Next Article in Journal
MMP3 as a Molecular Link: Unraveling the Connection Between Ankylosing Spondylitis and Acute Coronary Syndrome
Previous Article in Journal
Primary Cilia, Hypoxia, and Liver Dysfunction: A New Perspective on Biliary Atresia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phenotyping the Chemical Communications of the Intestinal Microbiota and the Host: Secondary Bile Acids as Postbiotics

1
Dipartimento di Medicina e Chirurgia, Università degli Studi di Perugia, 06123 Perugia, Italy
2
SC di Gastroenterologia ed Epatologia, Azienda Ospedaliera di Perugia, 06123 Perugia, Italy
*
Author to whom correspondence should be addressed.
Cells 2025, 14(8), 595; https://doi.org/10.3390/cells14080595
Submission received: 4 March 2025 / Revised: 10 April 2025 / Accepted: 12 April 2025 / Published: 15 April 2025

Abstract

:
The current definition of a postbiotic is a “preparation of inanimate microorganisms and/or their components that confers a health benefit on the host”. Postbiotics can be mainly classified as metabolites, derived from intestinal bacterial fermentation, or structural components, as intrinsic constituents of the microbial cell. Secondary bile acids deoxycholic acid (DCA) and lithocholic acid (LCA) are bacterial metabolites generated by the enzymatic modifications of primary bile acids by microbial enzymes. Secondary bile acids function as receptor ligands modulating the activity of a family of bile-acid-regulated receptors (BARRs), including GPBAR1, Vitamin D (VDR) receptor and RORγT expressed by various cell types within the entire human body. Secondary bile acids integrate the definition of postbiotics, exerting potential beneficial effects on human health given their ability to regulate multiple biological processes such as glucose metabolism, energy expenditure and inflammation/immunity. Although there is evidence that bile acids might be harmful to the intestine, most of this evidence does not account for intestinal dysbiosis. This review examines this novel conceptual framework of secondary bile acids as postbiotics and how these mediators participate in maintaining host health.

1. Definition of Postbiotics: Differences Between Prebiotics and Probiotics

According to the International Scientific Association for Probiotics and Prebiotics (ISAPP), a postbiotic is a “preparation of inanimate microorganisms and/or their components that confers a health benefit on the host” [1]. Despite the fact that multiple definitions of “postbiotic” have been previously proposed (Table 1), taking or not in consideration the inclusion of microbial cells in the preparation, none of them resulted in perfectly fitting with the intrinsic meaning of this concept.
Within the definition of a postbiotic, intact microorganisms are not required for health effects: what is needed, as a part of the manufacturing process of a postbiotic, is a deliberate chemical or physical process of viability termination (via heat, radiation, high pressure or lysis) that may maintain (or not) cell integrity [8]. To qualify as a postbiotic, the exact microbial composition must be characterized before inactivation processes: indeed, preparations obtained from undefined microorganisms do not fit with the postbiotic definition. On the contrary, metabolites generated during intestinal bacteria fermentation, i.e., secreted proteins and short-chain fatty acids (SCFAs), or molecules representing structural fragments of these bacteria, such as exopolysaccharides (EPS) and lipoteichoic acid (LTA), could be considered as postbiotics, as long as they are in the presence of the inactivated microbial cell and/or its cell components, too [9,10].
Thus, while probiotics are living nonpathogenic microorganisms (such as Saccharomyces boulardii yeast or Lactobacillus and Bifidobacterium species) able to provide health benefits if assumed in adequate amounts [11] and prebiotics are non-digestible substances (i.e., dietary fibers) that promote probiotics’ growth, contribute to gut health and exert immunomodulatory effects [12], postbiotics refer to bioactive compounds produced by microorganisms during their growth, including the inactivated form of the microbial cells—that does not necessarily mean that they need to be qualified as a probiotic (while living) to be accepted as a postbiotic. Currently, inanimate strains of genera from the Lactobacillaceae family and strains of the genus Bifidobacterium represent the bulk of known postbiotics [13].

2. Classification of Postbiotics

As mentioned above, it is possible to identify two main categories of postbiotics: (i) metabolites, derived from intestinal bacterial fermentation, and (ii) structural components, as intrinsic constituents of the microbial cell [14]. Despite SCFAs, vitamins and peptides represent the best characterized postbiotics, and a large variety of bioactive molecules fitting the definition of postbiotics have been identified (Table 2).

2.1. Short-Chain Fatty Acids (SCFAs)

Short-chain fatty acids (SCFAs) are a subset of fatty acids derived from the fermentation of partially and non-digestible polysaccharides (dietary fibers and resistant starches) carried by the enzymatic activity of specific taxa belonging to Bacteroidetes and Firmicutes species [18]. The carbon chain of SCFAs, for definition, is composed of less than six carbons, acetate (C2), propionate (C3) and butyrate (C4) being the most represented ones [19]. Indeed, these three SCFAs contribute to 80% of the total SCFA pool in the human body (~60% acetate, ~20% propionate and butyrate) [20].
The composition of SCFAs’ pool fluctuates throughout life, depending both on diet variety and gut microbiota composition [21]: high fiber–low fat diets expand the SCFAs’ pool and increase the fecal and blood SCFA content (fSCFAs, bSCFAs); a diet with low fiber content does the opposite [14,22]. However, obesity is related to increased levels of total fSCFAs, too [23,24], which are reduced following an anti-obesity treatment [25]. Also, changes in the intestinal microbiota, as mentioned above, have a direct impact of SCFAs’ pool heterogeneity: early stages of life (0–3 years) are characterized by high levels of acetate, mainly produced by Bifidobacteria (in particular Bifidobacterium breve and Bifidobacterium bifidum), due to human milk oligosaccharide (HMO) consumption in breastfeeding [26,27]; in adulthood, increased levels of propionate and butyrate positively correlate with an increase in Firmicutes (e.g., Lactobacillaceae, Ruminococcaceae, Lachnospiraceae) [28,29,30] and Bacteroidetes (e.g Bacteroides fragilis, Bacteroides thetaiotaomicron, Bacteroides caccae, Bacteroides ovatus) [31,32,33], which are reduced at the elderly stage due to a decrease in microbial diversity of commensal taxa accompanied by an increased abundance of pathobionts such as Enterobacteriaceae and Streptococcus spp. [34,35,36].
Thanks to their ability to bind several cell-surface G protein-coupled receptors (GPCRs) [37], SCFAs are responsible for the development of a tolerogenic immune system by (i) promoting regulatory T (Treg) cells and quiescent dendritic cell (DC) phenotypes; (ii) enhancing epithelial barrier function; (iii) stimulating mucus secretion by intestinal goblet cells; (iv) damping inflammatory response and suppressing autoimmune reactions [38,39,40,41] (Table 3).
Moreover, SCFAs have been proved responsible for increasing insulin sensitivity by stimulating the production of insulin-sensitizing hormones like glucagon-like peptide 1 (GLP-1) and by reducing inflammation in adipose tissue [59,60]; they promote hepatic and muscular FA oxidation and adipose tissue lipolysis while inhibiting hepatic FA synthesis, contributing to lipid metabolism homeostasis [61,62]. Last, SCFAs can be converted into acetyl-CoA, thus entering the Krebs Cycle and participating in energy production [59]. An overview of the main functions of SCFAs is shown in Figure 1.

2.2. Lactic Acid

Lactic acid (LA) is a carboxylic acid generated from pyruvate, the end product of glycolysis, by the activity of the lactate dehydrogenase (LDH) enzyme [63]. In hypoxia or anaerobic conditions, when oxygen is not available as the final acceptor of electrons, pyruvate conversion to lactate allows the regeneration of nicotinamide adenine nucleotide (NAD+), essential for sustained glycolysis [64].
Lactic acid is also produced via fermentation of multiple carbohydrate sources (e.g., glucose, cellulose, xylose, maltose, lactose and others) by different Gram-positive, catalase-negative, non-spore-forming bacteria with amylase activity referred to as lactic acid bacteria (LAB) [65,66]. In addition to LA, LAB generate other growth inhibitions molecules such as bacteriocins, antifungal peptides and hydrogen peroxide (H2O2), preventing the proliferation of putative pathogen microorganisms [67,68]. Lactobacillus strains are the most well-known LA producers, even if Lactococcus, Streptococcus, Pediococcus and Enterococcus are important, too [69]. Acting like an acidifying metabolite, LA is essential for the maintenance of a healthy vaginal environment, preventing the colonization or growth of pathogens responsible for common vaginal infections, such as Gardrenella vaginitis and Candida albicans [70].

2.3. Bacteriocins

Bacteriocins are a heterogeneous group of antimicrobial peptides (AMPs) naturally secreted by both Gram-positive and Gram-negative species belonging primarily to the genera Bifidobacterium and Lactobacillus [71,72]. Bacteriocins are synthesized with the purpose of killing other bacteria, particularly pathogenic ones such as Staphylococcus aureus, Pseudomonas aeruginosa, Salmonella typhi, Listeria monocytogenes and Clostridium botulinum, thus participating in microbial competition among prokaryotes [73,74]. Bacteriocins are very diverse in terms of length, molecular weight, genetic origins, immunity mechanisms, biochemical and/or structural features and may act via multiple modes of actions, such as pore formation on target cell membrane, inhibition of cell wall synthesis as well as nucleic acid degradation through DNase and RNase activity [65,72,75]. Differently from antibiotics, bacteriocins offer more benefits as they are natural bioactive peptides with no side effects and represent a possible solution to Multiple Drug Resistance (MDR) disease-causing bacteria, more than having multiple additional positive effects on human health [76,77]. A list of the best known bacteriocins is reported in Table 4.

2.4. Secondary Bile Acids

Secondary bile acids, specifically deoxycholic acid (DCA) and lithocholic acid (LCA), are steroidal bacterial metabolites produced in the colon derived from the conversion of primary bile acids cholic acid (CA) and chenodeoxycholic acid (CDCA) carried by specific bile salt hydrolase (BSH)-expressing microorganisms [89,90] (Figure 2). Despite the most well-known purpose of bile acids being to facilitate dietary lipid emulsion and absorption, recent studies demonstrated that bile acids have several biochemical and immunological effects by acting as ligands of both membrane and nuclear receptors referred to as Bile Acid Receptors (BARs) [91,92,93]. This function will be discussed in more detail later.

2.5. Bacterial Cell Wall Components (CWCs)

Recent studies include bacterial cell wall components (CWCs) in the postbiotic class: among these, exopolysaccharides (EPSs), peptidoglycan (PG) and lipoteichoic acid (LTA) are the most relevant ones [95].
EPSs are high-molecular-weight carbohydrate biopolymers synthesized by microorganisms that can be classified into (i) capsular polysaccharides, closely associated with the cell surface, and (ii) free slime polysaccharides, loosely attached or totally secreted into the extracellular environment [96]. Species belonging to both Gram+ and Gram bacteria produce EPSs, such as Acetobacter, Gluconobacter, Pseudomonas, Enterobacter, Klebsiella, Bacillus, Streptococcus and Clostridia genera [97]. Recent studies demonstrated how EPSs from Lactobacillus brevis could lower intestinal pH, upregulate SCFA production, especially propionate and butyrate via improving intestinal microbiota [98], while EPSs from Bifidobacterium longum managed to alleviate DSS-induced intestinal inflammation in mouse model modulating macrophage polarization toward the anti-inflammatory M2-type [99].
PG is a three-dimensional polymer representing the primary component of the Gram+ and Gram bacteria cell wall, responsible for cell shape maintenance and resistance to both extracellular environmental insults and intracellular osmotic pressure (or turgor) caused by cytosolic content [100,101,102]. PG administration in vitro results in the inhibition of pro-inflammatory cytokine release such as interleukin 6 (IL-6), IL-8, IL-1β and tumor necrosis factor α (TNFα) [103] while upregulating anti-inflammatory genes, including IL-10 and transforming growth factor β (TGF-β) [104]. In addition, PG promotes collagen synthesis, fibroblast proliferation and angiogenesis, thus promoting wound healing and tissue regeneration [105,106].
LTA, instead, is a surface-associated adhesion amphiphilic molecule found exclusively in Gram+ bacteria that works to maintain ion homeostasis, resist osmotic stress and regulate autolytic activity [107,108]. Studies showed that LTA from Lactobacillus plantarum exerts anti-inflammatory activity both in vitro and in vivo, by reducing Toll-like receptor 2 (TLR2) and subsequent nuclear factor-κB (NF-κB) activation in human intestinal epithelial cells [109,110], lowering the Colitis Disease Activity Index (CDAI) as well as TNFα levels in LTA-treated mice compared to untreated ones [110].

2.6. Plasmalogens (Pls)

Plasmalogens (Pls) are a unique class of membrane glycerophospholipids characterized by the presence of a fatty alcohol and several polyunsaturated fatty acids bound to the glycerol backbone [111]. In bacteria, major evidence supports the theory that Pls play an important role in exosome fission [112]. In the human body, Pls are mainly expressed in the heart, retina and innate immune cells and represent the main component (up to 80%) of neural tissue [113,114]. Pls account for around 20% of the total human phospholipids and play important roles in cell homeostasis, cell signaling and neural transmission [115,116]. Daily oral administration of Pls as postbiotics seems to (i) regulate adipogenesis [117], (ii) have anti-inflammatory and antioxidant activities [118] as well as (iii) improve cognitive function in patients with mild Alzheimer’s disease [119].

