Next Article in Journal
Parnassin, a Novel Therapeutic Peptide, Alleviates Skin Lesions in a DNCB-Induced Atopic Dermatitis Mouse Model
Next Article in Special Issue
Nutritional Metabolomics in Diet–Breast Cancer Relations: Current Research, Challenges, and Future Directions—A Review
Previous Article in Journal
Hereditary Breast Cancer in Romania—Molecular Particularities and Genetic Counseling Challenges in an Eastern European Country
Previous Article in Special Issue
Clinical Relevance of Gut Microbiota Alterations under the Influence of Selected Drugs—Updated Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Influence of Gut Microbiota on Oxidative Stress and the Immune System

Department of Internal Medicine I, Gastroenterology, Hepatology, Endocrinology, Rheumatology and Infectious Diseases, University Hospital Regensburg, 93053 Regensburg, Bavaria, Germany
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(5), 1388; https://doi.org/10.3390/biomedicines11051388
Submission received: 19 April 2023 / Revised: 4 May 2023 / Accepted: 5 May 2023 / Published: 8 May 2023

Abstract

:
The human gastrointestinal tract is home to a complex microbial community that plays an important role in the general well-being of the entire organism. The gut microbiota generates a variety of metabolites and thereby regulates many biological processes, such as the regulation of the immune system. In the gut, bacteria are in direct contact with the host. The major challenge here is to prevent unwanted inflammatory reactions on one hand and on the other hand to ensure that the immune system can be activated when pathogens invade. Here the REDOX equilibrium is of utmost importance. This REDOX equilibrium is controlled by the microbiota either directly or indirectly via bacterial-derived metabolites. A balanced microbiome sorts for a stable REDOX balance, whereas dysbiosis destabilizes this equilibrium. An imbalanced REDOX status directly affects the immune system by disrupting intracellular signaling and promoting inflammatory responses. Here we (i) focus on the most common reactive oxygen species (ROS) and (ii) define the transition from a balanced REDOX state to oxidative stress. Further, we (iii) describe the role of ROS in regulating the immune system and inflammatory responses. Thereafter, we (iv) examine the influence of microbiota on REDOX homeostasis and how shifts in pro- and anti-oxidative cellular conditions can suppress or promote immune responses or inflammation.

1. Introduction

Reactive oxygen species (ROS) are generated in living cells. Initially, it was discovered that ROS are formed as by-products of enzymatic reactions. However, a few years later, it became evident that ROS are also generated in a controlled manner by eukaryotic cells. Thus, it was obvious that ROS not only are harmful by-products but also exert physiological functions as intracellular and intercellular messengers. Various proteins can be modified by ROS, for example, p53, Jun, Fos, and the NF-κB subunits p50 and p65. The oxidation of these proteins leads either to the stimulation (p50) or inhibition (p53, Jun, Fos, and p65) of these proteins; thus, ROS play an important role in intracellular signaling [1,2]. Long-lived ROS that can cross membranes also affect neighboring cells. Thus, ROS also play a role in the communication between cells [1,3]. To enable these signaling functions and prevent oxidative damage, the balance between pro-oxidative and antioxidative molecules must be strictly controlled. If this REDOX equilibrium is disturbed, oxidative stress and cell and tissue damage will occur [4].
Oxidative signals are particularly important for the activation of our immune system. Here ROS are generated by a metabolic switch from cellular respiration to glycolysis. ROS production is essential for the regulation of an appropriate immune response [5,6,7]. In addition, ROS can also be used by antigen-presenting cells (e.g., monocytes and B cells) and neutrophils as a defense mechanism against pathogens [6].
The gut microbiome is unique to each individual. There appears to be a balance in composition and diversity that is beneficial for the host and suppresses inflammation. The intestine is the only place where continuous activation of the immune system through direct contact with microbiota occurs. Under physiological conditions, there is a balance of pro- and anti-inflammatory mechanisms. This balance is maintained by microbiota influencing the REDOX system. Commensal bacteria often exhibit anti-oxidative properties and suppress inflammatory reactions. Pathogenic microbiota induce inflammation and shift the REDOX balance toward a pro-oxidative status [8]. Therefore, the interaction between the intestinal microbiota and the host’s cells, especially the immune cells, is crucial in maintaining the REDOX equilibrium and suppressing unwanted inflammation.

2. The Cellular REDOX Equilibrium

In a physiological state, cells display a balanced REDOX equilibrium. This depends on one hand on the production of reactive molecules and on the other hand on the oxidative defense. Under physiological conditions, the REDOX balance allows oxidative signaling while inhibiting oxidative damage. Under pathophysiological conditions, this balance can shift so that either cellular signaling is impaired and/or oxidative damage is promoted. This is referred to as oxidative stress.

2.1. Reactive Oxygen (ROS)

To date, the best-characterized ROS include superoxide anions (O2·), hydroxyl radicals (·OH), and hydrogen peroxide (H2O2) [4,5]. Cellular ROS production starts with the transfer of an electron to oxygen. This leads to the formation of O2·. Due to their energetically unstable state, these molecules are highly reactive and have only a short half-life of about 1 µs. Furthermore, they cannot freely cross cellular membranes due to their charge. As a result, O2· has a locally limited effect and is responsible for oxidative damage rather than acting as a signaling molecule [1,7].
In an aqueous environment, O2· rapidly converts to H2O2. Intracellularly, this process is accelerated by superoxide dismutases (SODs) [9,10,11]. Although H2O2 is not a radical, it is classified as an ROS. H2O2 displays a longer half-life of about 1 ms compared with O2·. In addition, H2O2 can diffuse freely through membranes similar to water, and it targets mainly free thiols. This oxidation is generally reversible, which means that H2O2 fulfills the requirements for a secondary messenger [1,6,7,12]. To prevent excessively high H2O2 concentrations and thus oxidative damage, the amount of H2O2 is strictly controlled by enzymes such as catalase or in addition by thiol scavengers [13].
The accumulation of intracellular H2O2 increases the risk of a Fenton reaction [14], which can cause major cell and tissue damage. In this reaction, H2O2 interacts with free iron (Fe2+) to produce highly reactive ·OH [10,15,16]. The majority of iron in the cell is present in the bound form as Fe3+, which is unreactive. However, there is always a small amount of free Fe2+ in the cell, the so-called labile iron pool. In the case of excessive amounts of H2O2, Fe2+ and H2O2 can react with each other. The resulting ·OH is extremely reactive [1]. The oxidation of cellular macromolecules, especially lipids, by ·OH leads to uncontrollable chain reactions and massive cell and tissue damage, followed by inflammatory reactions [10,16,17].
In addition to ROS, nitric oxide (NO) can be generated in mammalian cells by oxidation of one of the terminal guanidino nitrogen atoms of L-arginine. This reaction is catalyzed by the enzyme NO synthase (NOS) [18,19]. Nitric oxide itself is less reactive and generally not harmful. However, if NO molecules accumulate, they rapidly react with O2· to form highly detrimental peroxynitrite (ONOO) [3,7]. Peroxynitrite reacts with multiple substrate derivatives and can induce cellular damage, whereas NO itself functions as a second messenger [20]. In this review article, we focus on ROS and its physiological and pathophysiological functions.

2.2. Sources of ROS Generation

Among the many different cellular sources of ROS (Table 1), mitochondria and NADPHoxidases are the most important ones. In mitochondria, the electron transport chain (ETC) is responsible for the release of ROS (Figure 1a). The main site of univalent reduction of oxygen and thus the production of O2· is ubisemiquinone, a component of the ETC localized in the mitochondrial matrix [3,21,22,23]. In addition to this unintentional production of ROS via the ETC, ROS can also be generated in a controlled manner via complex I and complex III of the ETC. Thus, the mitochondrion can act as an oxidative signaling platform in many physiological settings, e.g., in the regulation of a T-cell immune response [7,24,25,26,27,28].
NADPHoxidases are multicomponent complexes that catalyze a one-electron reduction of oxygen by NADPH. Among the various NADPHoxidases within a cell, it is certainly the phagocytic NADPHoxidase, also called NADPHoxidase 2, that is most relevant to inducible ROS generation (Figure 1b) [29]. This plasma membrane-associated enzyme complex is best studied in phagocytes. However, it is also found in other cells, such as neutrophils, B lymphocytes, and dendritic cells. During the catalytic reaction, NADPHoxidase transports electrons across the plasma membrane to extracellular oxygen to form extracellular O2·. These O2· are rapidly converted to H2O2, which can freely diffuse across the plasma membrane and thus translocate back into the cell [1,30]. NADPHoxidase 1 is another crucial ROS-producing enzyme. It is expressed in intestinal epithelial cells and plays an important role in cell migration, differentiation, and wound healing, and it can be induced by gut microbiota [29,31]. Like NADPHoxidase 2, NADPHoxidase 1 is a multicomponent membrane complex producing O2·, which is immediately converted into H2O2. However, aberrant NADPHoxidase 1 activation or expression is involved in a growing number of diseases, including cancer [29,32].
Table 1. Cellular ROS sources influenced by gut microbiota.
Table 1. Cellular ROS sources influenced by gut microbiota.
Cellular CompartmentROS SourceROSMechanism of ActionROS-Related DiseaseRole of Gut Microbiota
Mitochondria Complex I respiratory chainO2·Cellular signaling, immune cell activation, energy metabolismInflammatory bowel disease (IBD) and colorectal cancer (CRC)Induced by bacterial antigens, regulated by bacterial metabolites [7,33]
MitochondriaComplex III respiratory chainO2·Cellular signaling, immune cell activation, energy metabolismIBD, CRCInduced by bacterial antigens, regulated by bacterial metabolites [7,33]
Plasma membrane, vesicular membranes (endoplasmatic reticulum, endosome and lysosome)NADPHoxidasesO2·Cellular signaling, host defense, immune cell activation, inflammation, oxidative burstInflammation, CRC, ileitis, and IBDInduced by bacterial antigens [34,35,36]
LysosomeMyeloperoxidaseH2O2Neutrophile activation pathogen defenseIBDInduced by bacterial antigens [37,38,39,40]
PeroxisomesFlavoproteins include acyl-CoA oxidases, urate oxidase, D-amino acid oxidase, D-aspartate oxidase, L-pipecolic acid oxidase, L-α-hydroxyacid oxidase, and polyamine oxidase H2O2Combat/oxidize bacterial pathogensIBD and CRCInduced by bacterial antigens and metabolites [37,41,42,43]
PeroxisomesXanthine oxidaseO2·/H2O2Combat/oxidize bacterial pathogens, inflammationInflammation, type II diabetesInduced by bacterial antigens and metabolites, purine degradation [37,41,42,43,44,45]
CytosolXanthine oxidaseO2·/H2O2InflammationInflammation, type II diabetes, goutPurine degradation [37,44,45]

2.3. Oxidative Stress

Under physiologic conditions, free radicals and non-radical reactive compounds are present in tissues in low amounts. Their levels depend on the ratio of their production and degradation by anti-oxidants and anti-oxidative enzymes. This creates a REDOX equilibrium that allows oxidative signaling and prevents oxidation-induced damage. A shift of this REDOX equilibrium toward enhanced ROS production can result in unwanted and often non-reversible oxidation of macromolecules such as lipids, proteins, and DNA referred to as oxidative stress. However, it should be emphasized that the transition from REDOX balance to oxidative stress is different for each tissue and even from cell to cell, as it depends on the cellular background [3,4,37,46].
Antioxidants are classified as non-enzymatic or enzymatic. The most important non-enzymatic antioxidants include the tri-peptide glutathione and the proteins thioredoxin 1 and thioredoxin 2 [3,13]. Glutathione is a powerful antioxidant that can scavenge radical and non-radical reactive compounds such as H2O2, nitrites, nitrates, and benzoates. Thioredoxins display cytoprotective effects in various cellular responses by removing H2O2 and thus regulating the activity of redox-sensitive transcription factors, which often control the antioxidant defense system [2,13,47].
The major components of the enzymatic oxidative defense include catalase (CAT), superoxide dismutases (SOD1/2/3), glutathione peroxidase (GPX), and glutathione reductase (GSR) [13,48]. In particular, catalase and SODs form the main enzymatic defense against oxidative stress. SODs catalyze the conversion of O2· to H2O2, whereas CAT converts H2O2 to water [13,49]. GPX is part of the glutathione–REDOX system and converts glutathione to its oxidized form, thereby reducing H2O2 to water, and lipid hydroperoxides to their corresponding stable alcohols [13]. The GPX reaction is coupled to GSR, which reduces and thereby recycles oxidized glutathione [13].

