appl. Environ. Microbiol. 2005, 71, 4149–4152. [Google Scholar] [CrossRef] [Green Version]
  72. Rolfe, R.D. The role of probiotic cultures in the control of gastrointestinal health. J. Nutr. 2000, 130, 396S–402S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zhu, L.; Yang, B.; Ma, D.; Wang, L.; Duan, W. Hydrogen Sulfide, Adipose Tissue and Diabetes Mellitus. Diabetes Metab. Syndr. Obes. 2020, 13, 1873–1886. [Google Scholar] [CrossRef] [PubMed]
  74. Elson, C.O.; Sartor, R.B.; Tennyson, G.S.; Riddell, R.H. Experimental models of inflammatory bowel disease. Gastroenterology 1995, 109, 1344–1367. [Google Scholar] [CrossRef]
  75. Smith, E.A.; Macfarlane, G.T. Dissimilatory amino Acid metabolism in human colonic bacteria. Anaerobe 1997, 3, 327–337. [Google Scholar] [CrossRef] [PubMed]
  76. Florin, T.; Neale, G.; Gibson, G.R.; Christl, S.U.; Cummings, J.H. Metabolism of dietary sulphate: Absorption and excretion in humans. Gut 1991, 32, 766–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Gibson, G.R.; Cummings, J.H.; Macfarlane, G.T. Competition for hydrogen between sulphate-reducing bacteria and methanogenic bacteria from the human large intestine. J. Appl. Bacteriol. 1988, 65, 241–247. [Google Scholar] [CrossRef]
  78. Liau, Y.H.; Horowitz, M.I. The importance of PAPS in determining sulfation in gastrointestinal mucosa. Digestion 1976, 14, 372–375. [Google Scholar] [CrossRef]
  79. Sartor, R.B. Microbial influences in inflammatory bowel diseases. Gastroenterology 2008, 134, 577–594. [Google Scholar] [CrossRef]
  80. Levine, J.; Ellis, C.J.; Furne, J.K.; Springfield, J.; Levitt, M.D. Fecal hydrogen sulfide production in ulcerative colitis. Am. J. Gastroenterol. 1998, 93, 83–87. [Google Scholar] [CrossRef]
  81. Marchesi, J.R.; Holmes, E.; Khan, F.; Kochhar, S.; Scanlan, P.; Shanahan, F.; Wilson, I.D.; Wang, Y. Rapid and noninvasive metabonomic characterization of inflammatory bowel disease. J. Proteome Res. 2007, 6, 546–551. [Google Scholar] [CrossRef]
  82. Pitcher, M.C.; Beatty, E.R.; Cummings, J.H. The contribution of sulphate reducing bacteria and 5-aminosalicylic acid to faecal sulphide in patients with ulcerative colitis. Gut 2000, 46, 64–72. [Google Scholar] [CrossRef] [Green Version]
  83. Kushkevych, I.; Dordević, D.; Vítězová, M. Possible synergy effect of hydrogen sulfide and acetate produced by sulfate-reducing bacteria on inflammatory bowel disease development. J. Adv. Res. 2021, 27, 71–78. [Google Scholar] [CrossRef]
  84. Rowan, F.E.; Docherty, N.G.; Coffey, J.C.; O’Connell, P.R. Sulphate-reducing bacteria and hydrogen sulphide in the aetiology of ulcerative colitis. Br. J. Surg. 2009, 96, 151–158. [Google Scholar] [CrossRef] [PubMed]
  85. Conte, M.P.; Schippa, S.; Zamboni, I.; Penta, M.; Chiarini, F.; Seganti, L.; Osborn, J.; Falconieri, P.; Borrelli, O.; Cucchiara, S. Gut-associated bacterial microbiota in paediatric patients with inflammatory bowel disease. Gut 2006, 55, 1760–1767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Schwiertz, A.; Jacobi, M.; Frick, J.S.; Richter, M.; Rusch, K.; Köhler, H. Microbiota in pediatric inflammatory bowel disease. J. Pediatr. 2010, 157, 240–244.e241. [Google Scholar] [CrossRef] [PubMed]
  87. Sokol, H.; Pigneur, B.; Watterlot, L.; Lakhdari, O.; Bermúdez-Humarán, L.G.; Gratadoux, J.J.; Blugeon, S.; Bridonneau, C.; Furet, J.P.; Corthier, G.; et al. Faecalibacterium prausnitzii is an anti-inflammatory commensal bacterium identified by gut microbiota analysis of Crohn disease patients. Proc. Natl. Acad. Sci. USA 2008, 105, 16731–16736. [Google Scholar] [CrossRef]
  88. Sultan, S.; El-Mowafy, M.; Elgaml, A.; Ahmed, T.A.E.; Hassan, H.; Mottawea, W. Metabolic Influences of Gut Microbiota Dysbiosis on Inflammatory Bowel Disease. Front. Physiol. 2021, 12, 715506. [Google Scholar] [CrossRef]
  89. Gevers, D.; Kugathasan, S.; Denson, L.A.; Vázquez-Baeza, Y.; Van Treuren, W.; Ren, B.; Schwager, E.; Knights, D.; Song, S.J.; Yassour, M.; et al. The treatment-naive microbiome in new-onset Crohn’s disease. Cell Host Microbe 2014, 15, 382–392. [Google Scholar] [CrossRef] [Green Version]
  90. Finnie, I.A.; Dwarakanath, A.D.; Taylor, B.A.; Rhodes, J.M. Colonic mucin synthesis is increased by sodium butyrate. Gut 1995, 36, 93–99. [Google Scholar] [CrossRef] [Green Version]
  91. Ramasamy, S.; Singh, S.; Taniere, P.; Langman, M.J.; Eggo, M.C. Sulfide-detoxifying enzymes in the human colon are decreased in cancer and upregulated in differentiation. Am. J. Physiol. Gastrointest. Liver Physiol. 2006, 291, G288–G296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Byndloss, M.X.; Olsan, E.E.; Rivera-Chávez, F.; Tiffany, C.R.; Cevallos, S.A.; Lokken, K.L.; Torres, T.P.; Byndloss, A.J.; Faber, F.; Gao, Y.; et al. Microbiota-activated PPAR-γ signaling inhibits dysbiotic Enterobacteriaceae expansion. Science 2017, 357, 570–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Litvak, Y.; Byndloss, M.X.; Tsolis, R.M.; Bäumler, A.J. Dysbiotic Proteobacteria expansion: A microbial signature of epithelial dysfunction. Curr. Opin. Microbiol. 2017, 39, 1–6. [Google Scholar] [CrossRef]
  94. Babidge, W.; Millard, S.; Roediger, W. Sulfides impair short chain fatty acid beta-oxidation at acyl-CoA dehydrogenase level in colonocytes: Implications for ulcerative colitis. Mol. Cell. Biochem. 1998, 181, 117–124. [Google Scholar] [CrossRef] [PubMed]
  95. Teigen, L.M.; Geng, Z.; Sadowsky, M.J.; Vaughn, B.P.; Hamilton, M.J.; Khoruts, A. Dietary Factors in Sulfur Metabolism and Pathogenesis of Ulcerative Colitis. Nutrients 2019, 11, 931. [Google Scholar] [CrossRef] [Green Version]