2.7. Intestinal Bacteria-Derived Vitamins

Vitamins are organic micronutrients defined as essential constituents of the diet not endogenously synthesized by humans or not synthesized in an adequate amount to support human health [120,121]. Based on their biochemical composition, vitamins can be classified as water-soluble (B1 B2, B3, B5, B6, B7, B9, B12 and C) or fat-soluble (A, D, E and K) [122,123,124]. However, vitamins are not only introduced via exogenous sources: several gut bacteria, belonging predominantly to Bacteroides, Bifidobacterium and Enterococcus genera, contribute to vitamin synthesis with particular reference to thiamine (B1), riboflavin (B2), pantothenic acid (B5), biotin (B7), folate (B9), cobalamin (B12) and vitamin K [125,126]. Given the multiple metabolic functions of both hydro- and lipo-soluble vitamins in human health such as (i) immune-modulation, (ii) bone health maintenance, (iii) calcium balance, (iv) retinal health and sight protection, (v) antioxidant activity and (vi) blood clotting regulation, adequate amounts of vitamin uptake should be guaranteed to avoid syndromes and/or diseases derived from vitamin deficiencies [121,127,128,129].

2.8. Tryptophan Metabolites

Tryptophan is a widely investigated amino acid, essential for body health and homeostasis: it cannot be synthesized de novo by human cells but it must be supplied through the diet via bread, milk, chocolate, tuna fish and other foods rich in such amino acid [130,131]. Tryptophan is required for a normal body’s growth and development, being the in vivo precursor of several bioactive compounds such as nicotinamide (B6), serotonin, melatonin, tryptamine, kynurenine and others [132,133,134] as well as affecting metabolism of neurotransmitters and CNS compounds such as dopamine, norepinephrine and beta-endorphin [135,136,137]. Tryptophan is also converted into indoles via the action of the tryptophanase (TnaA) enzyme, expressed in various both Gram+ and Gram bacteria species including Escherichia coli, Lactobacillus spp., Clostridium spp. and Bacteroides spp. [138,139,140].
Tryptophan derivatives exert multiple beneficial effects: (i) they regulate neurotransmitters levels, thus having positive influence on the recognition of positive emotions [141,142]; (ii) they regulate both innate and adaptive immunity towards an antimicrobial, anti-inflammatory and tumor surveillance phenotype via kynurenine [143]; (iii) they enhance the function of the intestinal epithelial barrier via indoles [144] and inhibit LPS-induced pro-inflammatory interleukin expression through Aryl hydrocarbon receptor (AhR) signaling [145]; (iv) they improve insulin resistance [146] and lipid metabolism, reducing liver steatosis and inflammation, thus alleviating metabolic dysfunction-associated steatotic liver disease (MASLD) [147,148,149].

2.9. Conjugated Linoleic Acids (CLAs)

Conjugated linoleic acids (CLAs) refer to a pool of cis or trans isomers of the polyunsaturated omega-6 essential fatty acid linoleic acid, cis-9, cis-12 and octadecadienoic acid being the most represented ones (almost 95% of all linoleic acid isomers) [150]. CLAs derive from the biohydrogenation of linoleic acid carried by bacteria that express linoleic acid isomerase, such as Bifidobacterium breve, Bifidobacterium infantis, Bifidobacterium adolescentis, Lactobacillus reuteri, Roseburia spp. and others [151,152].
When administered as postbiotics, CLAs show multiple beneficial effects on human health, including: (i) anti-breast cancer properties [153]; (ii) body fat reduction via increased lipolysis and decreased FA accumulation in adipose tissue [154]; (iii) atherosclerosis inhibition [155]; (iv) improved immune system functions and reduced inflammation [156]; (v) osteoporosis prevention [157] and many others [158].

2.10. Polyamines

Polyamines are organic polycationic alkylamines synthesized from L-ornithine and/or arginine or by amino acid decarboxylation that play important roles in a huge variety of biological functions in all organisms, from cell metabolism to apoptosis and cell differentiation [159,160]. Among all, putrescine (PUT), spermine (SPE), spermidine (SPD) and cadaverine (CAD) are the most important ones and recent studies demonstrated their role as NLRP6 inflammasome inhibitors [161,162].
Interestingly, the human gut microbiota, with bacteria belonging to Bifidobacterium, Clostridium, Enterococcus, Lactobacillus and Enterobacter expressing arginine decarboxylase and/or ornithine decarboxylase, is a major contributor to the total polyamine pool in vivo [163,164].
Polyamines are essential for intestinal epithelial renewal and barrier integrity and homeostasis, acting through both transcriptional and posttranscriptional control of expression of multiple genes involved in intestinal epithelial cell (IEC) proliferation, migration and cell-to-cell interactions [165]. PAs are also fundamental for a proper immune system development, since PA depletion causes abnormal differentiation of cytolytic T lymphocytes and defective immunoglobulin-producing B cells [166,167,168]. In addition, SPE is able to reverse B cell senescence [169].
Despite that, elevated levels of polyamine can inhibit immune cell activity and have been associated with tumorigenesis, in particular with breast, colon, prostate and skin cancers [170,171].

2.11. Phenolic Compounds

Phenolic compounds (or polyphenols) are a heterogeneous group of natural bioactive molecules defined as secondary metabolites mainly found in plant tissues and generated during plant metabolism that play a pivotal role in protecting from UV radiations and pathogen aggression [172]. Polyphenols are largely found in fruits, vegetables and cereals, thus representing an important component of our diet [173].
After dietary ingestion, polyphenols are metabolized by the human gut microbiota (e.g., Eubacterium ramulus, Lactobacillus spp. and Gordonibacter urolothinfaciens) via multiple biotransformation processes such as esterification, glycosylation, hydrolysis and acylation [174]. Urolithins, derived from the microbiota transformation of ellagitannins (ETs) and ellagic acid (EA), represent one of the most common and important polyphenol-derived group of metabolites which have drawn the attention of the scientific community for the last few years for their pleiotropic health beneficial effects in preventing several conditions such as cardiovascular diseases (CVDs), diabetes, aging, asthma and infectious diseases thanks to their well-known antioxidant, anti-inflammatory, neuroprotective and cardioprotective effects [175,176,177].

2.12. Hydrogen Peroxide (H2O2)

Hydrogen peroxide (H2O2) is an endogenous reactive oxygen species (ROS) which naturally occurs as a byproduct of cellular respiration [178]. H2O2 contributes to oxidative stress both directly, acting as a molecular oxidant (e.g., peroxidation of membrane lipids which leads to membrane integrity disruption), and indirectly, through free radical generation, which penetrates cell membranes and reacts with intracellular molecules [179]. Moreover, in concentrations from 1% to 6%, H2O2 has antimicrobial properties [180].
Lactobacillaceae and their H2O2 production represent one of the most important mechanisms of colonization resistance against pathogen microbes, thus making it possible to be considered an interkingdom antivirulence strategy [181].
However, given its oxidizing activity, H2O2 is responsible for single- and double-strand DNA breaks and it seems to play a pivotal role in mutagenesis and tumorigenesis of thyroid cells, specifically when proper antioxidant defenses are lacking [182]. Moreover, in vitro studies suggest that exposure of cortical neural cells to H2O2 is toxic, being responsible for increased intracellular free calcium concentration and apoptotic cell death within 3 h [183]. Lastly, elevated levels of H2O2 cause a decrease in gap junction (GJ) resistance and well as a reduction in intracellular pH, leading to acidosis [184].

2.13. Organic Acids

Organic acids are a heterogenous class of low-molecular-weight (LMW) compounds containing at least one carboxylic acid group which are intermediate products of several cellular catabolic pathways including glycolysis, tricarboxylic acid (TCA) cycle and FA oxidation [185,186].
Similarly to H2O2, organic acids such as formic acid (FA), mainly produced by Lactobacillus spp. [187], are suitable as antibacterial agents by acting as inhibitory metabolites through colonization resistance mechanisms by preventing pathogens’ enzymes from working properly [188,189].

2.14. Glutathione (GSH)

γ-L-glutamyl-L-cysteinyl-glycine, known as glutathione (GSH), is one of the most important LMW antioxidant compounds produced by the cell. It is obtained by the sequential addition of cysteine and glycine to a glutamate molecule and its potential is majorly due to the sulfhydryl group (-SH) of the cysteine residue which is involved in reduction and conjugation reactions, making GSH essential for peroxide removal and xenobiotic metabolism [190,191].
Recent studies demonstrated how Lactobacillus salivarius can enhance GSH de novo synthesis, which in turn inhibits mitochondrial biogenesis in osteoclasts (OCs), thus representing an interesting approach for the treatment and prevention of osteoporosis [192]. Due to its antioxidant activity, GSH supplementation may be used as an anti-inflammatory and immunomodulatory compound [193,194,195].

2.15. Microbial Enzymes

Enzymes are proteins required to accelerate metabolic processes by decreasing activation energy for a chemical reaction to occur. This allows to speed up reaction rates; reactions that, otherwise, would not be time-compatible with physiological biological timelines [196].
Microbial enzymes are particularly interesting due to economic feasibility, high yields, rapid bacteria growth rates and inexpensive culture media as well as greater catalytic activity [197]. Microbial enzymes mediate several metabolic, physiological and regulatory processes and are able to resist to unusual temperatures and pH conditions, making them attractive not only for medical but also for industrial applications [198]. Some of the most common enzymes used as postbiotics are proteases and lipases.
Proteases, mainly synthesized by Lactobacillus spp. bacteria, not only participate in protein-rich food digestion: they have been shown to be responsible for the generation of bioactive peptides (2–20 amino acids) with immunomodulatory and anticancer activities [199].
Lipases, produced by bacterial Bacillus spp., Alcaligens spp., Pseudomonas spp. and fungi Penicillium spp. and Aspergillus spp., may act as postbiotics by exerting multiple effects via antioxidant influence, antimicrobial and lipolytic action [200]. Lipases may help in reducing inflammation via lipid metabolism modulation in conditions such as obesity or metabolic syndrome [201,202].

3. Secondary Bile Acids as Postbiotics

Bile acids are amphipathic molecules derived by cholesterol conversion via a chain of enzymatic reactions within hepatocytes: as a result, primary bile acids chenodeoxycholic acid (CDCA) and cholic acid (CA) are obtained and then conjugated with glycine (G) or taurine (T) residues to give rise to their respective bile salts which are finally secreted into the bile [203,204,205].
Other than being involved in dietary lipid emulsion and adsorption, once they have reached the small intestine, primary bile acids are processed and metabolized into secondary bile acids by intestinal microbial enzymatic activity [206]: such biotransformations are mainly represented by deamination (or deconjugation), carried by the Bile Salt Hydrolase (BSH) abundantly expressed by Lactobacillus, Bifidobacterium, Enterococcus and Clostridium species [207], and epimerization, carried out by hydroxysteroid dehydrogenases (HSDHs) [208]. While LCA, DCA and ursodeoxycholic acid (UDCA) are the best characterized secondary bile acids, actually, more than 692 novel bile acids [209], including over 200 microbiota-derived secondary bile acids [210,211] (MDBA), have been identified [209,212].
Thanks to their ability to bind a heterogeneous family of both membrane and nuclear receptors referred to as bile-acid-regulated receptors (BARRs) [213], which are ubiquitously expressed by different cell types of the human body such as enterocytes, hepatocytes, neurons, adipocytes and immune cells [214], secondary bile acids represent the most abundant families of chemical metabolites able to mediate mutual interactions between the intestinal microbiota and the host, regulating immune system, glucose and energy metabolism [215,216,217]. The two best characterized BARRs are FXR [218] and GPBAR1 (also known as TGR5) [219]. FXR functions as a bile acid sensor [94,220], regulates bile acid synthesis and homeostasis, and is mainly activated by primary bile acids [221]. In contrast, GPBAR1 [219] regulates energy expenditure [222] and glucose metabolism and is preferentially activated by secondary bile acids. In addition to these metabolic effects, both FXR and GPBAR1 exert immunoregulatory effects [203,213,223] in the liver, intestine [224] and cardiovascular system [225] (Figure 3).