3. REDOX Regulation of the Immune System

The state of the REDOX equilibrium is crucial for the immune system. Oxidative signals determine whether an immune response is switched on or off. To generate these important oxidative signals, most immune cells undergo a metabolic shift from mitochondrial respiration to glycolysis. This shift enables the cells to initiate oxidative signals. In addition, ROS are also used as a defense mechanism against pathogens. Here much higher concentrations of ROS are released by the cells to oxidize and destroy pathogens.

3.1. Activation-Induced Oxidative Signaling in Immune Cells

Oxidative signals in immune cells are very important. They not only regulate the expression of cytokines/interleukins (IL) but also control the differentiation into different cellular subsets. Here we describe the mechanisms of ROS generation in the most relevant immune cells.

3.1.1. T-Cell Activation-Induced Oxidative Signaling

After T-cell receptor (TCR) stimulation, two signals are induced: (i) a calcium influx into the cytosol and (ii) initiation of an oxidative signal. The calcium influx leads to the activation of calcium-dependent transcription factors such as NF-AT or the activation of the neuronal nitric oxide synthase (nNOS) and the epithelial nitric oxide synthase (eNOS) [26]. However, the exact function of nNOS and iNOS in T-cell activation is still unclear [50,51,52]. The oxidative signal is initiated by a switch in the metabolism from mitochondrial respiration toward glycolysis comparable to the Warburg effect in cancer cells [24]. This results in a controlled release of O2· from mitochondria [24,26,28,53,54]. O2· are then converted to H2O2 by SODs [55], which then activate the REDOX-sensitive transcription factors NF-κB and AP-1. The oxidative signaling pathway can then be further significantly enhanced by an additional release of ROS by NADPHoxidase 2 [26,36]. NF-kappaB and AP-1 in concert with the calcium-dependent NF-AT constitute the minimal requirement for the induction of the expression of various cytokines or the CD95 death ligand (CD95L) and can thus control the induction and the termination of a T-cell immune response [7,24,25,26,56].
However, ROS play a crucial role not only in the activation of T cells but also in the CD4 T helper cell distribution. Thus, the activity, as well as the release of ROS from complex I, is a crucial factor for the differentiation of CD4+ cells to Th17 or T regulatory cells (Treg). It is not the function of the T-cell subsets that are affected but rather the differentiation of the different T-cell types [57,58]. Therefore, it can be stated that ROS and RNS make an important contribution to T-cell differentiation and activation.

3.1.2. B-Cell Activation-Induced Oxidative Signaling

Unlike T cells, where an oxidative signal is first generated by the mitochondria, which is then amplified and prolonged by additional ROS production through the phagocytic oxidase 2, ROS are first produced by phagocytic NADPHoxidase after B-cell stimulation, and then the oxidative signal is prolonged by mitochondrial ROS generation [26,59]. In the absence of the second mitochondria-produced ROS signal, B-cell activation is inefficient, and cell proliferation is reduced [59,60]. Thus, cell mitochondrial ROS signaling participates in the activation processes of both B and T lymphocytes [61].

3.1.3. Macrophages and Oxidative Signaling

Macrophages can be activated by commensal LPS, cytokines, pathogen-associated molecular patterns (PAMPs), damage-associated molecular patterns (DAMPs), and ROS. These warning signals induce transcription factors such as NF-κB, as well as PI3K and mTOR. Downstream pathways lead to the maintenance of macrophage activation and metabolic reprogramming. This shift from aerobic mitochondrial energy production to anaerobic glycolysis is essential in M1 macrophages for increased phagocytosis, increased production of inflammatory cytokines, and an even higher increase in ROS production [62,63,64,65,66,67].
In an environment with elevated levels of interleukin (IL)-4 and IL-13, M1 macrophages change their phenotype, become M2 macrophages, and develop anti-inflammatory and wound-healing properties [68,69]. During the shift from M1 to M2, the metabolism of macrophages switches back to mitochondrial energy production, and these cells become independent of oxidative signals [68,69]. In summary, inflammatory M1 macrophages are induced by ROS and produce ROS themselves, whereas anti-inflammatory M2 macrophages act independently of ROS. Of note, permanently elevated ROS levels and oxidative stress correlate with the induction of senescence [70,71,72] and may trigger cell death in macrophages [73,74].

3.1.4. Dendritic Cells and Oxidative Signaling

Dendritic cells are typical antigen-presenting cells. Under physiological conditions, tissue-resident dendritic cells migrate to draining lymph nodes and present self-antigens, inducing tolerance [75]. However, after pathogen invasion, dendritic cells are activated via toll-like receptors (TLR), migrate to the lymph nodes, and undergo diverse changes in function and phenotype. Resting dendritic cells derive their energy from mitochondrial respiration. Upon activation by TLR stimulation, dendritic cells switch to glycolysis [76,77,78]. The intermediates produced by glycolysis are then shunted into the pentose phosphate pathway, which promotes the production of NO. NO can react with O2· to form highly reactive ONOO, which in turn blocks the ETC and leads to an increase in ROS generation [79,80,81]. ROS production then has a decisive impact on antigen presentation and determines the activation of CD8+ and CD4+ T cells [82,83,84].

3.2. ROS, the Oxidative Burst, and the Inflammasome

Inflammation is primarily a host-induced defense against pathogens. The production of ROS is essential for the progression of the inflammatory reaction. Primarily, ROS are produced by cells of the immune system, mainly by antigen-presenting cells and polymorphonuclear neutrophils.
NADPHoxidases are certainly the main source of pathogen-induced ROS production. Currently, seven isoforms of NADPHoxidases are known (NADPHoxidase 1–5 and DUOX 1/2). NADPHoxidase 2 has the most important role in immune cells. Activation of ROS production induced by this NADPHoxidase 2 results in the release of high concentrations of O2·. For example, neutrophils can produce ~10 nmol/min O2· per million neutrophils during the oxidative burst to oxidize and kill pathogens [85,86]. NADPHoxidase 2 deficiency leads to severe diseases such as chronic granulomatous disease [87]. In epithelial cells, NADPHoxidase 1 is mainly responsible for ROS production. The amounts of ROS produced by epithelial cells are much lower compared with an oxidative burst. Epithelial cell-derived ROS are not to oxidize and destroy bacteria directly; they rather work as messenger molecules controlling proliferation and cellular inflammatory responses [88,89].
Mitochondria-derived ROS also play an important role as messenger molecules in inflammatory responses. Thus, LPS can induce the production of mitochondrial ROS via TLR signaling. The ROS production is then involved in inducing the activation of the pro-inflammatory proteins IL-1β, IL-6, and TNF [90]. In addition, mitochondria-derived ROS play a crucial role in the activation of the inflammasome [91]. The inflammasome is a cytosolic multi-protein complex that regulates the activation of inflammatory caspases (caspase-1 and caspase-12) (Figure 2). Three different types of inflammasomes have been described: NALP1, NALP3, and IPAF. Of these, the NALP3 inflammasome is REDOX sensitive [92,93]. The activity of the NALP3 inflammasome is regulated by the thioredoxin binding protein (TXNIP, VDUP-1). TXNIP is bound to thioredoxin under non-inflammatory conditions. In the presence of ROS, it is released and can bind to NALP3 and activate the inflammasome. Activation of Caspase-1 and Caspase-12 induces cleavage and activation of IL-1β and IL-18 [94].

4. Gut Microbiome and the REDOX Status

The only site in vertebrates where a continuous activation of the immune system occurs is the gut. In humans, the intestine represents the largest contact surface of the body with the environment. The intestinal mucosa has two opposite functions: on one hand, nutrients have to be absorbed, and on the other hand, the infiltration of digestive enzymes, pathogens, and also commensal bacteria into the mucosa and circulation has to be prevented. Thus, the epithelial barrier with its diverse components (mucus, tight monolayer of epithelial cells, and the intestinal immune system) plays an essential role in homeostatic mechanisms. Although the intestine is the habitat for billions of bacteria, it is noteworthy that only a few individuals develop inflammatory diseases because of these commensal bacteria. Thus, in the physiologic situation, some mechanisms effectively differentiate between resident bacteria and invading pathogens and adjust their reaction accordingly. In contrast to other lymphoid organs, the intestinal immune system is not exclusively activated in the case of an infection. Moreover, a continuous immune response takes place. In a healthy organism, there is homeostasis, i.e., a balance of pro- and anti-inflammatory mechanisms. Only in the case of an actual infection, a temporary inflammatory reaction is initiated [8]. However, chronic intestinal inflammatory processes, such as those associated with chronic inflammatory bowel diseases (IBD), develop due to a complex disturbance of the immunological homeostasis.
Since the terms “microbiome” and “microbiota” have frequently been used as synonyms, they were clearly defined in 2020: The microbiota denotes the collection of microbial organisms within a community in an animal host and refers to the taxonomy of the microorganisms. The microbiome is the collection of microorganisms and their genes living in a particular environment. Thus, the microbiome contains the microbiota, its “activity”, and the surrounding environmental conditions [95], including bacteria, viruses, archaea, and fungi. The human gut microbiota is involved in various protective, structural, and metabolic functions and plays a central role in gut homeostasis and health. Here we focus on the role of bacteria in REDOX regulation in the gastrointestinal tract.
The gastrointestinal tract is a habitat for more than 100 trillion microorganisms with at least 1000 different species of bacteria [96]. Only about one-third of the gut microbiota constitutes a “common core”, whereas two-thirds of gut bacteria differ between individuals [97] and thus represent a kind of intestinal fingerprint. While being sterile at birth, the gastrointestinal tract is colonized thereafter, and a stable microbiota develops during the first two years of life. However, changes may occur over the course of a lifetime due to external factors such as health, age, and lifestyle [98]. Bacteroidetes (~15–50%), firmicutes (~20–50%), actinobacteria (<5%), and proteobacteria (<10%) constitute the predominant phyla of commensal bacteria in the human gut [99].
In addition to this plethora of microorganisms, the luminal side of the gastrointestinal tract is exposed to bacterial metabolites as well as to dietary components. This amount of antigen is countered on the intestinal side by the epithelial cells on one hand and the immune cells in the lamina propria on the other. It has long been known that the gut microbiome and its related low-molecular-weight metabolites play an essential role in the maturation of the host immune system but also in the homeostatic processes [100]. As a consequence, serious changes in the gut microbiome are closely associated with the development of inflammatory diseases such as IBD [101,102]. Of note, the gut microbiome cannot only induce local inflammatory reactions but moreover can induce systemic inflammation and intervene in the regulation of the function of extra-intestinal organs such as the brain, liver, and skin by inducing the production of specific metabolites [103,104,105,106,107,108]. Conversely, non-intestinal diseases can influence the interaction of gut bacteria with the epithelium, e.g., via the gut–liver axis in patients with liver cirrhosis [109]. This mutual interaction between the liver and the gut is mediated by the portal vein, which enables the transport of products originating from the intestine to the liver and the transport of bile and metabolites from the liver to the gut via the bile ducts. The mucosal and epithelial barrier of the intestine is the functional structure that serves as a connection point for the interactions between the intestine and the liver, which also limits the systemic spread of microbes. Control of microbial communities is critical for maintaining homeostasis of the gut–liver axis. As a result of this bidirectional communication, the liver has an impact on gut microbial communities [110]. However, emerging evidence has shown that the interaction of gut bacteria with intestinal epithelial cells or immune cells also exerts protective effects by regulating the REDOX status and thus contributes to the homeostasis.