  96. Ungaro, R.; Bernstein, C.N.; Gearry, R.; Hviid, A.; Kolho, K.L.; Kronman, M.P.; Shaw, S.; Van Kruiningen, H.; Colombel, J.F.; Atreja, A. Antibiotics associated with increased risk of new-onset Crohn’s disease but not ulcerative colitis: A meta-analysis. Am. J. Gastroenterol. 2014, 109, 1728–1738. [Google Scholar] [CrossRef]
  97. Ohge, H.; Furne, J.K.; Springfield, J.; Sueda, T.; Madoff, R.D.; Levitt, M.D. The effect of antibiotics and bismuth on fecal hydrogen sulfide and sulfate-reducing bacteria in the rat. FEMS Microbiol. Lett. 2003, 228, 137–142. [Google Scholar] [CrossRef] [Green Version]
  98. Singh, S.B.; Lin, H.C. Hydrogen Sulfide in Physiology and Diseases of the Digestive Tract. Microorganisms 2015, 3, 866–889. [Google Scholar] [CrossRef] [Green Version]
  99. Jiang, J.; Chan, A.; Ali, S.; Saha, A.; Haushalter, K.J.; Lam, W.L.; Glasheen, M.; Parker, J.; Brenner, M.; Mahon, S.B.; et al. Hydrogen Sulfide—Mechanisms of Toxicity and Development of an Antidote. Sci. Rep. 2016, 6, 20831. [Google Scholar] [CrossRef]
  100. De Preter, V.; Arijs, I.; Windey, K.; Vanhove, W.; Vermeire, S.; Schuit, F.; Rutgeerts, P.; Verbeke, K. Decreased mucosal sulfide detoxification is related to an impaired butyrate oxidation in ulcerative colitis. Inflamm. Bowel Dis. 2012, 18, 2371–2380. [Google Scholar] [CrossRef]
  101. Wolf, P.G.; Biswas, A.; Morales, S.E.; Greening, C.; Gaskins, H.R. H2 metabolism is widespread and diverse among human colonic microbes. Gut Microbes 2016, 7, 235–245. [Google Scholar] [CrossRef] [Green Version]
  102. Carbonero, F.; Benefiel, A.C.; Gaskins, H.R. Contributions of the microbial hydrogen economy to colonic homeostasis. Nat. Rev. Gastroenterol. Hepatol. 2012, 9, 504–518. [Google Scholar] [CrossRef] [PubMed]
  103. Macfarlane, S.; Macfarlane, G.T. Regulation of short-chain fatty acid production. Proc. Nutr. Soc. 2003, 62, 67–72. [Google Scholar] [CrossRef] [PubMed]
  104. Duboc, H.; Rajca, S.; Rainteau, D.; Benarous, D.; Maubert, M.A.; Quervain, E.; Thomas, G.; Barbu, V.; Humbert, L.; Despras, G.; et al. Connecting dysbiosis, bile-acid dysmetabolism and gut inflammation in inflammatory bowel diseases. Gut 2013, 62, 531–539. [Google Scholar] [CrossRef] [PubMed]
  105. Walker, A.; Schmitt-Kopplin, P. The role of fecal sulfur metabolome in inflammatory bowel diseases. Int. J. Med. Microbiol. 2021, 311, 151513. [Google Scholar] [CrossRef] [PubMed]
  106. Kostic, A.D.; Xavier, R.J.; Gevers, D. The microbiome in inflammatory bowel disease: Current status and the future ahead. Gastroenterology 2014, 146, 1489–1499. [Google Scholar] [CrossRef] [Green Version]
  107. Khodakivskyi, P.V.; Lauber, C.L.; Yevtodiyenko, A.; Bazhin, A.A.; Bruce, S.; Ringel-Kulka, T.; Ringel, Y.; Bétrisey, B.; Torres, J.; Hu, J.; et al. Noninvasive imaging and quantification of bile salt hydrolase activity: From bacteria to humans. Sci. Adv. 2021, 7, eaaz9857. [Google Scholar] [CrossRef]
  108. Jones, B.V.; Begley, M.; Hill, C.; Gahan, C.G.; Marchesi, J.R. Functional and comparative metagenomic analysis of bile salt hydrolase activity in the human gut microbiome. Proc. Natl. Acad. Sci. USA 2008, 105, 13580–13585. [Google Scholar] [CrossRef]
  109. Lopez-Siles, M.; Khan, T.M.; Duncan, S.H.; Harmsen, H.J.; Garcia-Gil, L.J.; Flint, H.J. Cultured representatives of two major phylogroups of human colonic Faecalibacterium prausnitzii can utilize pectin, uronic acids, and host-derived substrates for growth. Appl. Environ. Microbiol. 2012, 78, 420–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Song, Z.; Cai, Y.; Lao, X.; Wang, X.; Lin, X.; Cui, Y.; Kalavagunta, P.K.; Liao, J.; Jin, L.; Shang, J.; et al. Taxonomic profiling and populational patterns of bacterial bile salt hydrolase (BSH) genes based on worldwide human gut microbiome. Microbiome 2019, 7, 9. [Google Scholar] [CrossRef] [Green Version]
  111. Cummings, J.H.; Macfarlane, G.T.; Macfarlane, S. Intestinal bacteria and ulcerative colitis. Curr. Issues Intest. Microbiol. 2003, 4, 9–20. [Google Scholar] [PubMed]
  112. Kushkevych, I.; Kotrsová, V.; Dordević, D.; Buňková, L.; Vítězová, M.; Amedei, A. Hydrogen Sulfide Effects on the Survival of Lactobacilli with Emphasis on the Development of Inflammatory Bowel Diseases. Biomolecules 2019, 9, 752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Das, P.; Marcišauskas, S.; Ji, B.; Nielsen, J. Metagenomic analysis of bile salt biotransformation in the human gut microbiome. BMC Genom. 2019, 20, 517. [Google Scholar] [CrossRef] [Green Version]
  114. Lloyd-Price, J.; Arze, C.; Ananthakrishnan, A.N.; Schirmer, M.; Avila-Pacheco, J.; Poon, T.W.; Andrews, E.; Ajami, N.J.; Bonham, K.S.; Brislawn, C.J.; et al. Multi-omics of the gut microbial ecosystem in inflammatory bowel diseases. Nature 2019, 569, 655–662. [Google Scholar] [CrossRef]
  115. Laue, H.; Denger, K.; Cook, A.M. Taurine reduction in anaerobic respiration of Bilophila wadsworthia RZATAU. Appl. Environ. Microbiol. 1997, 63, 2016–2021. [Google Scholar] [CrossRef] [PubMed]
  116. Edmond, L.M.; Hopkins, M.J.; Magee, E.A.; Cummings, J.H. The effect of 5-aminosalicylic acid-containing drugs on sulfide production by sulfate-reducing and amino acid-fermenting bacteria. Inflamm. Bowel Dis. 2003, 9, 10–17. [Google Scholar] [CrossRef] [PubMed]
  117. Dubuquoy, L.; Rousseaux, C.; Thuru, X.; Peyrin-Biroulet, L.; Romano, O.; Chavatte, P.; Chamaillard, M.; Desreumaux, P. PPARgamma as a new therapeutic target in inflammatory bowel diseases. Gut 2006, 55, 1341–1349. [Google Scholar] [CrossRef] [Green Version]
  118. Alex, S.; Lange, K.; Amolo, T.; Grinstead, J.S.; Haakonsson, A.K.; Szalowska, E.; Koppen, A.; Mudde, K.; Haenen, D.; Al-Lahham, S.; et al. Short-chain fatty acids stimulate angiopoietin-like 4 synthesis in human colon adenocarcinoma cells by activating peroxisome proliferator-activated receptor γ. Mol. Cell. Biol. 2013, 33, 1303–1316. [Google Scholar] [CrossRef] [Green Version]