3.1. Secondary Bile Acids and Immunity

Due to the expression of different BARRs by both innate and adaptive immunity cells, secondary bile acids are responsible for the development of a tolerogenic immune system. Specifically, different secondary bile acids have different affinities for various receptors, each one responsible for the activation of specific pathways (Table 5).
DCA and LCA are the main physiological ligands of GPBAR1 in humans [219]. GPBAR1 is expressed by monocytes, macrophages, DCs and natural killer T (NKT) cells [267]. GPBAR1 activation upon bile acids binding on these cells promotes the development of a tolerogenic phenotype in the immune system via different mechanisms: (i) acting as a negative regulator of the pro-inflammatory NF-κB pathway by inhibiting IκBα phosphorylation and p65 nuclear translocation [268,269]; (ii) inducing CREB phosphorylation, responsible for NF-κB-responsive element repression [270]; (iii) inhibiting NLR family pyrin domain containing 3 (NLRP3) inflammasome, thus preventing the secretion of pro-inflammatory mediators like IL-6, IL-1 β, TNF-α [232]; (iv) stimulating, in monocytes and macrophages, the secretion of the IL-10 anti-inflammatory cytokine [226].
LCA and its derivative 3-keto-LCA are ligands for the Pregnane-X-Receptor (PXR), expressed by monocytes, macrophages, CD4+, CD8+ and B cells [271]. Similarly to GPBAR1, PXR activation inhibits NF-κB and NLRP3 inflammasome assembly [272]. Also, LCA binds and activates the Constitutive Androstane Receptor (CAR), predominantly expressed by T cells, inducing effector T cell reprogramming, IL-10 secretion increase and Treg cell pool expansion [273]. In addition, Vitamin D Receptor (VDR) is activated by LCA and its derivatives, too, reducing pro-inflammatory cytokine expression in monocytes, macrophages and Kupffer cells (KCs) while promoting Tregs expansion by increasing FOXP3 expression [203,252].
Simultaneously, LCA, DCA and their derivatives (3-oxo-LCA, iso-allo-LCA and iso-allo-DCA) contribute to the development of a tolerogenic reprogramming of the immune system by acting as inverse agonists for the Retinoid Orphan Receptor gamma T (RORγt), a typical transcription factor responsible for Th17 differentiation [274,275], thus inhibiting Th17 differentiation while promoting Treg expansion via FOXP3 expression [276,277].

3.2. Secondary Bile Acids and Glucose Metabolism

Intestinal FXR and GPBAR1 expression is essential for a proper regulation of glucose metabolism via the secretion of two enterokines, the Fibroblast Growth Factor (FGF)-19 [278] and Glucagon-Like Peptide (GLP)-1 [203,279]. In post-prandial conditions, primary and secondary bile acids form in the gastrointestinal tract bind and activate intestinal FXR [280,281,282] and GPBAR1 [283], promoting the release of FGF-19 from ileal enterocytes and GLP-1 from ileal and colonic L cells, respectively. Binding to its FGF receptor (FGFR) on hepatocytes, FGF-19 acts as a CYP7A1 repressor, thus contributing to the feed-back inhibition of bile acid synthesis [284,285]. Various FXR agonists are currently developed for clinical applications [220,286,287,288], although animal studies seem to suggest that FXR antagonists [289,290,291] might also have a potential therapeutic utility.
On the other hand, GLP-1 stimulates insulin secretion from pancreatic ϐ-cells promoting glucose uptake, delays in gastric emptying and appetite suppression [292]. Moreover, recent studies showed that pancreatic ϐ-cells express, themselves, both FXR and GPBAR1, thus making it possible for primary and secondary bile acids to directly induce insulin transcription and secretion [293,294].

3.3. Secondary Bile Acids as Exercise Mimetic and Longevity-Associated Molecules

The composition of intestinal microbiota is influenced by physical activity and dietary intake. In general, physical activity reduces the beta-diversity of gut microbiota composition and postbiotic production, particularly SCFAs. However, in addition to SCFAs, secondary bile acids are increasingly recognized as potent modulators of energy expenditure and metabolism.
GPBAR1 is robustly expressed in thermogenic competent tissues including striated muscle and white (WAT) and brown (BAT) adipose tissues [222,295]. In these tissues, GPBAR1 agonism by DCA and LCA promotes a cAMP-dependent expression of the type 2 iodothyronine deiodinase (DIO2), a thyroid hormone activating enzyme responsible for tetraiodothyronine (T4) conversion into active tri-iodothyronine (T3) [222]. T3 binds and activates the thyroid hormone receptor (THR) that acts as a transcription factor for various genes, increasing both energy expenditure via thermogenesis and basal metabolic rate. A key transcription factor involved in this effect is the Uncoupling Protein 1 (UCP1) [296].
UCP1 plays a crucial role as an exercise mimetic by mimicking some of the metabolic and thermogenic benefits of physical activity, particularly through its role in energy expenditure and metabolic regulation [296]. UCP1 is a mitochondrial protein primarily expressed in the BAT and beige fat, where it dissipates the proton gradient to generate heat instead of adenosine triphosphate (ATP). This process, known as non-shivering thermogenesis, mimics the metabolic effects of exercise by increasing caloric burn and lipid oxidation. UCP1 activation improves whole-body glucose metabolism and protects against insulin resistance, resembling the effects of endurance training. In muscle cells, UCP1 enhances mitochondrial biogenesis and functions in skeletal muscle; targeting UCP1 pharmacologically (e.g., with β3-adrenergic agonists or cold exposure) is an area of interest for combating obesity and metabolic diseases in individuals who are unable to exercise. The two main secondary bile acids DCA and LCA and their isoforms exert metabolic effects that mimic exercise by activating GPBAR1 and FXR signaling (including UCP1), promoting mitochondrial function, enhancing glucose metabolism and reducing inflammation. While they do not replace physical activity, pharmacological modulation of secondary bile acid signaling could serve as a therapeutic strategy for metabolic disorders, obesity and age-related metabolic decline, especially for individuals unable to exercise.
This view is further supported by the observation that secondary bile acids, specifically LCA, might be the mediator of beneficial effects exerted by calorie restriction on body weight and life span [16]. A recent study has shown that LCA accumulates in muscles and BAT and WAT in response to calorie restriction, while its administration promotes body weight reduction via activated protein kinase (AMPK)-dependent pathways [297]. Recent evidence suggests that LCA is able to recapitulate, at least in animal models, all the beneficial effects of calorie restriction [16,298], a dietary intervention that can promote overall health and lifespan extension through the reduction of inflammation and reactive oxygen species (ROS) production, generally altered in age-related metabolic disorders and immune mediated diseases [299,300,301]. Qu et al. demonstrated that LCA administration activates AMPK in muscle cells, enhancing muscle regeneration and strength in old mice as well as inducing life extension in Caenorhabditis elegans and Drosophila melanogaster, all effects abrogated after knocking-down AMPK [16]. Additionally, LCA enhances mitochondrial respiration and reduces reactive oxygen species (ROS), thereby improving cellular energy metabolism and expression of mitochondrial unfolded protein response (UPRmt) in C. elegans. It has also been shown that LCA might promote a sirtuin-dependent activation of AMPK. LCA induces sirtuin 3, a mitochondrial deacetylase that enhances energy metabolism and oxidative stress resistance. It is important to consider that while LCA could extend the lifespan of yeast and Caenorhabditis elegans, evidence in higher organisms such as mice and humans is still limited.
In addition to sirtuin 3, Qu et al. have shown that LCA activates sirtuin 1 [298] and identified TUB-like protein 3 (TULP3) as receptors for LCA. Specifically, LCA binding to TULP3 induces the allosteric activation of sirtuins, which subsequently deacetylates the V1E1 subunit of v-ATPase at residues K52, K99 and K191, promoting a robust activation of AMPK and muscle function in aged mice.
Partially in accordance with these results, a study carried out in 2021 in Japanese centenarians detected elevated fecal levels of LCA and its metabolites. Given the potent antimicrobial effects of iso-LCA, it has been proposed that this could contribute to centenarians’ long-lasting health [208]. Nevertheless, it is important to remember that the high iso-allo-LCA hydrophobicity has been associated with an increased risk for age-related cognitive impairment [302].
While these studies suggest that secondary bile acids have beneficial effects on metabolism, and the majority of reports published in the last decade envision LCA and DCA and their receptors as potential targets to treat metabolic and inflammatory disorders, there is older literature that raises concerns over the potential harmful effects of these relative hydrophobic bile acids [303]. Bile acids have been reported to promote inflammation, rather than reducing the activation of inflammatory cells including direct activation of inflammasome. However, some of these effects are obtained at relatively high concentrations, approximatively 100 µM or higher, while the EC50 for the activation of nuclear and G-protein coupled receptors are generally in nanomolar or low micromolar ranges.
In addition to these beneficial effects, early evidence has also raised some concerns over the possibility that secondary bile acids might be detrimental for the human intestine and might exert a role in the development of intestinal injury and cancers. High concentrations of bile acids promote cell damage, oxidative stress, ROS production and DNA damage [304], and have been considered as putative etiologic agents in the development of gastrointestinal cancers, including esophageal, gastric, liver, bile duct, pancreatic and colorectal cancers [305]. Consistent with this view, both prospective and retrospective studies in humans associate high circulating blood levels of secondary bile acids with increased risks of colon cancer [306]. However, it should be noted that more recent studies carried out in IBD patients have shown that IBD development associates with intestinal dysbiosis and reduced excretion of secondary bile acids [223,307]. Further on, secondary bile acids restrain intestinal inflammation in IBD patients and models, by GPBAR1, VDR and RORγT-dependent mechanisms [274].

4. Conclusions

The intestinal microbiota is a source of an extraordinary variety of bile acids. Gut bacteria transform primary bile acids into secondary bile acids, which then act as metabolic regulators through GPBAR1, PXR and RORγT and other canonical receptors such as VDR [224]. A balanced microbiota promotes healthy metabolism, while intestinal dysbiosis might participate in the development of several human disorders including obesity, diabetes and systemic inflammation. By modifying the intestinal microbiota or bile acid pathways, it might be possible to prevent and treat metabolic diseases. Secondary bile acids modulate inflammation, instructing the host immune system and regulating glucose and energy metabolism, making it possible to define them as one of the most interesting classes of postbiotics. Secondary bile acids are emerging as potent regulators of metabolism, inflammation, and immunity. By modifying the intestinal microbiota or designing bile acid-based drugs, it will be possible to design novel therapies for metabolic disorders, inflammatory diseases. These bile-acid–microbiota interactions’ growth could lead to transformative treatments for a wide range of human diseases.