4.1. Gut Bacteria and Chronic Inflammation

IBD are chronic diseases of the gastrointestinal (GI) tract that impair the quality of life. Crohn’s disease (CD) and ulcerative colitis (UC) are the main forms of IBD. Although it is widely known that IBD is characterized by an inappropriate immune response to environmental changes and by alterations in the intestinal microbiota, the underlying mechanisms of inflammation remain elusive [111]. The recognition of a REDOX imbalance in colonic tissues due to ROS overproduction has linked these reactive molecules to the development and the progression of IBD [34].
Intestinal epithelial and immune cells use pattern recognition receptors (PPRs) such as Toll-like receptors (TLRs) and NOD (nucleotide-binding oligomerization domain-containing protein) proteins to detect bacteria and their metabolites. These receptors, therefore, provide constant communication between the microbiota and the host [34]. Recognition of bacteria by host cells can lead to the release of O2· by NADPHoxidases and dual oxidase 2 (DUOX2). Increased expressions of NADPHoxidases, including NADPHoxidase 1, NADPHoxidase 2, and DUOX2, are therefore considered genetic risk factors for IBD [112]. O2· derived from the NADPHoxidases is rapidly converted to H2O2 [34] and re-enters the intestinal epithelial cells. This leads to alterations in signal transduction [113] as well as the induction of inflammatory processes [114]. Altered cellular signaling induces the expression of inflammatory cytokines, which in turn further increase NADPH-oxidase-dependent ROS production. In addition, the inducible nitric oxide synthase (iNOS) is induced [115,116]. The generation of O2· by NADPHoxidase and NO by NOS results in the generation of highly reactive and detrimental ONOO. ONOO efficiently destroys bacteria but also oxidizes the plasma membrane of host cells, leading to the release of damage-associated molecular patterns (DAMPs) that amplify the inflammatory response [117]. Other immune cells such as leukocytes and monocytes are subsequently activated and further increase ROS accumulation [118]. This oxidative stress results in a chain reaction, which culminates in increased epithelial permeability [115,116]. It is noteworthy that oxidative stress can also be amplified directly by bacteria-produced ROS. For example, O2· is produced by bacteria such as E. coli [119,120]. In addition, bacteria, including Lactobacillus and Bifidobacterium [121,122], have been shown to produce H2O2 as a by-product of enzymatic reactions. In turn, bacterially produced H2O2 can increase inflammatory reactions and increase the permeability of the epithelium in IBD [123,124]. This increased permeability facilitates bacterial penetration into the lamina propria and contributes to sustained immune activation and ROS release, leading to an even more proinflammatory microenvironment and thus ROS-dependent perpetuation of the vicious cycle of chronic inflammation. In addition, the persistence of severe inflammatory processes in IBD has been shown to increase the risk of developing colitis-associated cancer [125]. This type of cancer differs from sporadic colorectal cancer in that it is chronic inflammation and an increased turnover of epithelial cells that drive tumor development. ROS produced during these inflammatory processes essentially contribute to the generation of dysplastic lesions [126,127]. Thus, it can be assumed that a regulation of the REDOX status in the gut has second-line effects on the development and progression of cancer.

4.2. Direct Effects of the Microbiome on the REDOX Status in the Gut

The human gut microbiota restricts the proliferation of pathogenic bacteria in the GI tract, activates the immune system, regulates nutrient utilization and host metabolism, and controls vitamin and enzyme production. The microbiota also produces short-chain fatty acids (SCFAs), ethanol, lactate, phenols, and succinate; degrades proteins and carbohydrates; and transforms bile acids [128,129].
Bacteria capable of producing ROS have developed specific mechanisms to resist an oxidative environment [130]. The ability to produce ROS, but also to counteract ROS, together with a variety of mechanisms to influence the REDOX status of the surrounding environment, makes bacteria key players in coordinating the intestinal REDOX equilibrium (Figure 3).
Generally, both commensals and pathogens can alter the production of ROS by host cells by modulating their mitochondrial activity or activation of NADPHoxidases [131]. In addition, during the degradation of sulfur-containing amino acids, pathogenic bacteria such as Salmonella and Escherichia coli can produce H2S, which impairs essential metabolic functions of colonic epithelial cells [105]. Increased levels of H2S can lead to a blockage of complex IV of the electron transport chain, resulting in epithelial damage, prevention of SCFA metabolism, and a disruption of the mucus barrier [33].
Importantly, commensal bacterial communities can also essentially contribute to REDOX homeostasis in the gut [37]. A brief overview is given in Figure 4.
It is not surprising that these include predominantly probiotics, which are defined as live nonpathogenic microorganisms that, when administered in adequate amounts, are beneficial for the health of the host [132,133,134]. Probiotics include strains of lactic acid bacilli, nonpathogenic strains of Escherichia coli such as Escherichia coli Nissle 1917, and Saccharomyces boulardii [135], among others. Probiotics are known for several beneficial effects on the gut in treating gastrointestinal disorders such as ulcerative colitis [136], pouchitis [137], irritable bowel syndrome, and Clostridioides difficile infections [138,139]. Even so, there is currently no agreement between the guidelines of the various international professional societies on the use of probiotics in the treatment of gastrointestinal diseases [136,140].
Of note, probiotics are also able to counteract high ROS levels by inducing anti-oxidative processes via several mechanisms [132]: Probiotics produce their own anti-oxidases such as SOD or catalase, and they also generate anti-oxidative metabolites such as folate and GSH. Moreover, probiotic lactic acid bacteria produce antioxidant metabolites such as exopolysaccharides, carotenoids, ferulic acid, or histamine, which reduce oxidized molecules and thereby contribute to the REDOX equilibrium [141]. Furthermore, probiotics affect host cells by inducing their anti-oxidative capacities (activation of host SOD and catalase and upregulation of the production of folate and GSH) and, simultaneously, dampening the activities of ROS-producing enzymes [37,132,141]. The latter actions are based on the regulation of several signaling pathways in the host cell, including Nrf2 (nuclear erythroid 2-related factor 2), SIRT, MAPK, and PKC [141].

4.3. Indirect Effects of the Microbiome on the Redox Status in the Gut

Besides direct intervention in the production of ROS and antioxidants, bacteria use a range of options to indirectly control the intestinal REDOX state. Commensal bacteria are producers of a set of metabolites, which can influence the REDOX status in the intestine: formyl-peptides, reactive nitrogen species (RNS), and SCFAs.
ROS are generally considered to be destructive pro-inflammatory molecules. However, they also trigger a number of important physiological functions, such as cell proliferation and migration. In addition, ROS also play an important role in angiogenesis [35].
Here it is crucial where and how much ROS is produced. NADPHoxidase 1 plays a key role in this context. NADPHoxidase 1 is expressed in epithelial cells and produces many times fewer ROS than NADPHoxidase 2 from immune cells. Some strains of Lactobacilli induce NADPHoxidase 1. This leads to reversible oxidation and inactivation of the tyrosine phosphatases LMW-PTPase and SHP-2 resulting in the activation of the focal adhesion kinase (FAK). This promotes focal adhesion and accelerates wound healing in the intestine [31,142,143]. Furthermore, the influence of ROS on proliferation and epithelial growth can be detected. It has been shown that ROS induced by commensal bacteria also lead to the inactivation of ERK phosphatase DUSP6, resulting in enhanced ERK phosphorylation and activation of the ERK signaling pathway [144,145]. Thus, it can be stated that bacterial induction of NADPHoxidase 1 plays an important role in the growth and differentiation of intestinal epithelial cells [31].
Commensal-derived formylated peptides bind to G protein receptors on immune cells such as macrophages and neutrophils, as well as epithelial cells leading to inflammatory processes and an enhanced ROS generation in the gut via activation of NADPHoxidases [105,146]. Moreover, Lactobacilli and Bifidobacteria as well as Streptococcus and Bacilli are capable of synthesizing NO [147].
Anaerobic fermentation of dietary fibers by bacteria produces the SCFAs acetate, propionate, and butyrate [102,105,148]. These metabolites have a variety of functions, ranging from energy supply for epithelial cells to effectors on neuronal development and other physiologic functions of organs via systemic circulation [105]. SCFAs have been shown to enhance tight junction proteins between endothelial cells [149], and intestinal epithelial cells [150,151,152]. Thus, they exert a direct protective effect on both the stabilization and the recovery of the intestinal epithelial barrier. Most importantly, SCFAs have been shown to modulate oxidative stress. Specifically, SCFAs can activate the antioxidant defense system via the Keap1-Nrf2 defense pathway [149,153,154]. Nrf2 is a key transcription factor of the cellular antioxidant defense by controlling more than 200 genes (Figure 5) [155]. By enhancing antioxidant defenses, SCFAs can reduce the mitochondrial damage caused by ROS and improve mitochondrial function. This protects the mitochondrial metabolism and enables a better energy supply (ATP) to the cells. In addition, the respiratory chain of the mitochondria is also protected from oxidation. Damage to the respiratory chain leads to an electron release, which leads to further ROS generation and starts a chain reaction of increased mitochondrial damage and ROS generation. Thus, SCFAs form a protective shield that defends against oxidative and mitochondrial stress.
Another non-negligible indirect effect of the microbiome on the REDOX status in the gut is the regulation of immune homeostasis. Commensal bacteria are essential regulators of intestinal immune homeostasis by maintaining the balance of pro- versus anti-inflammatory cytokine production by Th17 and Treg cells [102,156]. The probiotic Lactobacillus acidophililus has been shown to restore a homeostatic ratio of both Th17 and Treg cells and pro- and anti-inflammatory cytokines in a mouse model of colitis [157]. Concordantly, Clostridia and Bacteroides augmented anti-inflammatory responses by the induction of Treg cells [158,159,160]. Since an excess of pro-inflammatory cytokines leads to increased ROS production as part of the inflammatory response [161], restoring the balance between Th17 and Treg cells by specific bacteria builds a second line of defense against excess ROS production in the epithelium.
Vice versa, NADPHoxidases in the intestinal epithelium have an impact on the gut microbiome. A partial or complete inactivation of NADPHoxidases in the epithelium results in an altered H2O2 gradient in the mucus layer and, consequently, a change in the composition of the microbiome [34]. Interestingly, the inactivation of Nox1–4 together or individually increased the abundance of facultative aerobic bacteria, especially Firmicutes [162]. This example impressively demonstrates how closely epithelial REDOX mechanisms and the intestinal microbiome are interlinked, so that insufficient ROS production can be compensated by the expansion of H2O2-producing bacteria [34].

4.4. Therapeutic Interventions Reconstituting a Physiologic REDOX State in the Gut

Given the major influence of the microbiome on the intestinal REDOX status, it is reasonable to consider how the ROS balance can be specifically supported by bacteria. Thus, enhancing antioxidant-producing bacteria will shift the ROS ratio. Restoring the REDOX equilibrium in the gastrointestinal tract, therefore, involves treatment strategies to enhance the microbiome exerting direct or indirect antioxidative effects as depicted above. Thus, excessive ROS production as it occurs in several gastrointestinal diseases such as IBD, intestinal infections, ischemic damage, and colorectal cancer [37] might be attenuated by the administration of specific antioxidant-producing bacteria. Again, it is the group of probiotics that appear of special interest in this context.
Firmicutes such as Lactobacillus, Bifidobacterium, and Butyricicoccus are known to exert protective and anti-inflammatory effects on the intestinal epithelium [34]. Bacteria of the genus Butyricicoccus are reduced in the stool of patients with inflammatory bowel disease, and administration of Butyricicoccus pullicaecorum reduced mucosal lesions and inflammation in a rat model of colitis [163].
In addition to the long-known protective probiotics Lactobacillus and Bifidobacterium, “novel” probiotics are discussed, including anaerobic SCFA-producing strains [37]. As outlined above, SCFAs play a crucial role in activating the antioxidant defense system and exert several additional protective functions to maintain intestinal homeostasis. Therefore, an increase in SCFA-producing bacteria not only enhances antioxidant mechanisms but also contributes to the stabilization of the epithelial barrier and thus to reduced inflammatory responses. One candidate might be Faecalibacterium prausnitzii, which is reduced in samples from patients with gastrointestinal diseases and metabolic disorders such as IBD, irritable bowel syndrome, colorectal cancer, obesity, and celiac disease [164]. In a recent study, Faecalibacterium prausnitzii not only enhanced SCFA production but also had ameliorating effects on non-alcoholic fatty liver disease in mice [165]. Another option to support the growth of SCFA-producing bacteria is to supplement the diet with prebiotics. Prebiotics are defined as non-digestible fibers and other dietary compounds such as glycans or non-digestible carbohydrates that are selectively utilized by beneficial microorganisms in the gut. Thus, providing specifically tailored dietary compounds will create a growth advantage for SCFA-producing microbiota and enhance antioxidative processes without the need to culture beneficial anaerobic intestinal bacteria in vitro.
Last but not least, every anti-inflammatory effect of beneficial bacteria will have a second-line effect on ROS production in the gut. Concordantly, enrichment with probiotics will (i) shift the composition of the microbiome, (ii) reduce pro-inflammatory reactions by enhancing Treg responses, and (iii) reduce ROS levels due to diminished inflammation. Faecalibacterium prausnitzii, which is reduced in patients with IBD [166], revealed anti-inflammatory effects in Crohn’s disease [102,167]. Moreover, microbial products such as polysaccharides from Bacteroides fragilis as well as protein compounds from Faecalibacterium prausnitzii have been shown to induce IL-10-producing Treg in mice [102,168].
Taken together, both a dysbiosis of the gut microbiome and a REDOX imbalance are involved in the pathogenesis of a range of diseases, i.e., IBD [34,37], non-alcoholic fatty liver disease (NAFLD) [169], neurodegenerative disorders [103], inflammatory skin disorders [170], and cancer [171]. Therefore, regulating gut immunity by optimizing its microbiome and REDOX status appears to be quite an attractive novel option to treat intestinal or gut-associated diseases.