  119. Xu, J.; Chen, N.; Wu, Z.; Song, Y.; Zhang, Y.; Wu, N.; Zhang, F.; Ren, X.; Liu, Y. 5-Aminosalicylic Acid Alters the Gut Bacterial Microbiota in Patients with Ulcerative Colitis. Front. Microbiol. 2018, 9, 1274. [Google Scholar] [CrossRef]
  120. Fiorucci, S.; Orlandi, S.; Mencarelli, A.; Caliendo, G.; Santagada, V.; Distrutti, E.; Santucci, L.; Cirino, G.; Wallace, J.L. Enhanced activity of a hydrogen sulphide-releasing derivative of mesalamine (ATB-429) in a mouse model of colitis. Br. J. Pharmacol. 2007, 150, 996–1002. [Google Scholar] [CrossRef] [Green Version]
  121. Roediger, W.E. Decreased sulphur aminoacid intake in ulcerative colitis. Lancet 1998, 351, 1555. [Google Scholar] [CrossRef]
  122. Ijssennagger, N.; van der Meer, R.; van Mil, S.W.C. Sulfide as a Mucus Barrier-Breaker in Inflammatory Bowel Disease? Trends Mol. Med. 2016, 22, 190–199. [Google Scholar] [CrossRef] [PubMed]
  123. Langendijk, P.S.; Hanssen, J.T.; Van der Hoeven, J.S. Sulfate-reducing bacteria in association with human periodontitis. J. Clin. Periodontol. 2000, 27, 943–950. [Google Scholar] [CrossRef] [PubMed]
  124. Ohge, H.; Furne, J.K.; Springfield, J.; Rothenberger, D.A.; Madoff, R.D.; Levitt, M.D. Association between fecal hydrogen sulfide production and pouchitis. Dis. Colon. Rectum 2005, 48, 469–475. [Google Scholar] [CrossRef]
  125. Sahami, S.; Kooij, I.A.; Meijer, S.L.; Van den Brink, G.R.; Buskens, C.J.; Te Velde, A.A. The Link between the Appendix and Ulcerative Colitis: Clinical Relevance and Potential Immunological Mechanisms. Am. J. Gastroenterol. 2016, 111, 163–169. [Google Scholar] [CrossRef]
  126. Rogers, M.B.; Brower-Sinning, R.; Firek, B.; Zhong, D.; Morowitz, M.J. Acute Appendicitis in Children Is Associated with a Local Expansion of Fusobacteria. Clin. Infect. Dis. 2016, 63, 71–78. [Google Scholar] [CrossRef] [Green Version]
  127. Basic, A.; Blomqvist, S.; Carlén, A.; Dahlén, G. Estimation of bacterial hydrogen sulfide production in vitro. J. Oral. Microbiol. 2015, 7, 28166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Zhong, D.; Brower-Sinning, R.; Firek, B.; Morowitz, M.J. Acute appendicitis in children is associated with an abundance of bacteria from the phylum Fusobacteria. J. Pediatr. Surg. 2014, 49, 441–446. [Google Scholar] [CrossRef] [PubMed]
  129. Moore, J.; Babidge, W.; Millard, S.; Roediger, W. Colonic luminal hydrogen sulfide is not elevated in ulcerative colitis. Dig. Dis. Sci. 1998, 43, 162–165. [Google Scholar] [CrossRef]
  130. Jørgensen, J.; Mortensen, P.B. Hydrogen sulfide and colonic epithelial metabolism: Implications for ulcerative colitis. Dig. Dis. Sci. 2001, 46, 1722–1732. [Google Scholar] [CrossRef]
  131. Levitt, M.D.; Springfield, J.; Furne, J.; Koenig, T.; Suarez, F.L. Physiology of sulfide in the rat colon: Use of bismuth to assess colonic sulfide production. J. Appl. Physiol. 2002, 92, 1655–1660. [Google Scholar] [CrossRef] [Green Version]
  132. Macfarlane, S.; McBain, A.J.; Macfarlane, G.T. Consequences of biofilm and sessile growth in the large intestine. Adv. Dent. Res. 1997, 11, 59–68. [Google Scholar] [CrossRef]
  133. Swidsinski, A.; Weber, J.; Loening-Baucke, V.; Hale, L.P.; Lochs, H. Spatial organization and composition of the mucosal flora in patients with inflammatory bowel disease. J. Clin. Microbiol. 2005, 43, 3380–3389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kleessen, B.; Kroesen, A.J.; Buhr, H.J.; Blaut, M. Mucosal and invading bacteria in patients with inflammatory bowel disease compared with controls. Scand. J. Gastroenterol. 2002, 37, 1034–1041. [Google Scholar] [CrossRef]
  135. Swidsinski, A.; Ladhoff, A.; Pernthaler, A.; Swidsinski, S.; Loening-Baucke, V.; Ortner, M.; Weber, J.; Hoffmann, U.; Schreiber, S.; Dietel, M.; et al. Mucosal flora in inflammatory bowel disease. Gastroenterology 2002, 122, 44–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Johansson, M.E.; Gustafsson, J.K.; Holmén-Larsson, J.; Jabbar, K.S.; Xia, L.; Xu, H.; Ghishan, F.K.; Carvalho, F.A.; Gewirtz, A.T.; Sjövall, H.; et al. Bacteria penetrate the normally impenetrable inner colon mucus layer in both murine colitis models and patients with ulcerative colitis. Gut 2014, 63, 281–291. [Google Scholar] [CrossRef] [PubMed]
  137. Swidsinski, A.; Loening-Baucke, V.; Theissig, F.; Engelhardt, H.; Bengmark, S.; Koch, S.; Lochs, H.; Dörffel, Y. Comparative study of the intestinal mucus barrier in normal and inflamed colon. Gut 2007, 56, 343–350. [Google Scholar] [CrossRef]
  138. Pullan, R.D.; Thomas, G.A.; Rhodes, M.; Newcombe, R.G.; Williams, G.T.; Allen, A.; Rhodes, J. Thickness of adherent mucus gel on colonic mucosa in humans and its relevance to colitis. Gut 1994, 35, 353–359. [Google Scholar] [CrossRef] [Green Version]
  139. Strugala, V.; Dettmar, P.W.; Pearson, J.P. Thickness and continuity of the adherent colonic mucus barrier in active and quiescent ulcerative colitis and Crohn’s disease. Int. J. Clin. Pract. 2008, 62, 762–769. [Google Scholar] [CrossRef]
  140. Rankin, B.J.; Srivastava, E.D.; Record, C.O.; Pearson, J.P.; Allen, A. Patients with ulcerative colitis have reduced mucin polymer content in the adherent colonic mucus gel. Biochem. Soc. Trans. 1995, 23, 104S. [Google Scholar] [CrossRef] [Green Version]
  141. Gardiner, K.R.; Halliday, M.I.; Barclay, G.R.; Milne, L.; Brown, D.; Stephens, S.; Maxwell, R.J.; Rowlands, B.J. Significance of systemic endotoxaemia in inflammatory bowel disease. Gut 1995, 36, 897–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Ijssennagger, N.; Belzer, C.; Hooiveld, G.J.; Dekker, J.; van Mil, S.W.; Müller, M.; Kleerebezem, M.; van der Meer, R. Gut microbiota facilitates dietary heme-induced epithelial hyperproliferation by opening the mucus barrier in colon. Proc. Natl. Acad. Sci. USA 2015, 112, 10038–10043. [Google Scholar] [CrossRef] [PubMed]