Author Contributions

G.U. and S.F., conceptualization and writing; E.R., E.D., S.M. and M.B., writing and editing. All authors have equally contributed to the conception, data preparation and writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Salminen, S.; Collado, M.C.; Endo, A.; Hill, C.; Lebeer, S.; Quigley, E.M.M.; Sanders, M.E.; Shamir, R.; Swann, J.R.; Szajewska, H.; et al. The International Scientific Association of Probiotics and Prebiotics (ISAPP) consensus statement on the definition and scope of postbiotics. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 649–667. [Google Scholar] [CrossRef] [PubMed]
  2. Tsilingiri, K.; Rescigno, M. Postbiotics: What else? Benef. Microbes 2013, 4, 101–107. [Google Scholar] [CrossRef] [PubMed]
  3. Aguilar-Toalá, J.E.; Garcia-Varela, R.; Garcia, H.S.; Mata-Haro, V.; González-Córdova, A.F.; Vallejo-Cordoba, B.; Hernández-Mendoza, A. Postbiotics: An evolving term within the functional foods field. Trends Food Sci. Technol. 2018, 75, 105–114. [Google Scholar] [CrossRef]
  4. Collado, M.C.; Vinderola, G.; Salminen, S. Postbiotics: Facts and open questions. A position paper on the need for a consensus definition. Benef. Microbes 2019, 10, 711–719. [Google Scholar] [CrossRef]
  5. Faintuch, J.; Faintuch, S. (Eds.) Microbiome and Metabolome in Diagnosis, Therapy, and Other Strategic Applications; Academic Press: London, UK, 2019; Available online: http://lib.ugent.be/catalog/ebk01:4100000007427446 (accessed on 1 February 2025).
  6. Johnson, C.N.; Kogut, M.H.; Genovese, K.; He, H.; Kazemi, S.; Arsenault, R.J. Administration of a Postbiotic Causes Immunomodulatory Responses in Broiler Gut and Reduces Disease Pathogenesis Following Challenge. Microorganisms 2019, 7, 268. [Google Scholar] [CrossRef]
  7. Wegh, C.A.M.; Geerlings, S.Y.; Knol, J.; Roeselers, G.; Belzer, C. Postbiotics and Their Potential Applications in Early Life Nutrition and Beyond. Int. J. Mol. Sci. 2019, 20, 4673. [Google Scholar] [CrossRef]
  8. Vinderola, G.; Sanders, M.E.; Cunningham, M.; Hill, C. Frequently asked questions about the ISAPP postbiotic definition. Front. Microbiol. 2023, 14, 1324565. [Google Scholar] [CrossRef]
  9. Zhao, X.; Liu, S.; Li, S.; Jiang, W.; Wang, J.; Xiao, J.; Chen, T.; Ma, J.; Khan, M.Z.; Wang, W.; et al. Unlocking the power of postbiotics: A revolutionary approach to nutrition for humans and animals. Cell Metab. 2024, 36, 725–744. [Google Scholar] [CrossRef]
  10. Chaluvadi, S.; Hotchkiss, A.T.; Yam, K.L. Gut microbiota: Impact of probiotics, prebiotics, synbiotics, pharmabiotics, and postbiotics on human health. In Probiotics, Prebiotics, and Synbiotics: Bioactive Foods in Health Promotion; Elsevier Inc.: Amsterdam, The Netherlands, 2015; pp. 515–523. [Google Scholar]
  11. Williams, N.T. Probiotics. Am. J. Health Pharm. AJHP Off. J. Am. Soc. Health Pharm. 2010, 67, 449–458. [Google Scholar] [CrossRef]
  12. Valcheva, R.; Dieleman, L.A. Prebiotics: Definition and protective mechanisms. Best Pract. Res. Clin. Gastroenterol. 2016, 30, 27–37. [Google Scholar] [CrossRef]
  13. Cicenia, A.; Scirocco, A.; Carabotti, M.; Pallotta, L.; Marignani, M.; Severi, C. Postbiotic activities of lactobacilli-derived factors. J. Clin. Gastroenterol. 2014, 48 (Suppl. S1), S18–S22. [Google Scholar] [CrossRef]
  14. Hernández-Granados, M.J.; Franco-Robles, E. Postbiotics in human health: Possible new functional ingredients? Food Res. Int. 2020, 137, 109660. [Google Scholar] [CrossRef] [PubMed]
  15. Sousa, T.; Castro, R.E.; Pinto, S.N.; Coutinho, A.; Lucas, S.D.; Moreira, R.; Rodrigues, C.M.P.; Prieto, M.; Fernandes, F. Deoxycholic acid modulates cell death signaling through changes in mitochondrial membrane properties. J. Lipid Res. 2015, 56, 2158–2171. [Google Scholar] [CrossRef]
  16. Qu, Q.; Chen, Y.; Wang, Y.; Long, S.; Wang, W.; Yang, H.-Y.; Li, M.; Tian, X.; Wei, X.; Liu, Y.-H.; et al. Lithocholic acid phenocopies anti-ageing effects of calorie restriction. Nature 2024, 638, E6. [Google Scholar] [CrossRef]
  17. Sheng, W.; Ji, G.; Zhang, L. The Effect of Lithocholic Acid on the Gut-Liver Axis. Front. Pharmacol. 2022, 13, 910493. [Google Scholar] [CrossRef]
  18. Houtman, T.A.; Eckermann, H.A.; Smidt, H.; de Weerth, C. Gut microbiota and BMI throughout childhood: The role of firmicutes, bacteroidetes, and short-chain fatty acid producers. Sci. Rep. 2022, 12, 3140. [Google Scholar] [CrossRef] [PubMed]
  19. Mortensen, P.B.; Clausen, M.R. Short-chain fatty acids in the human colon: Relation to gastrointestinal health and disease. Scand. J. Gastroenterol. Suppl. 1996, 216, 132–148. [Google Scholar] [CrossRef] [PubMed]
  20. Topping, D.L.; Clifton, P.M. Short-chain fatty acids and human colonic function: Roles of resistant starch and nonstarch polysaccharides. Physiol. Rev. 2001, 81, 1031–1064. [Google Scholar] [CrossRef]
  21. Ríos-Covián, D.; Ruas-Madiedo, P.; Margolles, A.; Gueimonde, M.; de Los Reyes-Gavilán, C.G.; Salazar, N. Intestinal Short Chain Fatty Acids and their Link with Diet and Human Health. Front. Microbiol. 2016, 7, 185. [Google Scholar] [CrossRef]
  22. Cuervo, A.; Salazar, N.; Ruas-Madiedo, P.; Gueimonde, M.; González, S. Fiber from a regular diet is directly associated with fecal short-chain fatty acid concentrations in the elderly. Nutr. Res. 2013, 33, 811–816. [Google Scholar] [CrossRef]
  23. Fernandes, J.; Su, W.; Rahat-Rozenbloom, S.; Wolever, T.M.S.; Comelli, E.M. Adiposity, gut microbiota and faecal short chain fatty acids are linked in adult humans. Nutr. Diabetes 2014, 4, e121. [Google Scholar] [CrossRef] [PubMed]
  24. Rahat-Rozenbloom, S.; Fernandes, J.; Gloor, G.B.; Wolever, T.M.S. Evidence for greater production of colonic short-chain fatty acids in overweight than lean humans. Int. J. Obes. 2014, 38, 1525–1531. [Google Scholar] [CrossRef] [PubMed]
  25. Patil, D.P.; Dhotre, D.P.; Chavan, S.G.; Sultan, A.; Jain, D.S.; Lanjekar, V.B.; Gangawani, J.; Shah, P.S.; Todkar, J.S.; Shah, S.; et al. Molecular analysis of gut microbiota in obesity among Indian individuals. J. Biosci. 2012, 37, 647–657. [Google Scholar] [CrossRef] [PubMed]
  26. Salli, K.; Hirvonen, J.; Anglenius, H.; Hibberd, A.A.; Ahonen, I.; Saarinen, M.T.; Maukonen, J.; Ouwehand, A.C. The Effect of Human Milk Oligosaccharides and Bifidobacterium longum subspecies infantis Bi-26 on Simulated Infant Gut Microbiome and Metabolites. Microorganisms 2023, 11, 1553. [Google Scholar] [CrossRef]
  27. Kijner, S.; Kolodny, O.; Yassour, M. Human milk oligosaccharides and the infant gut microbiome from an eco-evolutionary perspective. Curr. Opin. Microbiol. 2022, 68, 102156. [Google Scholar] [CrossRef]
  28. Abdugheni, R.; Wang, W.-Z.; Wang, Y.-J.; Du, M.-X.; Liu, F.-L.; Zhou, N.; Jiang, C.-Y.; Wang, C.-Y.; Wu, L.; Ma, J.; et al. Metabolite profiling of human-originated Lachnospiraceae at the strain level. iMeta 2022, 1, e58. [Google Scholar] [CrossRef]
  29. Xie, J.; Li, L.-F.; Dai, T.-Y.; Qi, X.; Wang, Y.; Zheng, T.-Z.; Gao, X.-Y.; Zhang, Y.-J.; Ai, Y.; Ma, L.; et al. Short-Chain Fatty Acids Produced by Ruminococcaceae Mediate α-Linolenic Acid Promote Intestinal Stem Cells Proliferation. Mol. Nutr. Food Res. 2022, 66, e2100408. [Google Scholar] [CrossRef]
  30. Zhou, Y.; Xu, H.; Xu, J.; Guo, X.; Zhao, H.; Chen, Y.; Zhou, Y.; Nie, Y.F. prausnitzii and its supernatant increase SCFAs-producing bacteria to restore gut dysbiosis in TNBS-induced colitis. AMB Express 2021, 11, 33. [Google Scholar] [CrossRef]
  31. Flint, H.J.; Duncan, S.H.; Scott, K.P.; Louis, P. Interactions and competition within the microbial community of the human colon: Links between diet and health. Environ. Microbiol. 2007, 9, 1101–1111. [Google Scholar] [CrossRef]
  32. Horvath, T.D.; Ihekweazu, F.D.; Haidacher, S.J.; Ruan, W.; Engevik, K.A.; Fultz, R.; Hoch, K.M.; Luna, R.A.; Oezguen, N.; Spinler, J.K.; et al. Bacteroides ovatus colonization influences the abundance of intestinal short chain fatty acids and neurotransmitters. iScience 2022, 25, 104158. [Google Scholar] [CrossRef]
  33. Chia, L.W.; Mank, M.; Blijenberg, B.; Aalvink, S.; Bongers, R.S.; Stahl, B.; Knol, J.; Belzer, C. Bacteroides thetaiotaomicron Fosters the Growth of Butyrate-Producing Anaerostipes caccae in the Presence of Lactose and Total Human Milk Carbohydrates. Microorganisms 2020, 8, 1513. [Google Scholar] [CrossRef] [PubMed]
  34. Ghosh, T.S.; Shanahan, F.; O’Toole, P.W. The gut microbiome as a modulator of healthy ageing. Nat. Rev. Gastroenterol. Hepatol. 2022, 19, 565–584. [Google Scholar] [CrossRef]
  35. Jeffery, I.B.; Lynch, D.B.; O’Toole, P.W. Composition and temporal stability of the gut microbiota in older persons. ISME J. 2015, 10, 170–182. [Google Scholar] [CrossRef] [PubMed]
  36. Fusco, W.; Lorenzo, M.B.; Cintoni, M.; Porcari, S.; Rinninella, E.; Kaitsas, F.; Lener, E.; Mele, M.C.; Gasbarrini, A.; Collado, M.C.; et al. Short-Chain Fatty-Acid-Producing Bacteria: Key Components of the Human Gut Microbiota. Nutrients 2023, 15, 2211. [Google Scholar] [CrossRef]
  37. Carretta, M.D.; Quiroga, J.; López, R.; Hidalgo, M.A.; Burgos, R.A. Participation of Short-Chain Fatty Acids and Their Receptors in Gut Inflammation and Colon Cancer. Front. Physiol. 2021, 12, 662739. [Google Scholar] [CrossRef] [PubMed]
  38. Ranjbar, R.; Vahdati, S.N.; Tavakoli, S.; Khodaie, R.; Behboudi, H. Immunomodulatory roles of microbiota-derived short-chain fatty acids in bacterial infections. Biomed. Pharmacother. 2021, 141, 111817. [Google Scholar] [CrossRef]
  39. Barcutean, L.; Maier, S.; Burai-Patrascu, M.; Farczadi, L.; Balasa, R. The Immunomodulatory Potential of Short-Chain Fatty Acids in Multiple Sclerosis. Int. J. Mol. Sci. 2024, 25, 3198. [Google Scholar] [CrossRef]
  40. Kim, C.H. Complex regulatory effects of gut microbial short-chain fatty acids on immune tolerance and autoimmunity. Cell. Mol. Immunol. 2023, 20, 341–350. [Google Scholar] [CrossRef]
  41. Goverse, G.; Molenaar, R.; Macia, L.; Tan, J.; Erkelens, M.N.; Konijn, T.; Knippenberg, M.; Cook, E.C.L.; Hanekamp, D.; Veldhoen, M.; et al. Diet-Derived Short Chain Fatty Acids Stimulate Intestinal Epithelial Cells to Induce Mucosal Tolerogenic Dendritic Cells. J. Immunol. 2017, 198, 2172–2181. [Google Scholar] [CrossRef]
  42. Tazoe, H.; Otomo, Y.; Karaki, S.-I.; Kato, I.; Fukami, Y.; Terasaki, M.; Kuwahara, A. Expression of short-chain fatty acid receptor GPR41 in the human colon. Biomed. Res. 2009, 30, 149–156. [Google Scholar] [CrossRef]