5. Conclusions

ROS production in the gut is a double-edged sword: on one hand, it is an indispensable mechanism in the defense against pathogens and mucosal healing. On the other hand, excessive ROS production can have detrimental effects on mucosal integrity and epithelial barrier function.
The effects of ROS production are a matter of dose, which is also true for pro- and anti-inflammatory mechanisms via cytokines, T-cell populations, or responses to bacterial components in the gut [172,173]. Finely adjusted ROS production has vital functions as it helps to defend against pathogens and induce repair mechanisms. Exaggerated ROS levels in the context of chronic inflammation can cause severe tissue damage if they are not counteracted by anti-oxidative mechanisms.
Oxidative stress due to exaggerated ROS production is a vital process in the body, but chronic exposure can lead to the oxidation of biomolecules and activation of inflammatory signaling pathways, resulting in genomic instability and the dysregulation of gene and protein expression and tumor initiation or cancer cell survival [127,174]. In addition, chronic intestinal diseases such as IBD increase the risk of tumor development by the promotion of cell proliferation and angiogenesis [127,175]. A recent study indicated that a “healthy” gut microbiome contains bacteria suppressing tumor growth by antioxidative metabolites [176]. Thus, an ROS-balanced microbiome can also contribute to tumor suppression.
The constant interaction of the host with its microbiome ensures the maintenance of the highly complex ecosystem in the gastrointestinal tract. This principle of host–microbiome interaction includes the regulation of the REDOX equilibrium. To take advantage of this interaction, oxidative conditions in certain intestinal diseases can be counteracted by the enrichment of bacteria with enhanced anti-oxidant activity. Thus, the modulation of the individual microbiome will be an important issue regarding future therapy for gastrointestinal disorders, cancer, and other diseases.