  143. Blackler, R.W.; Motta, J.P.; Manko, A.; Workentine, M.; Bercik, P.; Surette, M.G.; Wallace, J.L. Hydrogen sulphide protects against NSAID-enteropathy through modulation of bile and the microbiota. Br. J. Pharmacol. 2015, 172, 992–1004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Wallace, J.L.; Ianaro, A.; de Nucci, G. Gaseous Mediators in Gastrointestinal Mucosal Defense and Injury. Dig. Dis. Sci. 2017, 62, 2223–2230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Stummer, N.; Weghuber, D.; Feichtinger, R.G.; Huber, S.; Mayr, J.A.; Kofler, B.; Neureiter, D.; Klieser, E.; Hochmann, S.; Lauth, W.; et al. Hydrogen Sulfide Metabolizing Enzymes in the Intestinal Mucosa in Pediatric and Adult Inflammatory Bowel Disease. Antioxidants 2022, 11, 2235. [Google Scholar] [CrossRef] [PubMed]
  146. Eichele, D.D.; Kharbanda, K.K. Dextran sodium sulfate colitis murine model: An indispensable tool for advancing our understanding of inflammatory bowel diseases pathogenesis. World J. Gastroenterol. 2017, 23, 6016–6029. [Google Scholar] [CrossRef]
  147. Beaumont, M.; Andriamihaja, M.; Lan, A.; Khodorova, N.; Audebert, M.; Blouin, J.M.; Grauso, M.; Lancha, L.; Benetti, P.H.; Benamouzig, R.; et al. Detrimental effects for colonocytes of an increased exposure to luminal hydrogen sulfide: The adaptive response. Free Radic. Biol. Med. 2016, 93, 155–164. [Google Scholar] [CrossRef]
  148. Cooper, C.E.; Brown, G.C. The inhibition of mitochondrial cytochrome oxidase by the gases carbon monoxide, nitric oxide, hydrogen cyanide and hydrogen sulfide: Chemical mechanism and physiological significance. J. Bioenerg. Biomembr. 2008, 40, 533–539. [Google Scholar] [CrossRef]
  149. Andriamihaja, M.; Lan, A.; Beaumont, M.; Grauso, M.; Gotteland, M.; Pastene, E.; Cires, M.J.; Carrasco-Pozo, C.; Tomé, D.; Blachier, F. Proanthocyanidin-containing polyphenol extracts from fruits prevent the inhibitory effect of hydrogen sulfide on human colonocyte oxygen consumption. Amino Acids 2018, 50, 755–763. [Google Scholar] [CrossRef]
  150. Mimoun, S.; Andriamihaja, M.; Chaumontet, C.; Atanasiu, C.; Benamouzig, R.; Blouin, J.M.; Tomé, D.; Bouillaud, F.; Blachier, F. Detoxification of H2S by differentiated colonic epithelial cells: Implication of the sulfide oxidizing unit and of the cell respiratory capacity. Antioxid. Redox Signal 2012, 17, 1–10. [Google Scholar] [CrossRef]
  151. Hirata, I.; Naito, Y.; Takagi, T.; Mizushima, K.; Suzuki, T.; Omatsu, T.; Handa, O.; Ichikawa, H.; Ueda, H.; Yoshikawa, T. Endogenous hydrogen sulfide is an anti-inflammatory molecule in dextran sodium sulfate-induced colitis in mice. Dig. Dis. Sci. 2011, 56, 1379–1386. [Google Scholar] [CrossRef]
  152. Picton, R.; Eggo, M.C.; Langman, M.J.; Singh, S. Impaired detoxication of hydrogen sulfide in ulcerative colitis? Dig. Dis. Sci. 2007, 52, 373–378. [Google Scholar] [CrossRef] [PubMed]
  153. Zhang, J.; Cen, L.; Zhang, X.; Tang, C.; Chen, Y.; Zhang, Y.; Yu, M.; Lu, C.; Li, M.; Li, S.; et al. MPST deficiency promotes intestinal epithelial cell apoptosis and aggravates inflammatory bowel disease via AKT. Redox Biol. 2022, 56, 102469. [Google Scholar] [CrossRef]
  154. Arijs, I.; Vanhove, W.; Rutgeerts, P.; Schuit, F.; Verbeke, K.; De Preter, V. Decreased mucosal sulfide detoxification capacity in patients with Crohn’s disease. Inflamm. Bowel Dis. 2013, 19, E70–E72. [Google Scholar] [CrossRef] [PubMed]
  155. Furne, J.; Springfield, J.; Koenig, T.; DeMaster, E.; Levitt, M.D. Oxidation of hydrogen sulfide and methanethiol to thiosulfate by rat tissues: A specialized function of the colonic mucosa. Biochem. Pharmacol. 2001, 62, 255–259. [Google Scholar] [CrossRef]
  156. Flannigan, K.L.; Agbor, T.A.; Blackler, R.W.; Kim, J.J.; Khan, W.I.; Verdu, E.F.; Ferraz, J.G.; Wallace, J.L. Impaired hydrogen sulfide synthesis and IL-10 signaling underlie hyperhomocysteinemia-associated exacerbation of colitis. Proc. Natl. Acad. Sci. USA 2014, 111, 13559–13564. [Google Scholar] [CrossRef] [PubMed]
  157. Vagianos, K.; Bector, S.; McConnell, J.; Bernstein, C.N. Nutrition assessment of patients with inflammatory bowel disease. JPEN J. Parenter. Enteral Nutr. 2007, 31, 311–319. [Google Scholar] [CrossRef]
  158. Goubern, M.; Andriamihaja, M.; Nübel, T.; Blachier, F.; Bouillaud, F. Sulfide, the first inorganic substrate for human cells. FASEB J. 2007, 21, 1699–1706. [Google Scholar] [CrossRef] [PubMed]
  159. Zanardo, R.C.; Brancaleone, V.; Distrutti, E.; Fiorucci, S.; Cirino, G.; Wallace, J.L. Hydrogen sulfide is an endogenous modulator of leukocyte-mediated inflammation. FASEB J. 2006, 20, 2118–2120. [Google Scholar] [CrossRef] [PubMed]
  160. Wallace, J.L.; Caliendo, G.; Santagada, V.; Cirino, G.; Fiorucci, S. Gastrointestinal safety and anti-inflammatory effects of a hydrogen sulfide-releasing diclofenac derivative in the rat. Gastroenterology 2007, 132, 261–271. [Google Scholar] [CrossRef]
  161. Distrutti, E.; Sediari, L.; Mencarelli, A.; Renga, B.; Orlandi, S.; Antonelli, E.; Roviezzo, F.; Morelli, A.; Cirino, G.; Wallace, J.L.; et al. Evidence that hydrogen sulfide exerts antinociceptive effects in the gastrointestinal tract by activating KATP channels. J. Pharmacol. Exp. Ther. 2006, 316, 325–335. [Google Scholar] [CrossRef] [Green Version]
  162. Wallace, J.L.; Ferraz, J.G.; Muscara, M.N. Hydrogen sulfide: An endogenous mediator of resolution of inflammation and injury. Antioxid. Redox Signal 2012, 17, 58–67. [Google Scholar] [CrossRef] [PubMed]