  43. Pingitore, A.; Gonzalez-Abuin, N.; Ruz-Maldonado, I.; Huang, G.C.; Frost, G.; Persaud, S.J. Short chain fatty acids stimulate insulin secretion and reduce apoptosis in mouse and human islets in vitro: Role of free fatty acid receptor 2. Diabetes. Obes. Metab. 2019, 21, 330–339. [Google Scholar] [CrossRef] [PubMed]
  44. Priyadarshini, M.; Layden, B.T. FFAR3 modulates insulin secretion and global gene expression in mouse islets. Islets 2015, 7, e1045182. [Google Scholar] [CrossRef] [PubMed]
  45. Kimura, I.; Inoue, D.; Maeda, T.; Hara, T.; Ichimura, A.; Miyauchi, S.; Kobayashi, M.; Hirasawa, A.; Tsujimoto, G. Short-chain fatty acids and ketones directly regulate sympathetic nervous system via G protein-coupled receptor 41 (GPR41). Proc. Natl. Acad. Sci. USA 2011, 108, 8030–8035. [Google Scholar] [CrossRef]
  46. Rangan, P.; Mondino, A. Microbial short-chain fatty acids: A strategy to tune adoptive T cell therapy. J. Immunother. Cancer 2022, 10, e004147. [Google Scholar] [CrossRef] [PubMed]
  47. Kaisar, M.M.M.; Pelgrom, L.R.; van der Ham, A.J.; Yazdanbakhsh, M.; Everts, B. Butyrate Conditions Human Dendritic Cells to Prime Type 1 Regulatory T Cells via both Histone Deacetylase Inhibition and G Protein-Coupled Receptor 109A Signaling. Front. Immunol. 2017, 8, 1429. [Google Scholar] [CrossRef]
  48. Pérez-Reytor, D.; Puebla, C.; Karahanian, E.; García, K. Use of Short-Chain Fatty Acids for the Recovery of the Intestinal Epithelial Barrier Affected by Bacterial Toxins. Front. Physiol. 2021, 12, 650313. [Google Scholar] [CrossRef]
  49. Sun, M.; Wu, W.; Liu, Z.; Cong, Y. Microbiota metabolite short chain fatty acids, GPCR, and inflammatory bowel diseases. J. Gastroenterol. 2017, 52, 1–8. [Google Scholar] [CrossRef]
  50. Tolhurst, G.; Heffron, H.; Lam, Y.S.; Parker, H.E.; Habib, A.M.; Diakogiannaki, E.; Cameron, J.; Grosse, J.; Reimann, F.; Gribble, F.M. Short-chain fatty acids stimulate glucagon-like peptide-1 secretion via the G-protein-coupled receptor FFAR2. Diabetes 2012, 61, 364–371. [Google Scholar] [CrossRef]
  51. Bolognini, D.; Moss, C.E.; Nilsson, K.; Petersson, A.U.; Donnelly, I.; Sergeev, E.; König, G.M.; Kostenis, E.; Kurowska-Stolarska, M.; Miller, A.; et al. A Novel Allosteric Activator of Free Fatty Acid 2 Receptor Displays Unique Gi-functional Bias. J. Biol. Chem. 2016, 291, 18915–18931. [Google Scholar] [CrossRef]
  52. Maslowski, K.M.; Vieira, A.T.; Ng, A.; Kranich, J.; Sierro, F.; Yu, D.; Schilter, H.C.; Rolph, M.S.; Mackay, F.; Artis, D.; et al. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 2009, 461, 1282–1286. [Google Scholar] [CrossRef]
  53. Vinolo, M.A.R.; Ferguson, G.J.; Kulkarni, S.; Damoulakis, G.; Anderson, K.; Bohlooly-Y, M.; Stephens, L.; Hawkins, P.T.; Curi, R. SCFAs induce mouse neutrophil chemotaxis through the GPR43 receptor. PLoS ONE 2011, 6, e21205. [Google Scholar] [CrossRef] [PubMed]
  54. Kespohl, M.; Vachharajani, N.; Luu, M.; Harb, H.; Pautz, S.; Wolff, S.; Sillner, N.; Walker, A.; Schmitt-Kopplin, P.; Boettger, T.; et al. The Microbial Metabolite Butyrate Induces Expression of Th1-Associated Factors in CD4+ T Cells. Front. Immunol. 2017, 8, 1036. [Google Scholar] [CrossRef]
  55. Taggart, A.K.P.; Kero, J.; Gan, X.; Cai, T.-Q.; Cheng, K.; Ippolito, M.; Ren, N.; Kaplan, R.; Wu, K.; Wu, T.-J.; et al. (D)-beta-Hydroxybutyrate inhibits adipocyte lipolysis via the nicotinic acid receptor PUMA-G. J. Biol. Chem. 2005, 280, 26649–26652. [Google Scholar] [CrossRef] [PubMed]
  56. Kostylina, G.; Simon, D.; Fey, M.F.; Yousefi, S.; Simon, H.U. Neutrophil apoptosis mediated by nicotinic acid receptors (GPR109A). Cell Death Differ. 2008, 15, 134–142. [Google Scholar] [CrossRef]
  57. Rooks, M.G.; Garrett, W.S. Gut microbiota, metabolites and host immunity. Nat. Rev. Immunol. 2016, 16, 341–352. [Google Scholar] [CrossRef] [PubMed]
  58. Thangaraju, M.; Cresci, G.A.; Liu, K.; Ananth, S.; Gnanaprakasam, J.P.; Browning, D.D.; Mellinger, J.D.; Smith, S.B.; Digby, G.J.; Lambert, N.A.; et al. GPR109A is a G-protein-coupled receptor for the bacterial fermentation product butyrate and functions as a tumor suppressor in colon. Cancer Res. 2009, 69, 2826–2832. [Google Scholar] [CrossRef]
  59. Gao, Z.; Yin, J.; Zhang, J.; Ward, R.E.; Martin, R.J.; Lefevre, M.; Cefalu, W.T.; Ye, J. Butyrate improves insulin sensitivity and increases energy expenditure in mice. Diabetes 2009, 58, 1509–1517. [Google Scholar] [CrossRef]
  60. Yadav, H.; Lee, J.-H.; Lloyd, J.; Walter, P.; Rane, S.G. Beneficial metabolic effects of a probiotic via butyrate-induced GLP-1 hormone secretion. J. Biol. Chem. 2013, 288, 25088–25097. [Google Scholar] [CrossRef]
  61. Yamashita, H.; Maruta, H.; Jozuka, M.; Kimura, R.; Iwabuchi, H.; Yamato, M.; Saito, T.; Fujisawa, K.; Takahashi, Y.; Kimoto, M.; et al. Effects of acetate on lipid metabolism in muscles and adipose tissues of type 2 diabetic Otsuka Long-Evans Tokushima Fatty (OLETF) rats. Biosci. Biotechnol. Biochem. 2009, 73, 570–576. [Google Scholar] [CrossRef]
  62. Den Besten, G.; Bleeker, A.; Gerding, A.; van Eunen, K.; Havinga, R.; van Dijk, T.H.; Oosterveer, M.H.; Jonker, J.W.; Groen, A.K.; Reijngoud, D.-J.; et al. Short-chain fatty acids protect against high-fat diet–induced obesity via a PPARγ-dependent switch from lipogenesis to fat oxidation. Diabetes 2015, 64, 2398–2408. [Google Scholar] [CrossRef]
  63. Hamadneh, L.; Al-Lakkis, L.; Alhusban, A.A.; Tarawneh, S.; Abu-Irmaileh, B.; Albustanji, S.; Al-Bawab, A.Q. Changes in Lactate Production, Lactate Dehydrogenase Genes Expression and DNA Methylation in Response to Tamoxifen Resistance Development in MCF-7 Cell Line. Genes 2021, 12, 777. [Google Scholar] [CrossRef] [PubMed]
  64. Rabinowitz, J.D.; Enerbäck, S. Lactate: The ugly duckling of energy metabolism. Nat. Metab. 2020, 2, 566–571. [Google Scholar] [CrossRef]
  65. Mokoena, M.P. Lactic Acid Bacteria and Their Bacteriocins: Classification, Biosynthesis and Applications against Uropathogens: A Mini-Review. Molecules 2017, 22, 1255. [Google Scholar] [CrossRef]
  66. Kaban, G.; Kaya, M. Identification of Lactic Acid Bacteria and Gram-Positive Catalase-Positive Cocci Isolated from Naturally Fermented Sausage (Sucuk). J. Food Sci. 2008, 73, M385–M388. [Google Scholar] [CrossRef]
  67. De Vuyst, L.; Leroy, F. Bacteriocins from Lactic Acid Bacteria: Production, Purification, and Food Applications. J. Mol. Microbiol. Biotechnol. 2007, 13, 194–199. [Google Scholar] [CrossRef] [PubMed]
  68. Alakomi, H.L.; Skytta, E.; Saarela, M.; Mattila-Sandholm, T.; Latva-Kala, K.; Helander, I.M. Lactic Acid Permeabilizes Gram-Negative Bacteria by Disrupting the Outer Membrane. Appl. Environ. Microbiol. 2000, 66, 2001–2005. [Google Scholar] [CrossRef] [PubMed]
  69. Reddy, G.; Altaf, M.; Naveena, B.J.; Venkateshwar, M.; Kumar, E.V. Amylolytic bacterial lactic acid fermentation—A review. Biotechnol. Adv. 2008, 26, 22–34. [Google Scholar] [CrossRef]
  70. Nichols, R.G.; Peters, J.M.; Patterson, A.D. Interplay between the host, the human microbiome, and drug metabolism. Hum. Genomics 2019, 13, 27. [Google Scholar] [CrossRef]
  71. Daw, M.A.; Falkiner, F.R. Bacteriocins: Nature, function and structure. Micron 1996, 27, 467–479. [Google Scholar] [CrossRef]
  72. Darbandi, A.; Asadi, A.; Mahdizade Ari, M.; Ohadi, E.; Talebi, M.; Halaj Zadeh, M.; Darb Emamie, A.; Ghanavati, R.; Kakanj, M. Bacteriocins: Properties and potential use as antimicrobials. J. Clin. Lab. Anal. 2022, 36, e24093. [Google Scholar] [CrossRef]
  73. Cotter, P.D.; Ross, R.P.; Hill, C. Bacteriocins—A viable alternative to antibiotics? Nat. Rev. Microbiol. 2013, 11, 95–105. [Google Scholar] [CrossRef] [PubMed]
  74. Marshall, S.H.; Arenas, G. Antimicrobial peptides: A natural alternative to chemical antibiotics and a potential for applied biotechnology. Electron. J. Biotechnol. 2003, 6, 271–284. [Google Scholar] [CrossRef]
  75. Roces, C.; Rodríguez, A.; Martínez, B. Cell Wall-active Bacteriocins and Their Applications Beyond Antibiotic Activity. Probiotics Antimicrob. Proteins 2012, 4, 259–272. [Google Scholar] [CrossRef]
  76. Bonhi, K.L.R.; Imran, S. Role Of Bacteriocin In Tackling The Global Problem Of Multi-Drug Resistance: An Updated Review. Biosci. Biotechnol. Res. Commun. 2019, 12, 601–608. [Google Scholar] [CrossRef]
  77. Benítez-Chao, D.F.; León-Buitimea, A.; Lerma-Escalera, J.A.; Morones-Ramírez, J.R. Bacteriocins: An Overview of Antimicrobial, Toxicity, and Biosafety Assessment by in vivo Models. Front. Microbiol. 2021, 12, 630695. [Google Scholar] [CrossRef] [PubMed]
  78. Breukink, E.; Ganz, P.; De Kruijff, B.; Seelig, J. Binding of nisin Z to bilayer vesicles as determined with isothermal titration calorimetry. Biochemistry 2000, 39, 10247–10254. [Google Scholar] [CrossRef]
  79. Joo, N.E.; Ritchie, K.; Kamarajan, P.; Miao, D.; Kapila, Y.L. Nisin, an apoptogenic bacteriocin and food preservative, attenuates HNSCC tumorigenesis via CHAC1. Cancer Med. 2012, 1, 295–305. [Google Scholar] [CrossRef] [PubMed]
  80. Müller, A.; Ulm, H.; Reder-Christ, K.; Sahl, H.-G.; Schneider, T. Interaction of Type A Lantibiotics with Undecaprenol-Bound Cell Envelope Precursors. Microb. Drug Resist. 2012, 18, 261–270. [Google Scholar] [CrossRef]
  81. Saising, J.; Dube, L.; Ziebandt, A.-K.; Voravuthikunchai, S.P.; Nega, M.; Götz, F. Activity of gallidermin on Staphylococcus aureus and Staphylococcus epidermidis biofilms. Antimicrob. Agents Chemother. 2012, 56, 5804–5810. [Google Scholar] [CrossRef]
  82. Chatterjee, S.; Chatterjee, D.K.; Jani, R.H.; Blumbach, J.; Ganguli, B.N.; Klesel, N.; Limbert, M.; Seibert, G. Mersacidin, a new antibiotic from Bacillus. In vitro and in vivo antibacterial activity. J. Antibiot. 1992, 45, 839–845. [Google Scholar] [CrossRef]
  83. Wu, C.; Biswas, S.; Garcia De Gonzalo, C.V.; van der Donk, W.A. Investigations into the Mechanism of Action of Sublancin. ACS Infect. Dis. 2019, 5, 454–459. [Google Scholar] [CrossRef] [PubMed]
  84. Bastos, M.d.C.d.F.; Coutinho, B.G.; Coelho, M.L.V. Lysostaphin: A Staphylococcal Bacteriolysin with Potential Clinical Applications. Pharmaceuticals 2010, 3, 1139–1161. [Google Scholar] [CrossRef] [PubMed]