Author Contributions

Conceptualization, C.K. and K.G.; formal analysis, J.B.; writing—original draft preparation, C.K., K.G., S.S., D.T. and M.M. (Marlen Michalski); writing—review and editing, C.K., K.G. and M.M. (Martina Müller); visualization, K.G., C.K., D.T. and M.M. (Marlen Michalski); supervision, K.G. and M.M. (Martina Müller); funding acquisition, K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Wilhelm-Sander-Stiftung (2022.096.1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank all the staff of the Department of Internal Medicine I for their support. Furthermore, we would like to thank the Wilhelm-Sander-Stiftung for the support of Karsten Gülow and Deniz Tümen (2022.096.1).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Reth, M. Hydrogen peroxide as second messenger in lymphocyte activation. Nat. Immunol. 2002, 3, 1129–1134. [Google Scholar] [CrossRef] [PubMed]
  2. Schroeder, A.; Warnken, U.; Röth, D.; Klika, K.D.; Vobis, D.; Barnert, A.; Bujupi, F.; Oberacker, T.; Schnölzer, M.; Nicolay, J.P.; et al. Targeting Thioredoxin-1 by dimethyl fumarate induces ripoptosome-mediated cell death. Sci. Rep. 2017, 7, 43168. [Google Scholar] [CrossRef]
  3. Dröge, W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002, 82, 47–95. [Google Scholar] [CrossRef] [PubMed]
  4. Flohé, L. Looking Back at the Early Stages of Redox Biology. Antioxidants 2020, 9, 1254. [Google Scholar] [CrossRef] [PubMed]
  5. Zuo, J.; Zhang, Z.; Luo, M.; Zhou, L.; Nice, E.C.; Zhang, W.; Wang, C.; Huang, C. Redox signaling at the crossroads of human health and disease. MedComm 2022, 3, e127. [Google Scholar] [CrossRef]
  6. Andrés, C.M.C.; La Pérez de Lastra, J.M.; Andrés Juan, C.; Plou, F.J.; Pérez-Lebeña, E. Superoxide Anion Chemistry-Its Role at the Core of the Innate Immunity. Int. J. Mol. Sci. 2023, 24, 1841. [Google Scholar] [CrossRef] [PubMed]
  7. Kamiński, M.M.; Röth, D.; Krammer, P.H.; Gülow, K. Mitochondria as oxidative signaling organelles in T-cell activa-tion: Physiological role and pathological implications. Arch. Immunol. Ther. Exp. 2013, 61, 367–384. [Google Scholar] [CrossRef] [PubMed]
  8. De Vos, W.M.; Tilg, H.; van Hul, M.; Cani, P.D. Gut microbiome and health: Mechanistic insights. Gut 2022, 71, 1020–1032. [Google Scholar] [CrossRef]
  9. McCord, J.M.; Fridovich, I. Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 1969, 244, 6049–6055. [Google Scholar] [CrossRef]
  10. Birben, E.; Sahiner, U.M.; Sackesen, C.; Erzurum, S.; Kalayci, O. Oxidative stress and antioxidant defense. World Allergy Organ. J. 2012, 5, 9–19. [Google Scholar] [CrossRef]
  11. Brand, M.D. Riding the tiger—Physiological and pathological effects of superoxide and hydrogen peroxide generated in the mitochondrial matrix. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 592–661. [Google Scholar] [CrossRef] [PubMed]
  12. Sies, H.; Jones, D.P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383. [Google Scholar] [CrossRef] [PubMed]
  13. Vona, R.; Pallotta, L.; Cappelletti, M.; Severi, C.; Matarrese, P. The Impact of Oxidative Stress in Human Pathology: Focus on Gastrointestinal Disorders. Antioxidants 2021, 10, 201. [Google Scholar] [CrossRef] [PubMed]
  14. Jomova, K.; Makova, M.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Rhodes, C.J.; Valko, M. Essential metals in health and disease. Chem. Biol. Interact. 2022, 367, 110173. [Google Scholar] [CrossRef] [PubMed]
  15. Torti, F.M.; Torti, S.V. Regulation of ferritin genes and protein. Blood 2002, 99, 3505–3516. [Google Scholar] [CrossRef]
  16. Kiessling, M.K.; Klemke, C.D.; Kaminski, M.M.; Galani, I.E.; Krammer, P.H.; Gülow, K. Inhibition of constitutively activated nuclear factor-kappaB induces reactive oxygen species- and iron-dependent cell death in cutaneous T-cell lymphoma. Cancer Res. 2009, 69, 2365–2374. [Google Scholar] [CrossRef]
  17. Davies, M.J. The oxidative environment and protein damage. Biochim. Biophys. Acta 2005, 1703, 93–109. [Google Scholar] [CrossRef]
  18. Cinelli, M.A.; Do, H.T.; Miley, G.P.; Silverman, R.B. Inducible nitric oxide synthase: Regulation, structure, and inhibition. Med. Res. Rev. 2020, 40, 158–189. [Google Scholar] [CrossRef]
  19. Palmer, R.M.; Rees, D.D.; Ashton, D.S.; Moncada, S. L-arginine is the physiological precursor for the formation of nitric oxide in endothelium-dependent relaxation. Biochem. Biophys. Res. Commun. 1988, 153, 1251–1256. [Google Scholar] [CrossRef]
  20. Radi, R. Oxygen radicals, nitric oxide, and peroxynitrite: Redox pathways in molecular medicine. Proc. Natl. Acad. Sci. USA 2018, 115, 5839–5848. [Google Scholar] [CrossRef]
  21. Boveris, A.; Cadenas, E.; Stoppani, A.O. Role of ubiquinone in the mitochondrial generation of hydrogen peroxide. Biochem. J. 1976, 156, 435–444. [Google Scholar] [CrossRef] [PubMed]
  22. Boveris, A.; Chance, B. The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem. J. 1973, 134, 707–716. [Google Scholar] [CrossRef] [PubMed]
  23. Loschen, G.; Azzi, A.; Richter, C.; Flohé, L. Superoxide radicals as precursors of mitochondrial hydrogen peroxide. FEBS Lett. 1974, 42, 68–72. [Google Scholar] [CrossRef] [PubMed]
  24. Kamiński, M.M.; Sauer, S.W.; Kamiński, M.; Opp, S.; Ruppert, T.; Grigaravičius, P.; Grudnik, P.; Gröne, H.-J.; Krammer, P.H.; Gülow, K. T cell activation is driven by an ADP-dependent glucokinase linking enhanced glycolysis with mitochondrial reactive oxygen species generation. Cell Rep. 2012, 2, 1300–1315. [Google Scholar] [CrossRef] [PubMed]
  25. Kaminski, M.M.; Sauer, S.W.; Klemke, C.-D.; Süss, D.; Okun, J.G.; Krammer, P.H.; Gülow, K. Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: Mechanism of ciprofloxa-cin-mediated immunosuppression. J. Immunol. 2010, 184, 4827–4841. [Google Scholar] [CrossRef]
  26. Kaminski, M.; Kiessling, M.; Süss, D.; Krammer, P.H.; Gülow, K. Novel role for mitochondria: Protein kinase Cθ-dependent oxidative signaling organelles in activation-induced T-cell death. Mol. Cell. Biol. 2007, 27, 3625–3639. [Google Scholar] [CrossRef]
  27. Weinberg, S.E.; Singer, B.D.; Steinert, E.M.; Martinez, C.A.; Mehta, M.M.; Martínez-Reyes, I.; Gao, P.; Helmin, K.A.; Abdala-Valencia, H.; Sena, L.A.; et al. Mitochondrial complex III is essential for suppressive function of regulatory T cells. Nature 2019, 565, 495–499. [Google Scholar] [CrossRef]
  28. Sena, L.A.; Li, S.; Jairaman, A.; Prakriya, M.; Ezponda, T.; Hildeman, D.A.; Wang, C.-R.; Schumacker, P.T.; Licht, J.D.; Perlman, H.; et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity 2013, 38, 225–236. [Google Scholar] [CrossRef]
  29. Pecchillo Cimmino, T.; Ammendola, R.; Cattaneo, F.; Esposito, G. NOX Dependent ROS Generation and Cell Metabolism. Int. J. Mol. Sci. 2023, 24, 2086. [Google Scholar] [CrossRef]
  30. Di Marzo, N.; Chisci, E.; Giovannoni, R. The Role of Hydrogen Peroxide in Redox-Dependent Signaling: Homeo-static and Pathological Responses in Mammalian Cells. Cells 2018, 7, 156. [Google Scholar] [CrossRef]
  31. Neish, A.S. Redox signaling mediated by the gut microbiota. Free Radic. Res. 2013, 47, 950–957. [Google Scholar] [CrossRef] [PubMed]
  32. Bedard, K.; Krause, K.-H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef] [PubMed]
  33. Jackson, D.N.; Theiss, A.L. Gut bacteria signaling to mitochondria in intestinal inflammation and cancer. Gut Microbes 2020, 11, 285–304. [Google Scholar] [CrossRef]
  34. Aviello, G.; Knaus, U.G. NADPH oxidases and ROS signaling in the gastrointestinal tract. Mucosal Immunol. 2018, 11, 1011–1023. [Google Scholar] [CrossRef]
  35. Vermot, A.; Petit-Härtlein, I.; Smith, S.M.E.; Fieschi, F. NADPH Oxidases (NOX): An Overview from Discovery, Molecular Mechanisms to Physiology and Pathology. Antioxidants 2021, 10, 890. [Google Scholar] [CrossRef] [PubMed]
  36. Jackson, S.H.; Devadas, S.; Kwon, J.; Pinto, L.A.; Williams, M.S. T cells express a phagocyte-type NADPH oxidase that is activated after T cell receptor stimulation. Nat. Immunol. 2004, 5, 818–827. [Google Scholar] [CrossRef] [PubMed]
  37. Singh, V.; Ahlawat, S.; Mohan, H.; Gill, S.S.; Sharma, K.K. Balancing reactive oxygen species generation by rebooting gut microbiota. J. Appl. Microbiol. 2022, 132, 4112–4129. [Google Scholar] [CrossRef]
  38. Aratani, Y. Myeloperoxidase: Its role for host defense, inflammation, and neutrophil function. Arch. Biochem. Biophys. 2018, 640, 47–52. [Google Scholar] [CrossRef]
  39. Chami, B.; Martin, N.J.J.; Dennis, J.M.; Witting, P.K. Myeloperoxidase in the inflamed colon: A novel target for treating inflammatory bowel disease. Arch. Biochem. Biophys. 2018, 645, 61–71. [Google Scholar] [CrossRef]
  40. Winterbourn, C.C.; Kettle, A.J.; Hampton, M.B. Reactive Oxygen Species and Neutrophil Function. Annu. Rev. Biochem. 2016, 85, 765–792. [Google Scholar] [CrossRef]
  41. Sandalio, L.M.; Rodríguez-Serrano, M.; Romero-Puertas, M.C.; Del Río, L.A. Role of peroxisomes as a source of reactive oxygen species (ROS) signaling molecules. Subcell. Biochem. 2013, 69, 231–255. [Google Scholar] [CrossRef] [PubMed]
  42. Fransen, M.; Nordgren, M.; Wang, B.; Apanasets, O. Role of peroxisomes in ROS/RNS-metabolism: Implications for human disease. Biochim. Biophys. Acta 2012, 1822, 1363–1373. [Google Scholar] [CrossRef] [PubMed]
  43. Di Cara, F.; Bülow, M.H.; Simmonds, A.J.; Rachubinski, R.A. Dysfunctional peroxisomes compromise gut structure and host defense by increased cell death and Tor-dependent autophagy. Mol. Biol. Cell 2018, 29, 2766–2783. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Z.; Li, Y.; Liao, W.; Huang, J.; Liu, Y.; Li, Z.; Tang, J. Gut microbiota remodeling: A promising therapeutic strategy to confront hyperuricemia and gout. Front. Cell. Infect. Microbiol. 2022, 12, 935723. [Google Scholar] [CrossRef]
  45. Yu, Y.; Liu, Q.; Li, H.; Wen, C.; He, Z. Alterations of the Gut Microbiome Associated with the Treatment of Hyperuricaemia in Male Rats. Front. Microbiol. 2018, 9, 2233. [Google Scholar] [CrossRef] [PubMed]
  46. Sies, H. (Ed.) Oxidative Stress: Introductory Remarks. In Oxidative Stress; Academic Press: London, UK, 1985; pp. 1–8. ISBN 9780126427608. [Google Scholar]
  47. Oberacker, T.; Bajorat, J.; Ziola, S.; Schroeder, A.; Röth, D.; Kastl, L.; Edgar, B.A.; Wagner, W.; Gülow, K.; Krammer, P.H. Enhanced expression of thioredoxin-interacting-protein regulates oxidative DNA damage and aging. FEBS Lett. 2018, 592, 2297–2307. [Google Scholar] [CrossRef]
  48. Rahman, K. Studies on free radicals, antioxidants, and co-factors. Clin. Interv. Aging. 2007, 2, 219–236. [Google Scholar]
  49. Schrader, M.; Fahimi, H.D. Peroxisomes and oxidative stress. Biochim. Biophys. Acta 2006, 1763, 1755–1766. [Google Scholar] [CrossRef]
  50. Gergely, P.; Niland, B.; Gonchoroff, N.; Pullmann, R.; Phillips, P.E.; Perl, A. Persistent mitochondrial hyperpolarization, increased reactive oxygen intermediate production, and cytoplasmic alkalinization characterize altered IL-10 signaling in patients with systemic lupus erythematosus. J. Immunol. 2002, 169, 1092–1101. [Google Scholar] [CrossRef]
  51. Nagy, G.; Koncz, A.; Perl, A. T cell activation-induced mitochondrial hyperpolarization is mediated by Ca2+- and redox-dependent production of nitric oxide. J. Immunol. 2003, 171, 5188–5197. [Google Scholar] [CrossRef]
  52. Williams, M.S.; Noguchi, S.; Henkart, P.A.; Osawa, Y. Nitric oxide synthase plays a signaling role in TCR-triggered apoptotic death. J. Immunol. 1998, 161, 6526–6531. [Google Scholar] [CrossRef] [PubMed]
  53. Sena, L.A.; Chandel, N.S. Physiological roles of mitochondrial reactive oxygen species. Mol. Cell 2012, 48, 158–167. [Google Scholar] [CrossRef] [PubMed]
  54. Röth, D.; Krammer, P.H.; Gülow, K. Dynamin related protein 1-dependent mitochondrial fission regulates oxidative signalling in T cells. FEBS Lett. 2014, 588, 1749–1754. [Google Scholar] [CrossRef]
  55. Kamiński, M.M.; Röth, D.; Sass, S.; Sauer, S.W.; Krammer, P.H.; Gülow, K. Manganese superoxide dismutase: A regulator of T cell activation-induced oxidative signaling and cell death. Biochim. Biophys. Acta 2012, 1823, 1041–1052. [Google Scholar] [CrossRef] [PubMed]