  163. Lee, H.J.; Lee, H.G.; Choi, K.S.; Surh, Y.J.; Na, H.K. Diallyl trisulfide suppresses dextran sodium sulfate-induced mouse colitis: NF-κB and STAT3 as potential targets. Biochem. Biophys. Res. Commun. 2013, 437, 267–273. [Google Scholar] [CrossRef] [PubMed]
  164. Peake, B.F.; Nicholson, C.K.; Lambert, J.P.; Hood, R.L.; Amin, H.; Amin, S.; Calvert, J.W. Hydrogen sulfide preconditions the db/db diabetic mouse heart against ischemia-reperfusion injury by activating Nrf2 signaling in an Erk-dependent manner. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H1215–H1224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Yang, G.; Zhao, K.; Ju, Y.; Mani, S.; Cao, Q.; Puukila, S.; Khaper, N.; Wu, L.; Wang, R. Hydrogen sulfide protects against cellular senescence via S-sulfhydration of Keap1 and activation of Nrf2. Antioxid. Redox Signal 2013, 18, 1906–1919. [Google Scholar] [CrossRef]
  166. Elrod, J.W.; Calvert, J.W.; Morrison, J.; Doeller, J.E.; Kraus, D.W.; Tao, L.; Jiao, X.; Scalia, R.; Kiss, L.; Szabo, C.; et al. Hydrogen sulfide attenuates myocardial ischemia-reperfusion injury by preservation of mitochondrial function. Proc. Natl. Acad. Sci. USA 2007, 104, 15560–15565. [Google Scholar] [CrossRef] [PubMed]
  167. Chan, M.V.; Wallace, J.L. Hydrogen sulfide-based therapeutics and gastrointestinal diseases: Translating physiology to treatments. Am. J. Physiol. Gastrointest. Liver Physiol. 2013, 305, G467–G473. [Google Scholar] [CrossRef] [Green Version]
  168. Oh, G.S.; Pae, H.O.; Lee, B.S.; Kim, B.N.; Kim, J.M.; Kim, H.R.; Jeon, S.B.; Jeon, W.K.; Chae, H.J.; Chung, H.T. Hydrogen sulfide inhibits nitric oxide production and nuclear factor-kappaB via heme oxygenase-1 expression in RAW264.7 macrophages stimulated with lipopolysaccharide. Free Radic. Biol. Med. 2006, 41, 106–119. [Google Scholar] [CrossRef]
  169. Wallace, J.L.; Dicay, M.; McKnight, W.; Martin, G.R. Hydrogen sulfide enhances ulcer healing in rats. FASEB J. 2007, 21, 4070–4076. [Google Scholar] [CrossRef]
  170. Zhao, H.; Yan, R.; Zhou, X.; Ji, F.; Zhang, B. Hydrogen sulfide improves colonic barrier integrity in DSS-induced inflammation in Caco-2 cells and mice. Int. Immunopharmacol. 2016, 39, 121–127. [Google Scholar] [CrossRef]
  171. Li, T.; Zhao, B.; Wang, C.; Wang, H.; Liu, Z.; Li, W.; Jin, H.; Tang, C.; Du, J. Regulatory effects of hydrogen sulfide on IL-6, IL-8 and IL-10 levels in the plasma and pulmonary tissue of rats with acute lung injury. Exp. Biol. Med. 2008, 233, 1081–1087. [Google Scholar] [CrossRef]
  172. Wu, J.; Wei, J.; You, X.; Chen, X.; Zhu, H.; Zhu, X.; Liu, Y.; Xu, M. Inhibition of hydrogen sulfide generation contributes to lung injury after experimental orthotopic lung transplantation. J. Surg. Res. 2013, 182, e25–e33. [Google Scholar] [CrossRef] [PubMed]
  173. Zeng, J.; Lin, X.; Fan, H.; Li, C. Hydrogen sulfide attenuates the inflammatory response in a mouse burn injury model. Mol. Med. Rep. 2013, 8, 1204–1208. [Google Scholar] [CrossRef] [Green Version]
  174. Tokuda, K.; Kida, K.; Marutani, E.; Crimi, E.; Bougaki, M.; Khatri, A.; Kimura, H.; Ichinose, F. Inhaled hydrogen sulfide prevents endotoxin-induced systemic inflammation and improves survival by altering sulfide metabolism in mice. Antioxid. Redox Signal 2012, 17, 11–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Noguchi, E.; Homma, Y.; Kang, X.; Netea, M.G.; Ma, X. A Crohn’s disease-associated NOD2 mutation suppresses transcription of human IL10 by inhibiting activity of the nuclear ribonucleoprotein hnRNP-A1. Nat. Immunol. 2009, 10, 471–479. [Google Scholar] [CrossRef] [Green Version]
  176. van der Linde, K.; Boor, P.P.; Sandkuijl, L.A.; Meijssen, M.A.; Savelkoul, H.F.; Wilson, J.H.; de Rooij, F.W. A Gly15Arg mutation in the interleukin-10 gene reduces secretion of interleukin-10 in Crohn disease. Scand. J. Gastroenterol. 2003, 38, 611–617. [Google Scholar] [PubMed]
  177. Franke, A.; Balschun, T.; Karlsen, T.H.; Sventoraityte, J.; Nikolaus, S.; Mayr, G.; Domingues, F.S.; Albrecht, M.; Nothnagel, M.; Ellinghaus, D.; et al. Sequence variants in IL10, ARPC2 and multiple other loci contribute to ulcerative colitis susceptibility. Nat. Genet. 2008, 40, 1319–1323. [Google Scholar] [CrossRef]
  178. Shouval, D.S.; Biswas, A.; Goettel, J.A.; McCann, K.; Conaway, E.; Redhu, N.S.; Mascanfroni, I.D.; Al Adham, Z.; Lavoie, S.; Ibourk, M.; et al. Interleukin-10 receptor signaling in innate immune cells regulates mucosal immune tolerance and anti-inflammatory macrophage function. Immunity 2014, 40, 706–719. [Google Scholar] [CrossRef]
  179. Fuss, I.J.; Boirivant, M.; Lacy, B.; Strober, W. The interrelated roles of TGF-beta and IL-10 in the regulation of experimental colitis. J. Immunol. 2002, 168, 900–908. [Google Scholar] [CrossRef]
  180. Lesage, S.; Zouali, H.; Cézard, J.P.; Colombel, J.F.; Belaiche, J.; Almer, S.; Tysk, C.; O’Morain, C.; Gassull, M.; Binder, V.; et al. CARD15/NOD2 mutational analysis and genotype-phenotype correlation in 612 patients with inflammatory bowel disease. Am. J. Hum. Genet. 2002, 70, 845–857. [Google Scholar] [CrossRef] [Green Version]
  181. Zhang, H.X.; Liu, S.J.; Tang, X.L.; Duan, G.L.; Ni, X.; Zhu, X.Y.; Liu, Y.J.; Wang, C.N. H2S Attenuates LPS-Induced Acute Lung Injury by Reducing Oxidative/Nitrative Stress and Inflammation. Cell. Physiol. Biochem. 2016, 40, 1603–1612. [Google Scholar] [CrossRef] [PubMed]
  182. Younis, N.N.; Shaheen, M.A.; Mahmoud, M.F. Silymarin preconditioning protected insulin resistant rats from liver ischemia-reperfusion injury: Role of endogenous H2S. J. Surg. Res. 2016, 204, 398–409. [Google Scholar] [CrossRef] [PubMed]
  183. Liu, R.; Qu, H.; Xie, J.; Long, J.; Bai, Q.; Chen, Y.; Mao, H. H2S-mediated aerobic exercise antagonizes the hippocampal inflammatory response in CUMS-depressed mice. J. Affect. Disord. 2021, 283, 410–419. [Google Scholar] [CrossRef]