  85. García-Vela, S.; Guay, L.-D.; Rahman, M.R.T.; Biron, E.; Torres, C.; Fliss, I. Antimicrobial Activity of Synthetic Enterocins A, B, P, SEK4, and L50, Alone and in Combinations, against Clostridium perfringens. Int. J. Mol. Sci. 2024, 25, 1597. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, C.; Zink, D.L.; Ushio, M.; Burgess, B.; Onishi, R.; Masurekar, P.; Barrett, J.F.; Singh, S.B. Isolation, structure, and antibacterial activity of thiazomycin A, a potent thiazolyl peptide antibiotic from Amycolatopsis fastidiosa. Bioorg. Med. Chem. 2008, 16, 8818–8823. [Google Scholar] [CrossRef]
  87. Morin, N.; Lanneluc, I.; Connil, N.; Cottenceau, M.; Pons, A.M.; Sablé, S. Mechanism of bactericidal activity of microcin L in Escherichia coli and Salmonella enterica. Antimicrob. Agents Chemother. 2011, 55, 997–1007. [Google Scholar] [CrossRef]
  88. Destoumieux-Garzón, D.; Thomas, X.; Santamaria, M.; Goulard, C.; Barthélémy, M.; Boscher, B.; Bessin, Y.; Molle, G.; Pons, A.-M.; Letellier, L.; et al. Microcin E492 antibacterial activity: Evidence for a TonB-dependent inner membrane permeabilization on Escherichia coli. Mol. Microbiol. 2003, 49, 1031–1041. [Google Scholar] [CrossRef]
  89. Begley, M.; Gahan, C.G.M.; Hill, C. The interaction between bacteria and bile. FEMS Microbiol. Rev. 2005, 29, 625–651. [Google Scholar] [CrossRef]
  90. Ridlon, J.M.; Kang, D.J.; Hylemon, P.B. Bile salt biotransformations by human intestinal bacteria. J. Lipid Res. 2006, 47, 241–259. [Google Scholar] [CrossRef]
  91. Chen, I.; Cassaro, S. Physiology, Bile Acids. In StatPearls; StatPearls Publishing LLC: Treasure Island, FL, USA, 2025. [Google Scholar]
  92. Zeng, H.; Umar, S.; Rust, B.; Lazarova, D.; Bordonaro, M. Secondary Bile Acids and Short Chain Fatty Acids in the Colon: A Focus on Colonic Microbiome, Cell Proliferation, Inflammation, and Cancer. Int. J. Mol. Sci. 2019, 20, 1214. [Google Scholar] [CrossRef]
  93. Lenci, I.; Milana, M.; Signorello, A.; Grassi, G.; Baiocchi, L. Secondary bile acids and the biliary epithelia: The good and the bad. World J. Gastroenterol. 2023, 29, 357–366. [Google Scholar] [CrossRef]
  94. Renga, B.; Migliorati, M.; Mencarelli, A.; Cipriani, S.; D’Amore, C.; Distrutti, E.; Fiorucci, S. Farnesoid X receptor suppresses constitutive androstane receptor activity at the multidrug resistance protein-4 promoter. Biochim. Biophys. Acta 2011, 1809, 157–165. [Google Scholar] [CrossRef] [PubMed]
  95. Elmas Demiralp, B. A Postbıotıc Group: Cell Wall Components. 2022. Available online: https://www.researchgate.net/publication/361461471_A_Postbiotic_Group_Cell_Wall_Components (accessed on 3 March 2025).
  96. Xiao, M.; Ren, X.; Yu, Y.; Gao, W.; Zhu, C.; Sun, H.; Kong, Q.; Fu, X.; Mou, H. Fucose-containing bacterial exopolysaccharides: Sources, biological activities, and food applications. Food Chem. X 2022, 13, 100233. [Google Scholar] [CrossRef]
  97. Netrusov, A.I.; Liyaskina, E.V.; Kurgaeva, I.V.; Liyaskina, A.U.; Yang, G.; Revin, V.V. Exopolysaccharides Producing Bacteria: A Review. Microorganisms 2023, 11, 1541. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, Q.; Liu, P.; Peng, J.; Zhao, B.; Cai, J. Postbiotic properties of exopolysaccharide produced by Levilactobacillus brevis M-10 isolated from natural fermented sour porridge through in vitro simulated digestion and fermentation. J. Food Sci. 2024, 89, 3110–3128. [Google Scholar] [CrossRef]
  99. Ma, C.; Zheng, X.; Zhang, Q.; Renaud, S.J.; Yu, H.; Xu, Y.; Chen, Y.; Gong, J.; Cai, Y.; Hong, Y.; et al. A postbiotic exopolysaccharide synergizes with Lactobacillus acidophilus to reduce intestinal inflammation in a mouse model of colitis. Int. J. Biol. Macromol. 2024, 291, 138931. [Google Scholar] [CrossRef]
  100. Galinier, A.; Delan-Forino, C.; Foulquier, E.; Lakhal, H.; Pompeo, F. Recent Advances in Peptidoglycan Synthesis and Regulation in Bacteria. Biomolecules 2023, 13, 720. [Google Scholar] [CrossRef]
  101. Garde, S.; Chodisetti, P.K.; Reddy, M. Peptidoglycan: Structure, Synthesis, and Regulation. EcoSal Plus 2021, 9. [Google Scholar] [CrossRef] [PubMed]
  102. Pazos, M.; Peters, K. Peptidoglycan. Subcell. Biochem. 2019, 92, 127–168. [Google Scholar] [CrossRef]
  103. Wu, Z.; Pan, D.; Guo, Y.; Sun, Y.; Zeng, X. Peptidoglycan diversity and anti-inflammatory capacity in Lactobacillus strains. Carbohydr. Polym. 2015, 128, 130–137. [Google Scholar] [CrossRef]
  104. Wang, P.; Wang, S.; Wang, D.; Li, Y.; Yip, R.C.S.; Chen, H. Postbiotics-peptidoglycan, lipoteichoic acid, exopolysaccharides, surface layer protein and pili proteins—Structure, activity in wounds and their delivery systems. Int. J. Biol. Macromol. 2024, 274, 133195. [Google Scholar] [CrossRef]
  105. Kilcullen, J.K.; Ly, Q.P.; Chang, T.H.; Levenson, S.M.; Steinberg, J.J. Nonviable Staphylococcus aureas and its peptidoglycan stimulate macrophage recruitment angiogenesis, fibroplasia, and collagen accumulation in wounded rats. Wound Repair Regen. 1998, 6, 149–156. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, X.; Levenson, S.M.; Chang, T.H.; Steinberg, J.J.; Imegwu, O.; Rojkind, M. Molecular mechanisms underlying wound healing acceleration by Staphylococcus aureus peptidoglycan. Wound Repair Regen. 1996, 4, 470–476. [Google Scholar] [CrossRef]
  107. Burtchett, T.A.; Shook, J.C.; Hesse, L.E.; Delekta, P.C.; Brzozowski, R.S.; Nouri, A.; Calas, A.J.; Spanoudis, C.M.; Eswara, P.J.; Hammer, N.D. Crucial Role for Lipoteichoic Acid Assembly in the Metabolic Versatility and Antibiotic Resistance of Staphylococcus aureus. Infect. Immun. 2023, 91, e0055022. [Google Scholar] [CrossRef]
  108. Ginsburg, I. Role of lipoteichoic acid in infection and inflammation. Lancet Infect. Dis. 2002, 2, 171–179. [Google Scholar] [CrossRef]
  109. Noh, S.Y.; Kang, S.-S.; Yun, C.-H.; Han, S.H. Lipoteichoic acid from Lactobacillus plantarum inhibits Pam2CSK4-induced IL-8 production in human intestinal epithelial cells. Mol. Immunol. 2015, 64, 183–189. [Google Scholar] [CrossRef] [PubMed]
  110. Pradhan, D.; Gulati, G.; Avadhani, R.; Rashmi, H.M.; Soumya, K.; Kumari, A.; Gupta, A.; Dwivedi, D.; Kaushik, J.K.; Grover, S. Postbiotic Lipoteichoic acid of probiotic Lactobacillus origin ameliorates inflammation in HT-29 cells and colitis mice. Int. J. Biol. Macromol. 2023, 236, 123962. [Google Scholar] [CrossRef] [PubMed]
  111. Braverman, N.E.; Moser, A.B. Functions of plasmalogen lipids in health and disease. Biochim. Biophys. Acta 2012, 1822, 1442–1452. [Google Scholar] [CrossRef]
  112. Goldfine, H. Plasmalogens in bacteria, sixty years on. Front. Mol. Biosci. 2022, 9, 962757. Available online: https://www.frontiersin.org/journals/molecular-biosciences/articles/10.3389/fmolb.2022.962757 (accessed on 3 March 2025). [CrossRef]
  113. Zhang, F.; Yang, Z.; Zhou, Y.; Wang, B.; Xie, Z.; Yu, N.; Zhao, J.; Goldfine, H.; Dai, S.; Zhang, G.; et al. Characterization and heterologous expression of plasmalogen synthase MeHAD from Megasphaera elsdenii. Biochim. Biophys. Acta-Mol. Cell Biol. Lipids 2023, 1868, 159358. [Google Scholar] [CrossRef]
  114. Han, X.; Holtzman, D.M.; McKeel Jr, D.W. Plasmalogen deficiency in early Alzheimer’s disease subjects and in animal models: Molecular characterization using electrospray ionization mass spectrometry. J. Neurochem. 2001, 77, 1168–1180. [Google Scholar] [CrossRef]
  115. Dean, J.M.; Lodhi, I.J. Structural and functional roles of ether lipids. Protein Cell 2017, 9, 196–206. [Google Scholar] [CrossRef] [PubMed]
  116. Leßig, J.; Fuchs, B. Plasmalogens in Biological Systems: Their Role in Oxidative Processes in Biological Membranes, their Contribution to Pathological Processes and Aging and Plasmalogen Analysis. Curr. Med. Chem. 2009, 16, 2021–2041. [Google Scholar] [CrossRef]
  117. Gotoh, C.; Hong, Y.-H.; Iga, T.; Hishikawa, D.; Suzuki, Y.; Song, S.-H.; Choi, K.-C.; Adachi, T.; Hirasawa, A.; Tsujimoto, G. The regulation of adipogenesis through GPR120. Biochem. Biophys. Res. Commun. 2007, 354, 591–597. [Google Scholar] [CrossRef]
  118. Engelmann, B. Plasmalogens: Targets for oxidants and major lipophilic antioxidants. Biochem. Soc. Trans. 2004, 32, 147–150. [Google Scholar] [CrossRef] [PubMed]
  119. Fujino, T.; Yamada, T.; Asada, T.; Tsuboi, Y.; Wakana, C.; Mawatari, S.; Kono, S. Efficacy and blood plasmalogen changes by oral administration of plasmalogen in patients with mild Alzheimer’s disease and mild cognitive impairment: A multicenter, randomized, double-blind, placebo-controlled trial. EBioMedicine 2017, 17, 199–205. [Google Scholar] [CrossRef]
  120. Vitamins. Bethesda (MD). 2012. Available online: https://pubmed.ncbi.nlm.nih.gov/31644195/ (accessed on 3 March 2025).
  121. Barker, T. Vitamins and Human Health: Systematic Reviews and Original Research. Nutrients 2023, 15, 2888. [Google Scholar] [CrossRef] [PubMed]
  122. Morris, A.L.; Mohiuddin, S.S. Biochemistry, Nutrients. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  123. Reddy, P.; Jialal, I. Biochemistry, Fat Soluble Vitamins. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  124. Lykstad, J.; Sharma, S. Biochemistry, Water Soluble Vitamins. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  125. Morowitz, M.J.; Carlisle, E.M.; Alverdy, J.C. Contributions of intestinal bacteria to nutrition and metabolism in the critically ill. Surg. Clin. N. Am. 2011, 91, 771–785, viii. [Google Scholar] [CrossRef]
  126. Hill, M.J. Intestinal flora and endogenous vitamin synthesis. Eur. J. Cancer Prev. 1997, 6, S43–S45. [Google Scholar] [CrossRef]
  127. Thorakkattu, P.; Khanashyam, A.C.; Shah, K.; Babu, K.S.; Mundanat, A.S.; Deliephan, A.; Deokar, G.S.; Santivarangkna, C.; Nirmal, N.P. Postbiotics: Current Trends in Food and Pharmaceutical Industry. Foods 2022, 11, 3094. [Google Scholar] [CrossRef]
  128. Tanaka, K.; Taketani, Y.; Angeles-Agdeppa, I.; Kambe, T. Role of Vitamins and Minerals in Health and Diseases. J. Nutr. Sci. Vitaminol. 2022, 68, S70–S72. [Google Scholar] [CrossRef]
  129. Kiani, A.K.; Dhuli, K.; Donato, K.; Aquilanti, B.; Velluti, V.; Matera, G.; Iaconelli, A.; Connelly, S.T.; Bellinato, F.; Gisondi, P.; et al. Main nutritional deficiencies. J. Prev. Med. Hyg. 2022, 63, E93–E101. [Google Scholar] [CrossRef] [PubMed]
  130. Paeslack, N.; Mimmler, M.; Becker, S.; Gao, Z.; Khuu, M.P.; Mann, A.; Malinarich, F.; Regen, T.; Reinhardt, C. Microbiota-derived tryptophan metabolites in vascular inflammation and cardiovascular disease. Amino Acids 2022, 54, 1339–1356. [Google Scholar] [CrossRef] [PubMed]
  131. Zuraikat, F.M.; Wood, R.A.; Barragán, R.; St-Onge, M.-P. Sleep and Diet: Mounting Evidence of a Cyclical Relationship. Annu. Rev. Nutr. 2021, 41, 309–332. [Google Scholar] [CrossRef] [PubMed]