  56. Devadas, S.; Zaritskaya, L.; Rhee, S.G.; Oberley, L.; Williams, M.S. Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulation: Selective regulation of mitogen-activated protein kinase activation and fas ligand expression. J. Exp. Med. 2002, 195, 59–70. [Google Scholar] [CrossRef] [PubMed]
  57. Ozay, E.I.; Sherman, H.L.; Mello, V.; Trombley, G.; Lerman, A.; Tew, G.N.; Yadava, N.; Minter, L.M. Rotenone Treatment Reveals a Role for Electron Transport Complex I in the Subcellular Localization of Key Transcriptional Regulators During T Helper Cell Differentiation. Front. Immunol. 2018, 9, 1284. [Google Scholar] [CrossRef]
  58. Yin, M.; O’Neill, L.A.J. The role of the electron transport chain in immunity. FASEB J. 2021, 35, e21974. [Google Scholar] [CrossRef]
  59. Wheeler, M.L.; DeFranco, A.L. Prolonged production of reactive oxygen species in response to B cell receptor stimulation promotes B cell activation and proliferation. J. Immunol. 2012, 189, 4405–4416. [Google Scholar] [CrossRef]
  60. Waters, L.R.; Ahsan, F.M.; Wolf, D.M.; Shirihai, O.; Teitell, M.A. Initial B Cell Activation Induces Metabolic Reprogramming and Mitochondrial Remodeling. iScience 2018, 5, 99–109. [Google Scholar] [CrossRef]
  61. Bassoy, E.Y.; Walch, M.; Martinvalet, D. Reactive Oxygen Species: Do They Play a Role in Adaptive Immunity? Front. Immunol. 2021, 12, 755856. [Google Scholar] [CrossRef]
  62. Jha, A.K.; Huang, S.C.-C.; Sergushichev, A.; Lampropoulou, V.; Ivanova, Y.; Loginicheva, E.; Chmielewski, K.; Stew-art, K.M.; Ashall, J.; Everts, B.; et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity 2015, 42, 419–430. [Google Scholar] [CrossRef]
  63. Feingold, K.R.; Shigenaga, J.K.; Kazemi, M.R.; McDonald, C.M.; Patzek, S.M.; Cross, A.S.; Moser, A.; Grunfeld, C. Mechanisms of triglyceride accumulation in activated macrophages. J. Leukoc. Biol. 2012, 92, 829–839. [Google Scholar] [CrossRef]
  64. Van Uden, P.; Kenneth, N.S.; Rocha, S. Regulation of hypoxia-inducible factor-1alpha by NF-kappaB. Biochem. J. 2008, 412, 477–484. [Google Scholar] [CrossRef]
  65. Freemerman, A.J.; Johnson, A.R.; Sacks, G.N.; Milner, J.J.; Kirk, E.L.; Troester, M.A.; Macintyre, A.N.; Go-raksha-Hicks, P.; Rathmell, J.C.; Makowski, L. Metabolic reprogramming of macrophages: Glucose transporter 1 (GLUT1)-mediated glucose metabolism drives a proinflammatory phenotype. J. Biol. Chem. 2014, 289, 7884–7896. [Google Scholar] [CrossRef]
  66. Wang, T.; Liu, H.; Lian, G.; Zhang, S.-Y.; Wang, X.; Jiang, C. HIF1α-Induced Glycolysis Metabolism Is Essential to the Activation of Inflammatory Macrophages. Mediat. Inflamm. 2017, 2017, 9029327. [Google Scholar] [CrossRef]
  67. Pavlou, S.; Wang, L.; Xu, H.; Chen, M. Higher phagocytic activity of thioglycollate-elicited peritoneal macrophages is related to metabolic status of the cells. J. Inflamm. 2017, 14, 4. [Google Scholar] [CrossRef]
  68. Van den Bossche, J.; O’Neill, L.A.; Menon, D. Macrophage Immunometabolism: Where Are We (Going)? Trends Immunol. 2017, 38, 395–406. [Google Scholar] [CrossRef]
  69. Boscá, L.; González-Ramos, S.; Prieto, P.; Fernández-Velasco, M.; Mojena, M.; Martín-Sanz, P.; Alemany, S. Metabolic signatures linked to macrophage polarization: From glucose metabolism to oxidative phosphorylation. Biochem. Soc. Trans. 2015, 43, 740–744. [Google Scholar] [CrossRef]
  70. Burova, E.; Borodkina, A.; Shatrova, A.; Nikolsky, N. Sublethal oxidative stress induces the premature senescence of human mesenchymal stem cells derived from endometrium. Oxid. Med. Cell. Longev. 2013, 2013, 474931. [Google Scholar] [CrossRef]
  71. Li, H.; Luo, Y.-F.; Wang, Y.-S.; Yang, Q.; Xiao, Y.-L.; Cai, H.-R.; Xie, C.-M. Using ROS as a Second Messenger, NADPH Oxidase 2 Mediates Macrophage Senescence via Interaction with NF-κB during Pseudomonas aeruginosa Infection. Oxid. Med. Cell. Longev. 2018, 2018, 9741838. [Google Scholar] [CrossRef]
  72. Elder, S.S.; Emmerson, E. Senescent cells and macrophages: Key players for regeneration? Open Biol. 2020, 10, 200309. [Google Scholar] [CrossRef]
  73. Brown, L.A.S.; Ping, X.-D.; Harris, F.L.; Gauthier, T.W. Glutathione availability modulates alveolar macrophage function in the chronic ethanol-fed rat. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 292, L824–L832. [Google Scholar] [CrossRef]
  74. Fitzpatrick, A.M.; Teague, W.G.; Burwell, L.; Brown, M.S.; Brown, L.A.S. Glutathione oxidation is associated with airway macrophage functional impairment in children with severe asthma. Pediatr. Res. 2011, 69, 154–159. [Google Scholar] [CrossRef] [PubMed]
  75. Gallo, P.M.; Gallucci, S. The dendritic cell response to classic, emerging, and homeostatic danger signals. Implications for autoimmunity. Front. Immunol. 2013, 4, 138. [Google Scholar] [CrossRef]
  76. Thwe, P.M.; Pelgrom, L.R.; Cooper, R.; Beauchamp, S.; Reisz, J.A.; D’Alessandro, A.; Everts, B.; Amiel, E. Cell-Intrinsic Glycogen Metabolism Supports Early Glycolytic Reprogramming Required for Dendritic Cell Immune Responses. Cell Metab. 2017, 26, 558–567.e5. [Google Scholar] [CrossRef]
  77. Krawczyk, C.M.; Holowka, T.; Sun, J.; Blagih, J.; Amiel, E.; DeBerardinis, R.J.; Cross, J.R.; Jung, E.; Thompson, C.B.; Jones, R.G.; et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 2010, 115, 4742–4749. [Google Scholar] [CrossRef]
  78. Sukhbaatar, N.; Hengstschläger, M.; Weichhart, T. mTOR-Mediated Regulation of Dendritic Cell Differentiation and Function. Trends Immunol. 2016, 37, 778–789. [Google Scholar] [CrossRef]
  79. Everts, B.; Amiel, E.; van der Windt, G.J.W.; Freitas, T.C.; Chott, R.; Yarasheski, K.E.; Pearce, E.L.; Pearce, E.J. Commitment to glycolysis sustains survival of NO-producing inflammatory dendritic cells. Blood 2012, 120, 1422–1431. [Google Scholar] [CrossRef]
  80. Amiel, E.; Everts, B.; Fritz, D.; Beauchamp, S.; Ge, B.; Pearce, E.L.; Pearce, E.J. Mechanistic target of rapamycin inhibition extends cellular lifespan in dendritic cells by preserving mitochondrial function. J. Immunol. 2014, 193, 2821–2830. [Google Scholar] [CrossRef]
  81. Morris, G.; Gevezova, M.; Sarafian, V.; Maes, M. Redox regulation of the immune response. Cell. Mol. Immunol. 2022, 19, 1079–1101. [Google Scholar] [CrossRef]
  82. Mantegazza, A.R.; Savina, A.; Vermeulen, M.; Pérez, L.; Geffner, J.; Hermine, O.; Rosenzweig, S.D.; Faure, F.; Amigorena, S. NADPH oxidase controls phagosomal pH and antigen cross-presentation in human dendritic cells. Blood 2008, 112, 4712–4722. [Google Scholar] [CrossRef]
  83. Paardekooper, L.M.; Dingjan, I.; Linders, P.T.A.; Staal, A.H.J.; Cristescu, S.M.; Verberk, W.C.E.P.; van den Bogaart, G. Human Monocyte-Derived Dendritic Cells Produce Millimolar Concentrations of ROS in Phagosomes Per Second. Front. Immunol. 2019, 10, 1216. [Google Scholar] [CrossRef]
  84. Oberkampf, M.; Guillerey, C.; Mouriès, J.; Rosenbaum, P.; Fayolle, C.; Bobard, A.; Savina, A.; Ogier-Denis, E.; Enninga, J.; Amigorena, S.; et al. Mitochondrial reactive oxygen species regulate the induction of CD8+ T cells by plasmacytoid dendritic cells. Nat. Commun. 2018, 9, 2241. [Google Scholar] [CrossRef]
  85. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive oxygen species in inflammation and tissue injury. Antioxid. Redox Signal. 2014, 20, 1126–1167. [Google Scholar] [CrossRef]
  86. Babior, B.M. NADPH oxidase: An update. Blood 1999, 93, 1464–1476. [Google Scholar] [CrossRef]
  87. Nunoi, H.; Nakamura, H.; Nishimura, T.; Matsukura, M. Recent topics and advanced therapies in chronic granulomatous disease. Hum. Cell 2023, 36, 515–527. [Google Scholar] [CrossRef]
  88. Rokutan, K.; Kawahara, T.; Kuwano, Y.; Tominaga, K.; Sekiyama, A.; Teshima-Kondo, S. NADPH oxidases in the gastrointestinal tract: A potential role of Nox1 in innate immune response and carcinogenesis. Antioxid. Redox Signal. 2006, 8, 1573–1582. [Google Scholar] [CrossRef]
  89. van der Post, S.; Birchenough, G.M.H.; Held, J.M. NOX1-dependent redox signaling potentiates colonic stem cell proliferation to adapt to the intestinal microbiota by linking EGFR and TLR activation. Cell Rep. 2021, 35, 108949. [Google Scholar] [CrossRef]
  90. Bulua, A.C.; Simon, A.; Maddipati, R.; Pelletier, M.; Park, H.; Kim, K.-Y.; Sack, M.N.; Kastner, D.L.; Siegel, R.M. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J. Exp. Med. 2011, 208, 519–533. [Google Scholar] [CrossRef]
  91. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef]
  92. Dostert, C.; Pétrilli, V.; van Bruggen, R.; Steele, C.; Mossman, B.T.; Tschopp, J. Innate immune activation through Nalp3 inflammasome sensing of asbestos and silica. Science 2008, 320, 674–677. [Google Scholar] [CrossRef]
  93. Schroder, K.; Tschopp, J. The inflammasomes. Cell 2010, 140, 821–832. [Google Scholar] [CrossRef]
  94. Zhou, R.; Tardivel, A.; Thorens, B.; Choi, I.; Tschopp, J. Thioredoxin-interacting protein links oxidative stress to inflammasome activation. Nat. Immunol. 2010, 11, 136–140. [Google Scholar] [CrossRef]
  95. Berg, G.; Rybakova, D.; Fischer, D.; Cernava, T.; Vergès, M.-C.C.; Charles, T.; Chen, X.; Cocolin, L.; Eversole, K.; Corral, G.H.; et al. Microbiome definition re-visited: Old concepts and new challenges. Microbiome 2020, 8, 103. [Google Scholar] [CrossRef]
  96. Bermon, S.; Petriz, B.; Kajėnienė, A.; Prestes, J.; Castell, L.; Franco, O.L. The microbiota: An exercise immunology perspective. Exerc. Immunol. Rev. 2015, 21, 70–79. [Google Scholar] [PubMed]
  97. Qin, J.; Li, R.; Raes, J.; Arumugam, M.; Burgdorf, K.S.; Manichanh, C.; Nielsen, T.; Pons, N.; Levenez, F.; Yamada, T.; et al. A human gut microbial gene catalogue established by metagenomic sequencing. Nature 2010, 464, 59–65. [Google Scholar] [CrossRef]
  98. Cho, I.; Blaser, M.J. The human microbiome: At the interface of health and disease. Nature reviews. Genetics 2012, 13, 260–270. [Google Scholar] [CrossRef]
  99. Yilmaz, B.; Juillerat, P.; Øyås, O.; Ramon, C.; Bravo, F.D.; Franc, Y.; Fournier, N.; Michetti, P.; Mueller, C.; Geuking, M.; et al. Microbial network disturbances in relapsing refractory Crohn’s disease. Nat. Med. 2019, 25, 323–336. [Google Scholar] [CrossRef]
  100. Dodd, D.; Spitzer, M.H.; van Treuren, W.; Merrill, B.D.; Hryckowian, A.J.; Higginbottom, S.K.; Le, A.; Cowan, T.M.; Nolan, G.P.; Fischbach, M.A.; et al. A gut bacterial pathway metabolizes aromatic amino acids into nine circulating metabolites. Nature 2017, 551, 648–652. [Google Scholar] [CrossRef] [PubMed]
  101. Cohen, L.J.; Cho, J.H.; Gevers, D.; Chu, H. Genetic Factors and the Intestinal Microbiome Guide Development of Microbe-Based Therapies for Inflammatory Bowel Diseases. Gastroenterology 2019, 156, 2174–2189. [Google Scholar] [CrossRef] [PubMed]
  102. Caparrós, E.; Wiest, R.; Scharl, M.; Rogler, G.; Gutiérrez Casbas, A.; Yilmaz, B.; Wawrzyniak, M.; Francés, R. Dysbi-otic microbiota interactions in Crohn’s disease. Gut Microbes 2021, 13, 1949096. [Google Scholar] [CrossRef]
  103. Morais, L.H.; Schreiber, H.L.; Mazmanian, S.K. The gut microbiota-brain axis in behaviour and brain disorders. Nature reviews. Microbiology 2021, 19, 241–255. [Google Scholar] [CrossRef]
  104. Wang, Y.; Zhang, Z.; Li, B.; He, B.; Li, L.; Nice, E.C.; Zhang, W.; Xu, J. New Insights into the Gut Microbiota in Neurodegenerative Diseases from the Perspective of Redox Homeostasis. Antioxidants 2022, 11, 2287. [Google Scholar] [CrossRef]
  105. Shandilya, S.; Kumar, S.; Kumar Jha, N.; Kumar Kesari, K.; Ruokolainen, J. Interplay of gut microbiota and oxidative stress: Perspective on neurodegeneration and neuroprotection. J. Adv. Res. 2022, 38, 223–244. [Google Scholar] [CrossRef]
  106. Tranah, T.H.; Edwards, L.A.; Schnabl, B.; Shawcross, D.L. Targeting the gut-liver-immune axis to treat cirrhosis. Gut 2021, 70, 982–994. [Google Scholar] [CrossRef]
  107. Ni, Q.; Ye, Z.; Wang, Y.; Chen, J.; Zhang, W.; Ma, C.; Li, K.; Liu, Y.; Liu, L.; Han, Z.; et al. Gut Microbial Dysbiosis and Plasma Metabolic Profile in Individuals with Vitiligo. Front. Microbiol. 2020, 11, 592248. [Google Scholar] [CrossRef]
  108. Ni, Q.; Zhang, P.; Li, Q.; Han, Z. Oxidative Stress and Gut Microbiome in Inflammatory Skin Diseases. Front. Cell Dev. Biol. 2022, 10, 849985. [Google Scholar] [CrossRef]