  184. Martin, G.R.; Wallace, J.L. Gastrointestinal inflammation: A central component of mucosal defense and repair. Exp. Biol. Med. 2006, 231, 130–137. [Google Scholar] [CrossRef]
  185. Fiorucci, S.; Antonelli, E.; Distrutti, E.; Rizzo, G.; Mencarelli, A.; Orlandi, S.; Zanardo, R.; Renga, B.; Di Sante, M.; Morelli, A.; et al. Inhibition of hydrogen sulfide generation contributes to gastric injury caused by anti-inflammatory nonsteroidal drugs. Gastroenterology 2005, 129, 1210–1224. [Google Scholar] [CrossRef] [Green Version]
  186. Martin, G.R.; McKnight, G.W.; Dicay, M.S.; Coffin, C.S.; Ferraz, J.G.; Wallace, J.L. Hydrogen sulphide synthesis in the rat and mouse gastrointestinal tract. Dig. Liver Dis. 2010, 42, 103–109. [Google Scholar] [CrossRef] [PubMed]
  187. Gaudichon, C.; Bos, C.; Morens, C.; Petzke, K.J.; Mariotti, F.; Everwand, J.; Benamouzig, R.; Daré, S.; Tomé, D.; Metges, C.C. Ileal losses of nitrogen and amino acids in humans and their importance to the assessment of amino acid requirements. Gastroenterology 2002, 123, 50–59. [Google Scholar] [CrossRef]
  188. Florin, T.H. Hydrogen sulphide and total acid-volatile sulphide in faeces, determined with a direct spectrophotometric method. Clin. Chim. Acta 1991, 196, 127–134. [Google Scholar] [CrossRef]
  189. Buffière, C.; Gaudichon, C.; Hafnaoui, N.; Migné, C.; Scislowsky, V.; Khodorova, N.; Mosoni, L.; Blot, A.; Boirie, Y.; Dardevet, D.; et al. In the elderly, meat protein assimilation from rare meat is lower than that from meat that is well done. Am. J. Clin. Nutr. 2017, 106, 1257–1266. [Google Scholar] [CrossRef] [Green Version]
  190. Mills, D.J.; Tuohy, K.M.; Booth, J.; Buck, M.; Crabbe, M.J.; Gibson, G.R.; Ames, J.M. Dietary glycated protein modulates the colonic microbiota towards a more detrimental composition in ulcerative colitis patients and non-ulcerative colitis subjects. J. Appl. Microbiol. 2008, 105, 706–714. [Google Scholar] [CrossRef]
  191. Rémond, D.; Machebeuf, M.; Yven, C.; Buffière, C.; Mioche, L.; Mosoni, L.; Patureau Mirand, P. Postprandial whole-body protein metabolism after a meat meal is influenced by chewing efficiency in elderly subjects. Am. J. Clin. Nutr. 2007, 85, 1286–1292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Lewis, S.; Cochrane, S. Alteration of sulfate and hydrogen metabolism in the human colon by changing intestinal transit rate. Am. J. Gastroenterol. 2007, 102, 624–633. [Google Scholar] [CrossRef] [PubMed]
  193. Flynn, S.; Eisenstein, S. Inflammatory Bowel Disease Presentation and Diagnosis. Surg. Clin. N. Am. 2019, 99, 1051–1062. [Google Scholar] [CrossRef] [PubMed]
  194. Kashyap, P.C.; Reigstad, C.S.; Loftus, E.V. Role of diet and gut microbiota in management of inflammatory bowel disease in an Asian migrant. J. Allergy Clin. Immunol. 2013, 132, 250–250.e5. [Google Scholar] [CrossRef]
  195. Chiba, M.; Tsuda, S.; Komatsu, M.; Tozawa, H.; Takayama, Y. Onset of Ulcerative Colitis during a Low-Carbohydrate Weight-Loss Diet and Treatment with a Plant-Based Diet: A Case Report. Perm. J. 2016, 20, 80–84. [Google Scholar] [CrossRef] [Green Version]
  196. Chiba, M.; Sugawara, T.; Komatsu, M.; Tozawa, H. Onset of Ulcerative Colitis in the Second Trimester after Emesis Gravidarum: Treatment with Plant-based Diet. Inflamm. Bowel Dis. 2018, 24, e8–e9. [Google Scholar] [CrossRef]
  197. Chiba, M.; Nakane, K.; Tsuji, T.; Tsuda, S.; Ishii, H.; Ohno, H.; Watanabe, K.; Ito, M.; Komatsu, M.; Yamada, K.; et al. Relapse Prevention in Ulcerative Colitis by Plant-Based Diet Through Educational Hospitalization: A Single-Group Trial. Perm. J. 2018, 22, 17–167. [Google Scholar] [CrossRef] [Green Version]
  198. Roediger, W.E.; Moore, J.; Babidge, W. Colonic sulfide in pathogenesis and treatment of ulcerative colitis. Dig. Dis. Sci. 1997, 42, 1571–1579. [Google Scholar] [CrossRef]
  199. David, L.A.; Maurice, C.F.; Carmody, R.N.; Gootenberg, D.B.; Button, J.E.; Wolfe, B.E.; Ling, A.V.; Devlin, A.S.; Varma, Y.; Fischbach, M.A.; et al. Diet rapidly and reproducibly alters the human gut microbiome. Nature 2014, 505, 559–563. [Google Scholar] [CrossRef] [Green Version]
  200. Mu, C.; Yang, Y.; Luo, Z.; Guan, L.; Zhu, W. The Colonic Microbiome and Epithelial Transcriptome Are Altered in Rats Fed a High-Protein Diet Compared with a Normal-Protein Diet. J. Nutr. 2016, 146, 474–483. [Google Scholar] [CrossRef] [Green Version]
  201. Fuentes, S.; Rossen, N.G.; van der Spek, M.J.; Hartman, J.H.; Huuskonen, L.; Korpela, K.; Salojärvi, J.; Aalvink, S.; de Vos, W.M.; D’Haens, G.R.; et al. Microbial shifts and signatures of long-term remission in ulcerative colitis after faecal microbiota transplantation. ISME J. 2017, 11, 1877–1889. [Google Scholar] [CrossRef] [Green Version]
  202. Bajer, L.; Kverka, M.; Kostovcik, M.; Macinga, P.; Dvorak, J.; Stehlikova, Z.; Brezina, J.; Wohl, P.; Spicak, J.; Drastich, P. Distinct gut microbiota profiles in patients with primary sclerosing cholangitis and ulcerative colitis. World J. Gastroenterol. 2017, 23, 4548–4558. [Google Scholar] [CrossRef]
  203. Vermeire, S.; Joossens, M.; Verbeke, K.; Wang, J.; Machiels, K.; Sabino, J.; Ferrante, M.; Van Assche, G.; Rutgeerts, P.; Raes, J. Donor Species Richness Determines Faecal Microbiota Transplantation Success in Inflammatory Bowel Disease. J. Crohns Colitis 2016, 10, 387–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Macfarlane, G.T.; Gibson, G.R.; Cummings, J.H. Comparison of fermentation reactions in different regions of the human colon. J. Appl. Bacteriol. 1992, 72, 57–64. [Google Scholar] [CrossRef] [PubMed]