  132. Sainio, E.-L.; Pulkki, K.; Young, S.N. L-Tryptophan: Biochemical, nutritional and pharmacological aspects. Amino Acids 1996, 10, 21–47. [Google Scholar] [CrossRef]
  133. Palmieri, A.; Petrini, M. Tryptophol and derivatives: Natural occurrence and applications to the synthesis of bioactive compounds. Nat. Prod. Rep. 2019, 36, 490–530. [Google Scholar] [CrossRef]
  134. Barik, S. The Uniqueness of Tryptophan in Biology: Properties, Metabolism, Interactions and Localization in Proteins. Int. J. Mol. Sci. 2020, 21, 8776. [Google Scholar] [CrossRef]
  135. Richard, D.M.; Dawes, M.A.; Mathias, C.W.; Acheson, A.; Hill-Kapturczak, N.; Dougherty, D.M. L-Tryptophan: Basic Metabolic Functions, Behavioral Research and Therapeutic Indications. Int. J. Tryptophan Res. 2009, 2, 45–60. [Google Scholar] [CrossRef]
  136. Ardis, T.C.; Cahir, M.; Elliott, J.J.; Bell, R.; Reynolds, G.P.; Cooper, S.J. Effect of acute tryptophan depletion on noradrenaline and dopamine in the rat brain. J. Psychopharmacol. 2009, 23, 51–55. [Google Scholar] [CrossRef]
  137. Liu, H.-W.; Shi, B.-M.; Liu, D.-S.; Shan, A.-S. Supplemental dietary tryptophan modifies behavior, concentrations of salivary cortisol, plasma epinephrine, norepinephrine and hypothalamic 5-hydroxytryptamine in weaning piglets. Livest. Sci. 2013, 151, 213–218. [Google Scholar] [CrossRef]
  138. Lee, J.-H.; Lee, J. Indole as an intercellular signal in microbial communities. FEMS Microbiol. Rev. 2010, 34, 426–444. [Google Scholar] [CrossRef]
  139. Vieira-Silva, S.; Falony, G.; Darzi, Y.; Lima-Mendez, G.; Garcia Yunta, R.; Okuda, S.; Vandeputte, D.; Valles-Colomer, M.; Hildebrand, F.; Chaffron, S.; et al. Species–function relationships shape ecological properties of the human gut microbiome. Nat. Microbiol. 2016, 1, 16088. [Google Scholar] [CrossRef] [PubMed]
  140. Tennoune, N.; Andriamihaja, M.; Blachier, F. Production of Indole and Indole-Related Compounds by the Intestinal Microbiota and Consequences for the Host: The Good, the Bad, and the Ugly. Microorganisms 2022, 10, 930. [Google Scholar] [CrossRef]
  141. Zamoscik, V.; Schmidt, S.N.L.; Bravo, R.; Ugartemendia, L.; Plieger, T.; Rodríguez, A.B.; Reuter, M.; Kirsch, P. Tryptophan-enriched diet or 5-hydroxytryptophan supplementation given in a randomized controlled trial impacts social cognition on a neural and behavioral level. Sci. Rep. 2021, 11, 21637. [Google Scholar] [CrossRef] [PubMed]
  142. Miura, H.; Ozaki, N.; Sawada, M.; Isobe, K.; Ohta, T.; Nagatsu, T. A link between stress and depression: Shifts in the balance between the kynurenine and serotonin pathways of tryptophan metabolism and the etiology and pathophysiology of depression. Stress 2008, 11, 198–209. [Google Scholar] [CrossRef] [PubMed]
  143. Munn, D.H.; Mellor, A.L. Indoleamine 2,3 dioxygenase and metabolic control of immune responses. Trends Immunol. 2013, 34, 137–143. [Google Scholar] [CrossRef]
  144. Hou, Q.; Ye, L.; Liu, H.; Huang, L.; Yang, Q.; Turner, J.R.; Yu, Q. Lactobacillus accelerates ISCs regeneration to protect the integrity of intestinal mucosa through activation of STAT3 signaling pathway induced by LPLs secretion of IL-22. Cell Death Differ. 2018, 25, 1657–1670. [Google Scholar] [CrossRef]
  145. Walter, K.; Grosskopf, H.; Karkossa, I.; von Bergen, M.; Schubert, K. Proteomic Characterization of the Cellular Effects of AhR Activation by Microbial Tryptophan Catabolites in Endotoxin-Activated Human Macrophages. Int. J. Environ. Res. Public Health 2021, 18, 10336. [Google Scholar] [CrossRef]
  146. Zhang, B.; Jiang, M.; Zhao, J.; Song, Y.; Du, W.; Shi, J. The Mechanism Underlying the Influence of Indole-3-Propionic Acid: A Relevance to Metabolic Disorders. Front. Endocrinol. 2022, 13, 841703. [Google Scholar] [CrossRef]
  147. Ding, Y.; Yanagi, K.; Yang, F.; Callaway, E.; Cheng, C.; Hensel, M.E.; Menon, R.; Alaniz, R.C.; Lee, K.; Jayaraman, A. Oral supplementation of gut microbial metabolite indole-3-acetate alleviates diet-induced steatosis and inflammation in mice. Elife 2024, 12, RP87458. [Google Scholar] [CrossRef]
  148. Sehgal, R.; Ilha, M.; Vaittinen, M.; Kaminska, D.; Männistö, V.; Kärjä, V.; Tuomainen, M.; Hanhineva, K.; Romeo, S.; Pajukanta, P.; et al. Indole-3-Propionic Acid, a Gut-Derived Tryptophan Metabolite, Associates with Hepatic Fibrosis. Nutrients 2021, 13, 3509. [Google Scholar] [CrossRef]
  149. Wang, Y.; Wang, G.; Bai, J.; Zhao, N.; Wang, Q.; Zhou, R.; Li, G.; Hu, C.; Li, X.; Tao, K.; et al. Role of Indole-3-Acetic Acid in NAFLD Amelioration After Sleeve Gastrectomy. Obes. Surg. 2021, 31, 3040–3052. [Google Scholar] [CrossRef] [PubMed]
  150. Kelly, G.S. Conjugated linoleic acid: A review. Altern. Med. Rev. 2001, 6, 367–382. [Google Scholar] [PubMed]
  151. Elnar, A.G.; Jang, Y.; Kim, G.-B. Heterologous Expression and Polyphasic Analysis of CLA-Converting Linoleic Acid Isomerase from Bifidobacterium breve JKL2022. J. Agric. Food Chem. 2025, 73, 1425–1440. [Google Scholar] [CrossRef] [PubMed]
  152. Devillard, E.; McIntosh, F.M.; Duncan, S.H.; Wallace, R.J. Metabolism of Linoleic Acid by Human Gut Bacteria: Different Routes for Biosynthesis of Conjugated Linoleic Acid. J. Bacteriol. 2007, 189, 2566–2570. [Google Scholar] [CrossRef]
  153. Zeng, Y.; Liu, P.; Yang, X.; Li, H.; Li, H.; Guo, Y.; Meng, X.; Liu, X. The dietary c9,t11-conjugated linoleic acid enriched from butter reduces breast cancer progression in vivo. J. Food Biochem. 2020, 44, e13163. [Google Scholar] [CrossRef]
  154. Lehnen, T.E.; da Silva, M.R.; Camacho, A.; Marcadenti, A.; Lehnen, A.M. A review on effects of conjugated linoleic fatty acid (CLA) upon body composition and energetic metabolism. J. Int. Soc. Sports Nutr. 2015, 12, 36. [Google Scholar] [CrossRef]
  155. Toomey, S.; Harhen, B.; Roche, H.M.; Fitzgerald, D.; Belton, O. Profound resolution of early atherosclerosis with conjugated linoleic acid. Atherosclerosis 2006, 187, 40–49. [Google Scholar] [CrossRef]
  156. Viladomiu, M.; Hontecillas, R.; Bassaganya-Riera, J. Modulation of inflammation and immunity by dietary conjugated linoleic acid. Eur. J. Pharmacol. 2016, 785, 87–95. [Google Scholar] [CrossRef]
  157. Park, Y.; Kim, J.; Scrimgeour, A.G.; Condlin, M.L.; Kim, D.; Park, Y. Conjugated linoleic acid and calcium co-supplementation improves bone health in ovariectomised mice. Food Chem. 2013, 140, 280–288. [Google Scholar] [CrossRef]
  158. Nasrollahzadeh, A.; Mollaei Tavani, S.; Arjeh, E.; Jafari, S.M. Production of conjugated linoleic acid by lactic acid bacteria; important factors and optimum conditions. Food Chem. X 2023, 20, 100942. [Google Scholar] [CrossRef]
  159. Bae, D.-H.; Lane, D.J.R.; Jansson, P.J.; Richardson, D.R. The old and new biochemistry of polyamines. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 2053–2068. [Google Scholar] [CrossRef] [PubMed]
  160. Xuan, M.; Gu, X.; Li, J.; Huang, D.; Xue, C.; He, Y. Polyamines: Their significance for maintaining health and contributing to diseases. Cell Commun. Signal. 2023, 21, 348. [Google Scholar] [CrossRef] [PubMed]
  161. Sagar, N.A.; Tarafdar, S.; Agarwal, S.; Tarafdar, A.; Sharma, S. Polyamines: Functions, Metabolism, and Role in Human Disease Management. Med. Sci. 2021, 9, 44. [Google Scholar] [CrossRef]
  162. Hoyles, L.; Wallace, R.J. Gastrointestinal Tract: Intestinal Fatty Acid Metabolism and Implications for Health. In Handbook of Hydrocarbon and Lipid Microbiology; Timmis, K.N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 3119–3132. [Google Scholar] [CrossRef]
  163. Bui, T.I.; Britt, E.A.; Muthukrishnan, G.; Gill, S.R. Probiotic induced synthesis of microbiota polyamine as a nutraceutical for metabolic syndrome and obesity-related type 2 diabetes. Front. Endocrinol. 2022, 13, 1094258. [Google Scholar] [CrossRef]
  164. Pugin, B.; Barcik, W.; Westermann, P.; Heider, A.; Wawrzyniak, M.; Hellings, P.; Akdis, C.A.; O’Mahony, L. A wide diversity of bacteria from the human gut produces and degrades biogenic amines. Microb. Ecol. Health Dis. 2017, 28, 1353881. [Google Scholar] [CrossRef] [PubMed]
  165. Rao, J.N.; Xiao, L.; Wang, J.-Y. Polyamines in Gut Epithelial Renewal and Barrier Function. Physiology 2020, 35, 328–337. [Google Scholar] [CrossRef]
  166. Puleston, D.J.; Baixauli, F.; Sanin, D.E.; Edwards-Hicks, J.; Villa, M.; Kabat, A.M.; Kamiński, M.M.; Stanckzak, M.; Weiss, H.J.; Grzes, K.M.; et al. Polyamine metabolism is a central determinant of helper T cell lineage fidelity. Cell 2021, 184, 4186–4202.e20. [Google Scholar] [CrossRef]
  167. Wang, Z. Polyamines instruct T-cell differentiation. Nat. Cell Biol. 2021, 23, 811. [Google Scholar] [CrossRef]
  168. Sugahara, T.; Nishimoto, S.; Miyazaki, Y. Effects of polyamines on proliferation and IgM productivity of human-human hybridoma, HB4C5 cells. Cytotechnology 2008, 57, 115–122. [Google Scholar] [CrossRef]
  169. Metur, S.P.; Klionsky, D.J. The curious case of polyamines: Spermidine drives reversal of B cell senescence. Autophagy 2020, 16, 389–390. [Google Scholar] [CrossRef]
  170. Nowotarski, S.L.; Woster, P.M.; Casero, R.A.J. Polyamines and cancer: Implications for chemotherapy and chemoprevention. Expert Rev. Mol. Med. 2013, 15, e3. [Google Scholar] [CrossRef] [PubMed]
  171. Wu, J.-Y.; Zeng, Y.; You, Y.-Y.; Chen, Q.-Y. Polyamine metabolism and anti-tumor immunity. Front. Immunol. 2025, 16, 1529337. Available online: https://www.frontiersin.org/journals/immunology/articles/10.3389/fimmu.2025.1529337 (accessed on 9 April 2025). [CrossRef]
  172. Pandey, K.B.; Rizvi, S.I. Plant polyphenols as dietary antioxidants in human health and disease. Oxid. Med. Cell. Longev. 2009, 2, 270–278. [Google Scholar] [CrossRef]
  173. Williamson, G. The role of polyphenols in modern nutrition. Nutr. Bull. 2017, 42, 226–235. [Google Scholar] [CrossRef] [PubMed]
  174. Pasinetti, G.M.; Singh, R.; Westfall, S.; Herman, F.; Faith, J.; Ho, L. The Role of the Gut Microbiota in the Metabolism of Polyphenols as Characterized by Gnotobiotic Mice. J. Alzheimer’s Dis. 2018, 63, 409–421. [Google Scholar] [CrossRef]
  175. García-Villalba, R.; Giménez-Bastida, J.A.; Cortés-Martín, A.; Ávila-Gálvez, M.Á.; Tomás-Barberán, F.A.; Selma, M.V.; Espín, J.C.; González-Sarrías, A. Urolithins: A Comprehensive Update on their Metabolism, Bioactivity, and Associated Gut Microbiota. Mol. Nutr. Food Res. 2022, 66, e2101019. [Google Scholar] [CrossRef]