  109. Haderer, M.; Neubert, P.; Rinner, E.; Scholtis, A.; Broncy, L.; Gschwendtner, H.; Kandulski, A.; Pavel, V.; Mehrl, A.; Brochhausen, C.; et al. Novel pathomechanism for spontaneous bacterial peritonitis: Disruption of cell junctions by cellular and bacterial proteases. Gut 2022, 71, 580–592. [Google Scholar] [CrossRef] [PubMed]
  110. Albillos, A.; de Gottardi, A.; Rescigno, M. The gut-liver axis in liver disease: Pathophysiological basis for therapy. J. Hepatol. 2020, 72, 558–577. [Google Scholar] [CrossRef]
  111. Bao, M.; Wang, K.; Li, J.; Li, Y.; Zhu, H.; Lu, M.; Zhang, Y.; Fan, Q.; Han, L.; Wang, K.; et al. ROS Scavenging and inflammation-directed polydopamine nanoparticles regulate gut immunity and flora therapy in inflammatory bowel disease. Acta Biomater. 2023, 161, 250–264. [Google Scholar] [CrossRef]
  112. O’Neill, S.; Brault, J.; Stasia, M.-J.; Knaus, U.G. Genetic disorders coupled to ROS deficiency. Redox Biol. 2015, 6, 135–156. [Google Scholar] [CrossRef]
  113. Thiagarajah, J.R.; Chang, J.; Goettel, J.A.; Verkman, A.S.; Lencer, W.I. Aquaporin-3 mediates hydrogen peroxide-dependent responses to environmental stress in colonic epithelia. Proc. Natl. Acad. Sci. USA 2017, 114, 568–573. [Google Scholar] [CrossRef]
  114. Rogler, G. Resolution of inflammation in inflammatory bowel disease. Lancet Gastroenterol. Hepatol. 2017, 2, 521–530. [Google Scholar] [CrossRef]
  115. Banan, A.; Choudhary, S.; Zhang, Y.; Fields, J.Z.; Keshavarzian, A. Ethanol-induced barrier dysfunction and its pre-vention by growth factors in human intestinal monolayers: Evidence for oxidative and cytoskeletal mechanisms. J. Pharmacol. Exp. Ther. 1999, 291, 1075–1085. [Google Scholar]
  116. Rao, R.; Baker, R.D.; Baker, S.S. Inhibition of oxidant-induced barrier disruption and protein tyrosine phosphoryla-tion in Caco-2 cell monolayers by epidermal growth factor. Biochem. Pharmacol. 1999, 57, 685–695. [Google Scholar] [CrossRef]
  117. Pacher, P.; Beckman, J.S.; Liaudet, L. Nitric oxide and peroxynitrite in health and disease. Physiol. Rev. 2007, 87, 315–424. [Google Scholar] [CrossRef]
  118. Babbs, C.F. Oxygen radicals in ulcerative colitis. Free. Radic. Biol. Med. 1992, 13, 169–181. [Google Scholar] [CrossRef]
  119. McBee, M.E.; Chionh, Y.H.; Sharaf, M.L.; Ho, P.; Cai, M.W.L.; Dedon, P.C. Production of Superoxide in Bacteria Is Stress- and Cell State-Dependent: A Gating-Optimized Flow Cytometry Method that Minimizes ROS Measurement Artifacts with Fluorescent Dyes. Front. Microbiol. 2017, 8, 459. [Google Scholar] [CrossRef]
  120. Imlay, J.A. Cellular defenses against superoxide and hydrogen peroxide. Annu. Rev. Biochem. 2008, 77, 755–776. [Google Scholar] [CrossRef]
  121. Voltan, S.; Martines, D.; Elli, M.; Brun, P.; Longo, S.; Porzionato, A.; Macchi, V.; D’Incà, R.; Scarpa, M.; Palù, G.; et al. Lactobacillus crispatus M247-derived H2O2 acts as a signal transducing molecule activating peroxisome prolif-erator activated receptor-gamma in the intestinal mucosa. Gastroenterology 2008, 135, 1216–1227. [Google Scholar] [CrossRef]
  122. Ito, A.; Sato, Y.; Kudo, S.; Sato, S.; Nakajima, H.; Toba, T. The screening of hydrogen peroxide-producing lactic acid bacteria and their application to inactivating psychrotrophic food-borne pathogens. Curr. Microbiol. 2003, 47, 231–236. [Google Scholar] [CrossRef]
  123. Vetrano, S.; Rescigno, M.; Cera, M.R.; Correale, C.; Rumio, C.; Doni, A.; Fantini, M.; Sturm, A.; Borroni, E.; Repici, A.; et al. Unique role of junctional adhesion molecule-a in maintaining mucosal homeostasis in inflammatory bowel disease. Gastroenterology 2008, 135, 173–184. [Google Scholar] [CrossRef] [PubMed]
  124. Zeissig, S.; Bürgel, N.; Günzel, D.; Richter, J.; Mankertz, J.; Wahnschaffe, U.; Kroesen, A.J.; Zeitz, M.; Fromm, M.; Schulzke, J.-D. Changes in expression and distribution of claudin 2, 5 and 8 lead to discontinuous tight junctions and barrier dysfunction in active Crohn’s disease. Gut 2007, 56, 61–72. [Google Scholar] [CrossRef] [PubMed]
  125. Fantini, M.C.; Guadagni, I. From inflammation to colitis-associated colorectal cancer in inflammatory bowel disease: Pathogenesis and impact of current therapies. Dig. Liver Dis. 2021, 53, 558–565. [Google Scholar] [CrossRef]
  126. Rogler, G. Chronic ulcerative colitis and colorectal cancer. Cancer Lett. 2014, 345, 235–241. [Google Scholar] [CrossRef] [PubMed]
  127. Bardelčíková, A.; Šoltys, J.; Mojžiš, J. Oxidative Stress, Inflammation and Colorectal Cancer: An Overview. Antioxidants 2023, 12, 901. [Google Scholar] [CrossRef] [PubMed]
  128. Kelder, T.; Stroeve, J.H.M.; Bijlsma, S.; Radonjic, M.; Roeselers, G. Correlation network analysis reveals relationships between diet-induced changes in human gut microbiota and metabolic health. Nutr. Diabetes 2014, 4, e122. [Google Scholar] [CrossRef]
  129. Rajilić-Stojanović, M.; Jonkers, D.M.; Salonen, A.; Hanevik, K.; Raes, J.; Jalanka, J.; de Vos, W.M.; Manichanh, C.; Golic, N.; Enck, P.; et al. Intestinal microbiota and diet in IBS: Causes, consequences, or epiphenomena? Am. J. Gastroenterol. 2015, 110, 278–287. [Google Scholar] [CrossRef]
  130. Chooruk, A.; Piwat, S.; Teanpaisan, R. Antioxidant activity of various oral Lactobacillus strains. J. Appl. Microbiol. 2017, 123, 271–279. [Google Scholar] [CrossRef]
  131. Zorov, D.B.; Plotnikov, E.Y.; Silachev, D.N.; Zorova, L.D.; Pevzner, I.B.; Zorov, S.D.; Babenko, V.A.; Jankauskas, S.S.; Popkov, V.A.; Savina, P.S. Microbiota and mitobiota. Putting an equal sign between mitochondria and bacteria. Biochemistry 2014, 79, 1017–1031. [Google Scholar] [CrossRef]
  132. Wang, Y.; Wu, Y.; Wang, Y.; Xu, H.; Mei, X.; Yu, D.; Wang, Y.; Li, W. Antioxidant Properties of Probiotic Bacteria. Nutrients 2017, 9, 521. [Google Scholar] [CrossRef] [PubMed]
  133. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.; Salminen, S.; et al. Expert consensus document. The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 506–514. [Google Scholar] [CrossRef] [PubMed]
  134. Guarner, F.; Khan, A.G.; Garisch, J.; Eliakim, R.; Gangl, A.; Thomson, A.; Krabshuis, J.; Lemair, T.; Kaufmann, P.; de Paula, J.A.; et al. World Gastroenterology Organisation Global Guidelines: Probiotics and prebiotics October 2011. J. Clin. Gastroenterol. 2012, 46, 468–481. [Google Scholar] [CrossRef] [PubMed]
  135. Wieërs, G.; Belkhir, L.; Enaud, R.; Leclercq, S.; Philippart de Foy, J.-M.; Dequenne, I.; de Timary, P.; Cani, P.D. How Probiotics Affect the Microbiota. Front. Cell. Infect. Microbiol. 2019, 9, 454. [Google Scholar] [CrossRef]
  136. Kucharzik, T.; Dignass, A.; Areya, R.; Bokemeyer, B.; Esters, P.; Herrlinger, K.; Kannengiesser, K.; Kienle, P.; Langhorst, J.; Lügering, A.; et al. Aktualisierte S3-Leitlinie Colitis ulcerosa (Version 6.1). Available online: https://www.dgvs.de/wp-content/uploads/2023/03/LL-CU_Leitlinie_final_20.03.223.pdf (accessed on 20 March 2023).
  137. Su, G.L.; Ko, C.W.; Bercik, P.; Falck-Ytter, Y.; Sultan, S.; Weizman, A.V.; Morgan, R.L. AGA Clinical Practice Guidelines on the Role of Probiotics in the Management of Gastrointestinal Disorders. Gastroenterology 2020, 159, 697–705. [Google Scholar] [CrossRef]
  138. Ragan, M.V.; Wala, S.J.; Goodman, S.D.; Bailey, M.T.; Besner, G.E. Next-Generation Probiotic Therapy to Protect the Intestines from Injury. Front. Cell. Infect. Microbiol. 2022, 12, 863949. [Google Scholar] [CrossRef]
  139. Clostridioides Difficile. Available online: https://www.rki.de/DE/Content/Infekt/EpidBull/Merkblaetter/Ratgeber_Clostridium.html (accessed on 1 May 2023).
  140. Layer, P.; Andresen, V.; Allescher, H.; Bischoff, S.C.; Claßen, M.; Elsenbruch, S.; Freitag, M.; Frieling, T.; Gebhard, M.; Goebel-Stengel, M.; et al. Update S3-Leitlinie Reizdarmsyndrom: Definition, Pathophysiologie, Diagnostik und Therapie. Gemeinsame Leitlinie der Deutschen Gesellschaft für Gastroenterologie, Verdauungs- und Stoffwechselkrankheiten (DGVS) und der Deutschen Gesellschaft für Neurogastroenterologie und Motilität (DGNM)—Juni 2021—AWMF-Registriernummer: 021/016. Z. Gastroenterol. 2021, 59, 1323–1415. [Google Scholar] [CrossRef]
  141. Feng, T.; Wang, J. Oxidative stress tolerance and antioxidant capacity of lactic acid bacteria as probiotic: A systematic review. Gut Microbes 2020, 12, 1801944. [Google Scholar] [CrossRef]
  142. Chiarugi, P.; Pani, G.; Giannoni, E.; Taddei, L.; Colavitti, R.; Raugei, G.; Symons, M.; Borrello, S.; Galeotti, T.; Ramponi, G. Reactive oxygen species as essential mediators of cell adhesion: The oxidative inhibition of a FAK tyrosine phosphatase is required for cell adhesion. J. Cell Biol. 2003, 161, 933–944. [Google Scholar] [CrossRef]
  143. Diaz, B.; Shani, G.; Pass, I.; Anderson, D.; Quintavalle, M.; Courtneidge, S.A. Tks5-dependent, nox-mediated generation of reactive oxygen species is necessary for invadopodia formation. Sci. Signal. 2009, 2, ra53. [Google Scholar] [CrossRef]
  144. Wentworth, C.C.; Jones, R.M.; Kwon, Y.M.; Nusrat, A.; Neish, A.S. Commensal-epithelial signaling mediated via formyl peptide receptors. Am. J. Pathol. 2010, 177, 2782–2790. [Google Scholar] [CrossRef] [PubMed]
  145. Swanson, P.A.; Kumar, A.; Samarin, S.; Vijay-Kumar, M.; Kundu, K.; Murthy, N.; Hansen, J.; Nusrat, A.; Neish, A.S. Enteric commensal bacteria potentiate epithelial restitution via reactive oxygen species-mediated inactivation of focal adhesion kinase phosphatases. Proc. Natl. Acad. Sci. USA 2011, 108, 8803–8808. [Google Scholar] [CrossRef] [PubMed]
  146. Migeotte, I.; Communi, D.; Parmentier, M. Formyl peptide receptors: A promiscuous subfamily of G pro-tein-coupled receptors controlling immune responses. Cytokine Growth Factor Rev. 2006, 17, 501–519. [Google Scholar] [CrossRef]
  147. Tiso, M.; Schechter, A.N. Nitrate reduction to nitrite, nitric oxide and ammonia by gut bacteria under physiological conditions. PLoS ONE 2015, 10, e0119712. [Google Scholar] [CrossRef]
  148. Silva, Y.P.; Bernardi, A.; Frozza, R.L. The Role of Short-Chain Fatty Acids from Gut Microbiota in Gut-Brain Communication. Front. Endocrinol. 2020, 11, 25. [Google Scholar] [CrossRef]
  149. Hoyles, L.; Snelling, T.; Umlai, U.-K.; Nicholson, J.K.; Carding, S.R.; Glen, R.C.; McArthur, S. Microbiome-host sys-tems interactions: Protective effects of propionate upon the blood-brain barrier. Microbiome 2018, 6, 55. [Google Scholar] [CrossRef]
  150. Peng, L.; Li, Z.-R.; Green, R.S.; Holzman, I.R.; Lin, J. Butyrate enhances the intestinal barrier by facilitating tight junction assembly via activation of AMP-activated protein kinase in Caco-2 cell monolayers. J. Nutr. 2009, 139, 1619–1625. [Google Scholar] [CrossRef]
  151. Saleri, R.; Borghetti, P.; Ravanetti, F.; Cavalli, V.; Ferrari, L.; de Angelis, E.; Andrani, M.; Martelli, P. Effects of dif-ferent short-chain fatty acids (SCFA) on gene expression of proteins involved in barrier function in IPEC-J2. Porc. Health Manag. 2022, 8, 21. [Google Scholar] [CrossRef]
  152. Pérez-Reytor, D.; Puebla, C.; Karahanian, E.; García, K. Use of Short-Chain Fatty Acids for the Recovery of the In-testinal Epithelial Barrier Affected by Bacterial Toxins. Front. Physiol. 2021, 12, 650313. [Google Scholar] [CrossRef]
  153. González-Bosch, C.; Boorman, E.; Zunszain, P.A.; Mann, G.E. Short-chain fatty acids as modulators of redox signaling in health and disease. Redox Biol. 2021, 47, 102165. [Google Scholar] [CrossRef]
  154. Guo, W.; Liu, J.; Sun, J.; Gong, Q.; Ma, H.; Kan, X.; Cao, Y.; Wang, J.; Fu, S. Butyrate alleviates oxidative stress by regulating NRF2 nuclear accumulation and H3K9/14 acetylation via GPR109A in bovine mammary epithelial cells and mammary glands. Free Radic. Biol. Med. 2020, 152, 728–742. [Google Scholar] [CrossRef] [PubMed]
  155. Yamamoto, M.; Kensler, T.W.; Motohashi, H. The KEAP1-NRF2 System: A Thiol-Based Sensor-Effector Apparatus for Maintaining Redox Homeostasis. Physiol. Rev. 2018, 98, 1169–1203. [Google Scholar] [CrossRef]
  156. Azad, M.A.K.; Sarker, M.; Li, T.; Yin, J. Probiotic Species in the Modulation of Gut Microbiota: An Overview. Biomed Res. Int. 2018, 2018, 9478630. [Google Scholar] [CrossRef] [PubMed]