  205. Rao, S.S.; Kuo, B.; McCallum, R.W.; Chey, W.D.; DiBaise, J.K.; Hasler, W.L.; Koch, K.L.; Lackner, J.M.; Miller, C.; Saad, R.; et al. Investigation of colonic and whole-gut transit with wireless motility capsule and radiopaque markers in constipation. Clin. Gastroenterol. Hepatol. 2009, 7, 537–544. [Google Scholar] [CrossRef] [PubMed]
  206. Walker, A.W.; Duncan, S.H.; McWilliam Leitch, E.C.; Child, M.W.; Flint, H.J. pH and peptide supply can radically alter bacterial populations and short-chain fatty acid ratios within microbial communities from the human colon. Appl. Environ. Microbiol. 2005, 71, 3692–3700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Gibson, G.R.; Cummings, J.H.; Macfarlane, G.T.; Allison, C.; Segal, I.; Vorster, H.H.; Walker, A.R. Alternative pathways for hydrogen disposal during fermentation in the human colon. Gut 1990, 31, 679–683. [Google Scholar] [CrossRef] [PubMed]
  208. Zimmer, J.; Lange, B.; Frick, J.S.; Sauer, H.; Zimmermann, K.; Schwiertz, A.; Rusch, K.; Klosterhalfen, S.; Enck, P. A vegan or vegetarian diet substantially alters the human colonic faecal microbiota. Eur. J. Clin. Nutr. 2012, 66, 53–60. [Google Scholar] [CrossRef]
  209. Zhu, Q.C.; Gao, R.Y.; Wu, W.; Guo, B.M.; Peng, J.Y.; Qin, H.L. Effect of a high-fat diet in development of colonic adenoma in an animal model. World J. Gastroenterol. 2014, 20, 8119–8129. [Google Scholar] [CrossRef]
  210. Devkota, S.; Wang, Y.; Musch, M.W.; Leone, V.; Fehlner-Peach, H.; Nadimpalli, A.; Antonopoulos, D.A.; Jabri, B.; Chang, E.B. Dietary-fat-induced taurocholic acid promotes pathobiont expansion and colitis in Il10-/- mice. Nature 2012, 487, 104–108. [Google Scholar] [CrossRef] [Green Version]
  211. de Vries, J.H.M.; Dijkhuizen, M.; Tap, P.; Witteman, B.J.M. Patient’s Dietary Beliefs and Behaviours in Inflammatory Bowel Disease. Dig. Dis. 2019, 37, 131–139. [Google Scholar] [CrossRef] [PubMed]
  212. Ruemmele, F.M.; Veres, G.; Kolho, K.L.; Griffiths, A.; Levine, A.; Escher, J.C.; Amil Dias, J.; Barabino, A.; Braegger, C.P.; Bronsky, J.; et al. Consensus guidelines of ECCO/ESPGHAN on the medical management of pediatric Crohn’s disease. J. Crohns Colitis 2014, 8, 1179–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Day, A.S.; Whitten, K.E.; Sidler, M.; Lemberg, D.A. Systematic review: Nutritional therapy in paediatric Crohn’s disease. Aliment. Pharmacol. Ther. 2008, 27, 293–307. [Google Scholar] [CrossRef] [PubMed]
  214. Otley, A.R.; Russell, R.K.; Day, A.S. Nutritional therapy for the treatment of pediatric Crohn’s disease. Expert. Rev. Clin. Immunol. 2010, 6, 667–676. [Google Scholar] [CrossRef]
  215. Levine, A.; Turner, D.; Pfeffer Gik, T.; Amil Dias, J.; Veres, G.; Shaoul, R.; Staiano, A.; Escher, J.; Kolho, K.L.; Paerregaard, A.; et al. Comparison of outcomes parameters for induction of remission in new onset pediatric Crohn’s disease: Evaluation of the porto IBD group “growth relapse and outcomes with therapy” (GROWTH CD) study. Inflamm. Bowel Dis. 2014, 20, 278–285. [Google Scholar] [CrossRef]
  216. MacLellan, A.; Moore-Connors, J.; Grant, S.; Cahill, L.; Langille, M.G.I.; Van Limbergen, J. The Impact of Exclusive Enteral Nutrition (EEN) on the Gut Microbiome in Crohn’s Disease: A Review. Nutrients 2017, 9, 447. [Google Scholar] [CrossRef]
  217. Levine, A.; Wine, E. Effects of enteral nutrition on Crohn’s disease: Clues to the impact of diet on disease pathogenesis. Inflamm. Bowel Dis. 2013, 19, 1322–1329. [Google Scholar] [CrossRef] [PubMed]
  218. Day, A.S.; Lopez, R.N. Exclusive enteral nutrition in children with Crohn’s disease. World J. Gastroenterol. 2015, 21, 6809–6816. [Google Scholar] [CrossRef]
  219. Quince, C.; Ijaz, U.Z.; Loman, N.; Eren, A.M.; Saulnier, D.; Russell, J.; Haig, S.J.; Calus, S.T.; Quick, J.; Barclay, A.; et al. Extensive Modulation of the Fecal Metagenome in Children With Crohn’s Disease During Exclusive Enteral Nutrition. Am. J. Gastroenterol. 2015, 110, 1718–1729; quiz 1730. [Google Scholar] [CrossRef] [Green Version]
  220. Wu, G.D.; Chen, J.; Hoffmann, C.; Bittinger, K.; Chen, Y.Y.; Keilbaugh, S.A.; Bewtra, M.; Knights, D.; Walters, W.A.; Knight, R.; et al. Linking long-term dietary patterns with gut microbial enterotypes. Science 2011, 334, 105–108. [Google Scholar] [CrossRef] [Green Version]
  221. Gerasimidis, K.; Bertz, M.; Hanske, L.; Junick, J.; Biskou, O.; Aguilera, M.; Garrick, V.; Russell, R.K.; Blaut, M.; McGrogan, P.; et al. Decline in presumptively protective gut bacterial species and metabolites are paradoxically associated with disease improvement in pediatric Crohn’s disease during enteral nutrition. Inflamm. Bowel Dis. 2014, 20, 861–871. [Google Scholar] [CrossRef] [PubMed]
  222. Colombel, J.F.; Mahadevan, U. Inflammatory Bowel Disease 2017: Innovations and Changing Paradigms. Gastroenterology 2017, 152, 309–312. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hydrogen sulfide (H2S) production and detoxification. Exogenous H2S is a metabolic product of the degradation of sulfate (SO42−) through sulfate-reducing bacteria (SRB) or degradation of sulfur-containing compounds by intestinal bacteria. Endogenous H2S results from degradation of L-cysteine with or without homocysteine by cystathionine-γ-lyase (CSE) and cystathionine-β-synthase (CBS) in the cytosol and from 3-mercaptopyruvate (3MP) by 3-mercapto-sulfurtransferase (3-MST) in mitochondria. The detoxification process is catalyzed by sulfide:quinone oxidoreductase (SQOR) and subsequently by thiosulfate sulfurtransferase (TST) or ethylmalonic encephalopathy 1 protein (ETHE1). Detoxification occurs solely in the mitochondria. Cysteine aminotransferase (CAT), D-amino acid oxidase (DAO), glutathione persulfide (GSSH), thiosulfate (SSO32−), sulfite (SO32−).
Figure 1. Hydrogen sulfide (H2S) production and detoxification. Exogenous H2S is a metabolic product of the degradation of sulfate (SO42−) through sulfate-reducing bacteria (SRB) or degradation of sulfur-containing compounds by intestinal bacteria. Endogenous H2S results from degradation of L-cysteine with or without homocysteine by cystathionine-γ-lyase (CSE) and cystathionine-β-synthase (CBS) in the cytosol and from 3-mercaptopyruvate (3MP) by 3-mercapto-sulfurtransferase (3-MST) in mitochondria. The detoxification process is catalyzed by sulfide:quinone oxidoreductase (SQOR) and subsequently by thiosulfate sulfurtransferase (TST) or ethylmalonic encephalopathy 1 protein (ETHE1). Detoxification occurs solely in the mitochondria. Cysteine aminotransferase (CAT), D-amino acid oxidase (DAO), glutathione persulfide (GSSH), thiosulfate (SSO32−), sulfite (SO32−).
Antioxidants 12 01570 g001
Figure 2. Flow chart of the screening and selection process.
Figure 2. Flow chart of the screening and selection process.
Antioxidants 12 01570 g002
Table 1. Overview of studies on changes in hydrogen sulfide-producing and -detoxifying enzymes (cystathionine-β-synthase (CBS), cystathionine-γ-lyase (CSE), 3-mercapto-sulfurtransferase (3-MST), ethylmalonic encephalopathy 1 protein (ETHE1), sulfide:quinone oxidoreductase (SQOR) and thiosulfate sulfurtransferase (TST)) in human and animal samples. Ulcerative colitis (UC), Crohn’s disease (CD), dextran sulfate sodium (DSS).
Table 1. Overview of studies on changes in hydrogen sulfide-producing and -detoxifying enzymes (cystathionine-β-synthase (CBS), cystathionine-γ-lyase (CSE), 3-mercapto-sulfurtransferase (3-MST), ethylmalonic encephalopathy 1 protein (ETHE1), sulfide:quinone oxidoreductase (SQOR) and thiosulfate sulfurtransferase (TST)) in human and animal samples. Ulcerative colitis (UC), Crohn’s disease (CD), dextran sulfate sodium (DSS).
EnzymeMethodDifference in Enzyme Expression in Comparison to the Control GroupSpeciesReferenceYear
CBSmRNA measurementincreasedMouse (DSS-induced colitis)Hirata et al. [151]2011
CBSmRNA and protein expressiondecreasedMouse (Helicobacter hepaticus induced colitis in mice without adaptive immune systemDe Cicco et al. [16]2018
CBSProtein expression (Western blot)first lower (days 3–7 after induction), then increased Rat (trinitrobenzene sulfonic acid induced colitis)Wallace et al. [14]2009
CBSImmunohistochemical stainingincreasedRat (trinitrobenzene sulfonic acid induced colitis)Wallace et al. [14]2009
CSEmRNA measurementincreasedMouse (DSS-induced colitis)Hirata et al. [151]2011
CSEProtein expression (Western blot)decreasedRat (trinitrobenzene sulfonic acid induced colitis)Wallace et al. [14]2009
CSEImmunohistochemical staining Unstained epithelial cells, while increased staining of mucosa and submucosa Rat (trinitrobenzene sulfonic acid induced colitis)Wallace et al. [14]2009
CSEmRNA and protein expression (Western blot)decreased protein levels, decreased mRNARat and mouse (DSS-induced colitis)Taniguchi et al. [17]2009
CSEImmunohistochemical stainingdecreasedHuman (CD + UC)Stummer et al. [145]2022
3-MSTProtein levels (Western blot)decreasedHuman (CD + UC)Zhang et al. [153]2022
3-MSTImmunohistochemical stainingdecreased Human (CD + UC)Zhang et al. [153]2022
3-MSTmRNA and protein expressiondecreasedMouse (DSS induced colitis)Zhang et al. [153]2022
3-MSTImmunohistochemical stainingdecreasedHuman (CD + UC)Stummer et al. [145]2022
ETHE1mRNAdecreasedHuman (CD)Mottawea et al. [39]2016
ETHE1Immunohistochemical stainingdecreased, except for the terminal ileum in pediatric patientsHuman (CD + UC)Stummer et al. [145]2022
SQORmRNAdecreasedHuman (CD)Mottawea et al. [39]2016
SQORImmunohistochemical stainingdecreased, except for the terminal ileum Human (CD + UC)Stummer et al. [145]2022
TSTmRNA and protein expression (Western blot)decreasedRat and mouse (DSS induced colitis)Taniguchi et al. [17]2009
TSTmRNAdecreasedHuman (UC)De Preter et al. [100]2012
TSTmRNAdecreasedHuman (CD)Mottawea et al. [39]2016
TSTImmunohistochemical stainingdecreasedHuman (CD + UC)Stummer et al. [145]2022
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

Stummer, N.; Feichtinger, R.G.; Weghuber, D.; Kofler, B.; Schneider, A.M. Role of Hydrogen Sulfide in Inflammatory Bowel Disease. Antioxidants 2023, 12, 1570. https://doi.org/10.3390/antiox12081570

AMA Style

Stummer N, Feichtinger RG, Weghuber D, Kofler B, Schneider AM. Role of Hydrogen Sulfide in Inflammatory Bowel Disease. Antioxidants. 2023; 12(8):1570. https://doi.org/10.3390/antiox12081570

Chicago/Turabian Style

Stummer, Nathalie, René G. Feichtinger, Daniel Weghuber, Barbara Kofler, and Anna M. Schneider. 2023. "Role of Hydrogen Sulfide in Inflammatory Bowel Disease" Antioxidants 12, no. 8: 1570. https://doi.org/10.3390/antiox12081570

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