  157. Park, J.-S.; Choi, J.W.; Jhun, J.; Kwon, J.Y.; Lee, B.-I.; Yang, C.W.; Park, S.-H.; Cho, M.-L. Lactobacillus acidophilus Improves Intestinal Inflammation in an Acute Colitis Mouse Model by Regulation of Th17 and Treg Cell Balance and Fibrosis Development. J. Med. Food 2018, 21, 215–224. [Google Scholar] [CrossRef]
  158. Geuking, M.B.; Cahenzli, J.; Lawson, M.A.E.; Ng, D.C.K.; Slack, E.; Hapfelmeier, S.; McCoy, K.D.; Macpherson, A.J. Intestinal bacterial colonization induces mutualistic regulatory T cell responses. Immunity 2011, 34, 794–806. [Google Scholar] [CrossRef]
  159. Mazmanian, S.K.; Liu, C.H.; Tzianabos, A.O.; Kasper, D.L. An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell 2005, 122, 107–118. [Google Scholar] [CrossRef] [PubMed]
  160. Atarashi, K.; Tanoue, T.; Oshima, K.; Suda, W.; Nagano, Y.; Nishikawa, H.; Fukuda, S.; Saito, T.; Narushima, S.; Hase, K.; et al. Treg induction by a rationally selected mixture of Clostridia strains from the human microbiota. Nature 2013, 500, 232–236. [Google Scholar] [CrossRef]
  161. Lin, S.; Li, Y.; Zamyatnin, A.A.; Werner, J.; Bazhin, A.V. Reactive oxygen species and colorectal cancer. J. Cell. Physiol. 2018, 233, 5119–5132. [Google Scholar] [CrossRef]
  162. Pircalabioru, G.; Aviello, G.; Kubica, M.; Zhdanov, A.; Paclet, M.-H.; Brennan, L.; Hertzberger, R.; Papkovsky, D.; Bourke, B.; Knaus, U.G. Defensive Mutualism Rescues NADPH Oxidase Inactivation in Gut Infection. Cell Host Microbe 2016, 19, 651–663. [Google Scholar] [CrossRef]
  163. Eeckhaut, V.; Machiels, K.; Perrier, C.; Romero, C.; Maes, S.; Flahou, B.; Steppe, M.; Haesebrouck, F.; Sas, B.; Ducatelle, R.; et al. Butyricicoccus pullicaecorum in inflammatory bowel disease. Gut 2013, 62, 1745–1752. [Google Scholar] [CrossRef]
  164. Martín, R.; Miquel, S.; Benevides, L.; Bridonneau, C.; Robert, V.; Hudault, S.; Chain, F.; Berteau, O.; Azevedo, V.; Chatel, J.M.; et al. Functional Characterization of Novel Faecalibacterium prausnitzii Strains Isolated from Healthy Volunteers: A Step Forward in the Use of F. prausnitzii as a Next-Generation Probiotic. Front. Microbiol. 2017, 8, 1226. [Google Scholar] [CrossRef] [PubMed]
  165. Hu, W.; Gao, W.; Liu, Z.; Fang, Z.; Wang, H.; Zhao, J.; Zhang, H.; Lu, W.; Chen, W. Specific Strains of Faecalibacterium prausnitzii Ameliorate Nonalcoholic Fatty Liver Disease in Mice in Association with Gut Microbiota Regulation. Nutrients 2022, 14, 2945. [Google Scholar] [CrossRef] [PubMed]
  166. Mah, C.; Jayawardana, T.; Leong, G.; Koentgen, S.; Lemberg, D.; Connor, S.J.; Rokkas, T.; Grimm, M.C.; Leach, S.T.; Hold, G.L. Assessing the Relationship between the Gut Microbiota and Inflammatory Bowel Disease Therapeutics: A Systematic Review. Pathogens 2023, 12, 262. [Google Scholar] [CrossRef] [PubMed]
  167. Glassner, K.L.; Abraham, B.P.; Quigley, E.M.M. The microbiome and inflammatory bowel disease. J. Allergy Clin. Immunol. 2020, 145, 16–27. [Google Scholar] [CrossRef]
  168. Round, J.L.; Lee, S.M.; Li, J.; Tran, G.; Jabri, B.; Chatila, T.A.; Mazmanian, S.K. The Toll-like receptor 2 pathway establishes colonization by a commensal of the human microbiota. Science 2011, 332, 974–977. [Google Scholar] [CrossRef]
  169. Hrncir, T.; Hrncirova, L.; Kverka, M.; Hromadka, R.; Machova, V.; Trckova, E.; Kostovcikova, K.; Kralickova, P.; Krejsek, J.; Tlaskalova-Hogenova, H. Gut Microbiota and NAFLD: Pathogenetic Mechanisms, Microbiota Signa-tures, and Therapeutic Interventions. Microorganisms 2021, 9, 957. [Google Scholar] [CrossRef]
  170. Widhiati, S.; Purnomosari, D.; Wibawa, T.; Soebono, H. The role of gut microbiome in inflammatory skin disorders: A systematic review. Dermatol. Rep. 2022, 14, 9188. [Google Scholar] [CrossRef]
  171. Kim, J.; Lee, H.K. Potential Role of the Gut Microbiome in Colorectal Cancer Progression. Front. Immunol. 2021, 12, 807648. [Google Scholar] [CrossRef]
  172. Obermeier, F.; Hofmann, C.; Falk, W. Inflammatory bowel diseases: When natural friends turn into enemies-the importance of CpG motifs of bacterial DNA in intestinal homeostasis and chronic intestinal inflammation. Int. J. Inflam. 2010, 2010, 641910. [Google Scholar] [CrossRef]
  173. Hofmann, C.; Dunger, N.; Doser, K.; Lippert, E.; Siller, S.; Edinger, M.; Falk, W.; Obermeier, F. Physiologic TLR9-CpG-DNA interaction is essential for the homeostasis of the intestinal immune system. Inflamm. Bowel Dis. 2014, 20, 136–143. [Google Scholar] [CrossRef]
  174. Cheung, E.C.; Vousden, K.H. The role of ROS in tumour development and progression. Nat. Rev. Cancer 2022, 22, 280–297. [Google Scholar] [CrossRef] [PubMed]
  175. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef] [PubMed]
  176. Bell, H.N.; Rebernick, R.J.; Goyert, J.; Singhal, R.; Kuljanin, M.; Kerk, S.A.; Huang, W.; Das, N.K.; Andren, A.; Solanki, S.; et al. Reuterin in the healthy gut microbiome suppresses colorectal cancer growth through altering redox balance. Cancer Cell 2022, 40, 185–200.e6. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic illustration of the mitochondrial respiratory chain. Electrons are fed into the respiratory chain via NADH in complex I or via FADH2 in complex II. The electrons are passed through the respiratory chain and transferred to O2, resulting in the formation of water. However, some of the electrons can leave the respiratory chain, leading to a univalent reduction of oxygen and O2· formation. O2· is converted into H2O2 by SODs. (b) Scheme of phagocytic NADPHoxidase. The phagocytic NADPHoxidases are multicomponent complexes that catalyze a one-electron reduction of oxygen by NADPH. The resulting O2· is generated to the outside of the cell or to the inside of phagosomes. Q: ubiquinol; C: cytochrome C; gp: glycoprotein; RAC: subfamily of the Rho family of GTPases. The figure was created with BioRender.com (accessed on 1 May 2023).
Figure 1. (a) Schematic illustration of the mitochondrial respiratory chain. Electrons are fed into the respiratory chain via NADH in complex I or via FADH2 in complex II. The electrons are passed through the respiratory chain and transferred to O2, resulting in the formation of water. However, some of the electrons can leave the respiratory chain, leading to a univalent reduction of oxygen and O2· formation. O2· is converted into H2O2 by SODs. (b) Scheme of phagocytic NADPHoxidase. The phagocytic NADPHoxidases are multicomponent complexes that catalyze a one-electron reduction of oxygen by NADPH. The resulting O2· is generated to the outside of the cell or to the inside of phagosomes. Q: ubiquinol; C: cytochrome C; gp: glycoprotein; RAC: subfamily of the Rho family of GTPases. The figure was created with BioRender.com (accessed on 1 May 2023).
Biomedicines 11 01388 g001
Figure 2. Formation of the NALP3 inflammasome. Mitochondria-derived ROS induce the release of TXNIP (VDUP-1) from the Thioredoxin/TXNIP complex. TXNIP (VDUP-1) binds to NALP3 and leads to the formation of the inflammasome. The NALP3 inflammasome induces activation of caspase 1 and caspase 12. These proinflammatory caspases in turn cleave and activate proinflammatory cytokines IL1β and IL18. NALP: NLR family pyrin domain containing; TRX: thioredoxin; TXNIP: thioredoxin interacting protein. The figure was created with BioRender.com (accessed on 1 May 2023).
Figure 2. Formation of the NALP3 inflammasome. Mitochondria-derived ROS induce the release of TXNIP (VDUP-1) from the Thioredoxin/TXNIP complex. TXNIP (VDUP-1) binds to NALP3 and leads to the formation of the inflammasome. The NALP3 inflammasome induces activation of caspase 1 and caspase 12. These proinflammatory caspases in turn cleave and activate proinflammatory cytokines IL1β and IL18. NALP: NLR family pyrin domain containing; TRX: thioredoxin; TXNIP: thioredoxin interacting protein. The figure was created with BioRender.com (accessed on 1 May 2023).
Biomedicines 11 01388 g002
Figure 3. Direct and indirect influence of the gut microbiome on the intestinal REDOX status under physiological conditions. Gut bacteria can produce oxidative compounds but also have developed specific mechanisms to resist an oxidative environment. Moreover, gut bacteria can induce ROS production by intestinal epithelial cells. Beyond that, bacteria, especially probiotics, induce antioxidative activities in epithelial cells. This can occur directly or indirectly via SCFA production and Nrf2 activity. Furthermore, probiotics have beneficial effects on rebalancing immune responses mediated by Th17 and Treg cells. The figure was created with BioRender.com (accessed on 1 May 2023).
Figure 3. Direct and indirect influence of the gut microbiome on the intestinal REDOX status under physiological conditions. Gut bacteria can produce oxidative compounds but also have developed specific mechanisms to resist an oxidative environment. Moreover, gut bacteria can induce ROS production by intestinal epithelial cells. Beyond that, bacteria, especially probiotics, induce antioxidative activities in epithelial cells. This can occur directly or indirectly via SCFA production and Nrf2 activity. Furthermore, probiotics have beneficial effects on rebalancing immune responses mediated by Th17 and Treg cells. The figure was created with BioRender.com (accessed on 1 May 2023).
Biomedicines 11 01388 g003
Figure 4. Overview of bacteria that favor a pro-oxidative state and bacteria that favor an anti-oxidative environment. The figure was created with BioRender.com (accessed on 1 May 2023).
Figure 4. Overview of bacteria that favor a pro-oxidative state and bacteria that favor an anti-oxidative environment. The figure was created with BioRender.com (accessed on 1 May 2023).
Biomedicines 11 01388 g004
Figure 5. SCFAs induce proliferation and strengthen oxidative defense. SFCAs bind to FFARs and activate the RAS/RAF/MEK/ERK signaling pathway, inducing proliferation. Furthermore, stimulation of the FFAR receptor by SCFA butyrate shifts the AMP/ATP ratio and induces AMPK. Activated AMPK induces translocation of Nrf2 into the nucleus. In addition, inhibition of HDACs also increases Nrf2 synthesis. Thus, Nrf2-mediated activation of oxidative defense mechanisms occurs. SCFA: short- chain fatty acid; FFAR: free fatty acid receptor; RAS: rat sarcoma virus protein; RAF: rapidly accelerated fibrosarcoma protein; MEK: mitogen activated kinase; ERK: extracellular signal-regulated kinase; AMPK: AMP-activated protein kinase; HDAC: histone deacetylase; Nrf2: Nuclear Factor Erythroid 2-related Factor 2. The figure was created with BioRender.com (accessed on 1 May 2023).
Figure 5. SCFAs induce proliferation and strengthen oxidative defense. SFCAs bind to FFARs and activate the RAS/RAF/MEK/ERK signaling pathway, inducing proliferation. Furthermore, stimulation of the FFAR receptor by SCFA butyrate shifts the AMP/ATP ratio and induces AMPK. Activated AMPK induces translocation of Nrf2 into the nucleus. In addition, inhibition of HDACs also increases Nrf2 synthesis. Thus, Nrf2-mediated activation of oxidative defense mechanisms occurs. SCFA: short- chain fatty acid; FFAR: free fatty acid receptor; RAS: rat sarcoma virus protein; RAF: rapidly accelerated fibrosarcoma protein; MEK: mitogen activated kinase; ERK: extracellular signal-regulated kinase; AMPK: AMP-activated protein kinase; HDAC: histone deacetylase; Nrf2: Nuclear Factor Erythroid 2-related Factor 2. The figure was created with BioRender.com (accessed on 1 May 2023).
Biomedicines 11 01388 g005
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

Kunst, C.; Schmid, S.; Michalski, M.; Tümen, D.; Buttenschön, J.; Müller, M.; Gülow, K. The Influence of Gut Microbiota on Oxidative Stress and the Immune System. Biomedicines 2023, 11, 1388. https://doi.org/10.3390/biomedicines11051388

AMA Style

Kunst C, Schmid S, Michalski M, Tümen D, Buttenschön J, Müller M, Gülow K. The Influence of Gut Microbiota on Oxidative Stress and the Immune System. Biomedicines. 2023; 11(5):1388. https://doi.org/10.3390/biomedicines11051388

Chicago/Turabian Style

Kunst, Claudia, Stephan Schmid, Marlen Michalski, Deniz Tümen, Jonas Buttenschön, Martina Müller, and Karsten Gülow. 2023. "The Influence of Gut Microbiota on Oxidative Stress and the Immune System" Biomedicines 11, no. 5: 1388. https://doi.org/10.3390/biomedicines11051388

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