  176. Rahman, M.M.; Rahaman, M.S.; Islam, M.R.; Rahman, F.; Mithi, F.M.; Alqahtani, T.; Almikhlafi, M.A.; Alghamdi, S.Q.; Alruwaili, A.S.; Hossain, M.S.; et al. Role of Phenolic Compounds in Human Disease: Current Knowledge and Future Prospects. Molecules 2021, 27, 233. [Google Scholar] [CrossRef]
  177. Ji, Y.; Yin, Y.; Sun, L.; Zhang, W. The Molecular and Mechanistic Insights Based on Gut-Liver Axis: Nutritional Target for Non-Alcoholic Fatty Liver Disease (NAFLD) Improvement. Int. J. Mol. Sci. 2020, 21, 3066. [Google Scholar] [CrossRef] [PubMed]
  178. Nelson, A.L.; Porter, L. Hydrogen Peroxide Toxicity. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  179. Abdelshafy, A.M.; Neetoo, H.; Al-Asmari, F. Antimicrobial Activity of Hydrogen Peroxide for Application in Food Safety and COVID-19 Mitigation: An Updated Review. J. Food Prot. 2024, 87, 100306. [Google Scholar] [CrossRef]
  180. Murphy, E.C.; Friedman, A.J. Hydrogen peroxide and cutaneous biology: Translational applications, benefits, and risks. J. Am. Acad. Dermatol. 2019, 81, 1379–1386. [Google Scholar] [CrossRef]
  181. Knaus, U.G.; Hertzberger, R.; Pircalabioru, G.G.; Yousefi, S.P.M.; Branco Dos Santos, F. Pathogen control at the intestinal mucosa—H2O2 to the rescue. Gut Microbes 2017, 8, 67–74. [Google Scholar] [CrossRef] [PubMed]
  182. Driessens, N.; Versteyhe, S.; Ghaddhab, C.; Burniat, A.; De Deken, X.; Van Sande, J.; Dumont, J.-E.; Miot, F.; Corvilain, B. Hydrogen peroxide induces DNA single- and double-strand breaks in thyroid cells and is therefore a potential mutagen for this organ. Endocr. Relat. Cancer 2009, 16, 845–856. [Google Scholar] [CrossRef] [PubMed]
  183. Whittemore, E.R.; Loo, D.T.; Watt, J.A.; Cotman, C.W. A detailed analysis of hydrogen peroxide-induced cell death in primary neuronal culture. Neuroscience 1995, 67, 921–932. [Google Scholar] [CrossRef]
  184. Varadaraj, K.; Gao, J.; Mathias, R.T.; Kumari, S. Effect of hydrogen peroxide on lens transparency, intracellular pH, gap junction coupling, hydrostatic pressure and membrane water permeability. Exp. Eye Res. 2024, 245, 109957. [Google Scholar] [CrossRef] [PubMed]
  185. Ramsay, J.; Morton, J.; Norris, M.; Kanungo, S. Organic acid disorders. Ann. Transl. Med. 2018, 6, 472. [Google Scholar] [CrossRef]
  186. Nuttall, K.L.; Guzman, N.A. Organic Acids. In Clinical and Forensic Applications of Capillary Electrophoresis; Petersen, J.R., Mohammad, A.A., Eds.; Humana Press: Totowa, NJ, USA, 2001; pp. 193–208. [Google Scholar] [CrossRef]
  187. Neal-McKinney, J.M.; Lu, X.; Duong, T.; Larson, C.L.; Call, D.R.; Shah, D.H.; Konkel, M.E. Production of organic acids by probiotic lactobacilli can be used to reduce pathogen load in poultry. PLoS ONE 2012, 7, e43928. [Google Scholar] [CrossRef]
  188. Kareem, K.Y.; Hooi Ling, F.; Teck Chwen, L.; May Foong, O.; Anjas Asmara, S. Inhibitory activity of postbiotic produced by strains of Lactobacillus plantarum using reconstituted media supplemented with inulin. Gut Pathog. 2014, 6, 23. [Google Scholar] [CrossRef]
  189. Rad, A.H.; Aghebati-Maleki, L.; Kafil, H.S.; Gilani, N.; Abbasi, A.; Khani, N. Postbiotics, as dynamic biomolecules, and their promising role in promoting food safety. Biointerface Res. Appl. Chem. 2021, 11, 14529–14544. [Google Scholar]
  190. Forman, H.J.; Zhang, H.; Rinna, A. Glutathione: Overview of its protective roles, measurement, and biosynthesis. Mol. Asp. Med. 2009, 30, 1–12. [Google Scholar] [CrossRef]
  191. Alanazi, A.M.; Mostafa, G.A.E.; Al-Badr, A.A. Glutathione. Profiles Drug Subst. Excip. Relat. Methodol. 2015, 40, 43–158. [Google Scholar] [CrossRef]
  192. Yuan, Y.; Yang, J.; Zhuge, A.; Li, L.; Ni, S. Gut microbiota modulates osteoclast glutathione synthesis and mitochondrial biogenesis in mice subjected to ovariectomy. Cell Prolif. 2022, 55, e13194. [Google Scholar] [CrossRef] [PubMed]
  193. Silvagno, F.; Vernone, A.; Pescarmona, G.P. The Role of Glutathione in Protecting against the Severe Inflammatory Response Triggered by COVID-19. Antioxidants 2020, 9, 624. [Google Scholar] [CrossRef] [PubMed]
  194. Limongi, D.; Baldelli, S.; Checconi, P.; Marcocci, M.E.; De Chiara, G.; Fraternale, A.; Magnani, M.; Ciriolo, M.R.; Palamara, A.T. GSH-C4 Acts as Anti-inflammatory Drug in Different Models of Canonical and Cell Autonomous Inflammation Through NFκB Inhibition. Front. Immunol. 2019, 10, 155. [Google Scholar] [CrossRef]
  195. Fraternale, A.; Paoletti, M.F.; Casabianca, A.; Oiry, J.; Clayette, P.; Vogel, J.-U.; Cinatl, J.J.; Palamara, A.T.; Sgarbanti, R.; Garaci, E.; et al. Antiviral and immunomodulatory properties of new pro-glutathione (GSH) molecules. Curr. Med. Chem. 2006, 13, 1749–1755. [Google Scholar] [CrossRef]
  196. Lewis, T.; Stone, W.L. Biochemistry, Proteins Enzymes. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  197. Gurung, N.; Ray, S.; Bose, S.; Rai, V. A broader view: Microbial enzymes and their relevance in industries, medicine, and beyond. BioMed Res. Int. 2013, 2013, 329121. [Google Scholar] [CrossRef]
  198. Anbu, P.; Gopinath, S.C.B.; Chaulagain, B.P.; Lakshmipriya, T. Microbial Enzymes and Their Applications in Industries and Medicine 2016. BioMed Res. Int. 2017, 2017, 2195808. [Google Scholar] [CrossRef]
  199. Chalamaiah, M.; Yu, W.; Wu, J. Immunomodulatory and anticancer protein hydrolysates (peptides) from food proteins: A review. Food Chem. 2018, 245, 205–222. [Google Scholar] [CrossRef] [PubMed]
  200. Osman, A.; El-Gazzar, N.; Almanaa, T.N.; El-Hadary, A.; Sitohy, M. Lipolytic Postbiotic from Lactobacillus paracasei Manages Metabolic Syndrome in Albino Wistar Rats. Molecules 2021, 26, 472. [Google Scholar] [CrossRef]
  201. Lai, H.-H.; Yeh, K.-Y.; Hsu, H.-M.; Her, G.M. Deficiency of Adipose Triglyceride Lipase Induces Metabolic Syndrome and Cardiomyopathy in Zebrafish. Int. J. Mol. Sci. 2023, 24, 117. [Google Scholar] [CrossRef]
  202. Hasham, S.N.; Pillarisetti, S. Vascular lipases, inflammation and atherosclerosis. Clin. Chim. Acta 2006, 372, 179–183. [Google Scholar] [CrossRef]
  203. Fiorucci, S.; Marchianò, S.; Urbani, G.; Di Giorgio, C.; Distrutti, E.; Zampella, A.; Biagioli, M. Immunology of bile acids regulated receptors. Prog. Lipid Res. 2024, 95, 101291. [Google Scholar] [CrossRef] [PubMed]
  204. Ridlon, J.M.; Kang, D.J.; Hylemon, P.B.; Bajaj, J.S. Bile acids and the gut microbiome. Curr. Opin. Gastroenterol. 2014, 30, 332–338. [Google Scholar] [CrossRef] [PubMed]
  205. Chiang, J.Y.L.; Ferrell, J.M. Up to date on cholesterol 7 alpha-hydroxylase (CYP7A1) in bile acid synthesis. Liver Res. 2020, 4, 47–63. [Google Scholar] [CrossRef]
  206. Franzosa, E.A.; Sirota-Madi, A.; Avila-Pacheco, J.; Fornelos, N.; Haiser, H.J.; Reinker, S.; Vatanen, T.; Hall, A.B.; Mallick, H.; McIver, L.J.; et al. Gut microbiome structure and metabolic activity in inflammatory bowel disease. Nat. Microbiol. 2019, 4, 293–305. [Google Scholar] [CrossRef]
  207. Agolino, G.; Pino, A.; Vaccalluzzo, A.; Cristofolini, M.; Solieri, L.; Caggia, C.; Randazzo, C.L. Bile salt hydrolase: The complexity behind its mechanism in relation to lowering-cholesterol lactobacilli probiotics. J. Funct. Foods 2024, 120, 106357. [Google Scholar] [CrossRef]
  208. Sato, Y.; Atarashi, K.; Plichta, D.R.; Arai, Y.; Sasajima, S.; Kearney, S.M.; Suda, W.; Takeshita, K.; Sasaki, T.; Okamoto, S.; et al. Novel bile acid biosynthetic pathways are enriched in the microbiome of centenarians. Nature 2021, 599, 458–464. [Google Scholar] [CrossRef] [PubMed]
  209. Mohanty, I.; Mannochio-Russo, H.; Schweer, J.V.; El Abiead, Y.; Bittremieux, W.; Xing, S.; Schmid, R.; Zuffa, S.; Vasquez, F.; Muti, V.B.; et al. The underappreciated diversity of bile acid modifications. Cell 2024, 187, 1801–1818.e20. [Google Scholar] [CrossRef]
  210. Cai, J.; Sun, L.; Gonzalez, F.J. Gut microbiota-derived bile acids in intestinal immunity, inflammation, and tumorigenesis. Cell Host Microbe 2022, 30, 289–300. [Google Scholar] [CrossRef] [PubMed]
  211. Guzior, D.V.; Okros, M.; Shivel, M.; Armwald, B.; Bridges, C.; Fu, Y.; Martin, C.; Schilmiller, A.L.; Miller, W.M.; Ziegler, K.M.; et al. Bile salt hydrolase acyltransferase activity expands bile acid diversity. Nature 2024, 626, 852–858. [Google Scholar] [CrossRef]
  212. Guzior, D.V.; Quinn, R.A. Review: Microbial transformations of human bile acids. Microbiome 2021, 9, 140. [Google Scholar] [CrossRef]
  213. Fiorucci, S.; Distrutti, E.; Carino, A.; Zampella, A.; Biagioli, M. Bile acids and their receptors in metabolic disorders. Prog. Lipid Res. 2021, 82, 101094. [Google Scholar] [CrossRef] [PubMed]
  214. Biagioli, M.; Fiorucci, S. Bile acid activated receptors: Integrating immune and metabolic regulation in non-alcoholic fatty liver disease. Liver Res. 2021, 5, 119–141. [Google Scholar] [CrossRef]
  215. Quinn, R.A.; Melnik, A.V.; Vrbanac, A.; Fu, T.; Patras, K.A.; Christy, M.P.; Bodai, Z.; Belda-Ferre, P.; Tripathi, A.; Chung, L.K.; et al. Global chemical effects of the microbiome include new bile-acid conjugations. Nature 2020, 579, 123–129. [Google Scholar] [CrossRef] [PubMed]
  216. Ridlon, J.M.; Gaskins, H.R. Another renaissance for bile acid gastrointestinal microbiology. Nat. Rev. Gastroenterol. Hepatol. 2024, 21, 348–364. [Google Scholar] [CrossRef]
  217. Mohanty, I.; Allaband, C.; Mannochio-Russo, H.; El Abiead, Y.; Hagey, L.R.; Knight, R.; Dorrestein, P.C. The changing metabolic landscape of bile acids—Keys to metabolism and immune regulation. Nat. Rev. Gastroenterol. Hepatol. 2024, 21, 493–516. [Google Scholar] [CrossRef]
  218. Parks, D.J.; Blanchard, S.G.; Bledsoe, R.K.; Chandra, G.; Consler, T.G.; Kliewer, S.A.; Stimmel, J.B.; Willson, T.M.; Zavacki, A.M.; Moore, D.D.; et al. Bile acids: Natural ligands for an orphan nuclear receptor. Science 1999, 284, 1365–1368. [Google Scholar] [CrossRef]
  219. Maruyama, T.; Miyamoto, Y.; Nakamura, T.; Tamai, Y.; Okada, H.; Sugiyama, E.; Nakamura, T.; Itadani, H.; Tanaka, K. Identification of membrane-type receptor for bile acids (M-BAR). Biochem. Biophys. Res. Commun. 2002, 298, 714–719. [Google Scholar] [CrossRef] [PubMed]
  220. Fiorucci, S.; Urbani, G.; Di Giorgio, C.; Biagioli, M.; Distrutti, E. Bile Acids-Based Therapies for Primary Sclerosing Cholangitis: Current Landscape and Future Developments. Cells 2024, 13, 1650. [Google Scholar] [CrossRef]
  221. Wang, H.; Chen, J.; Hollister, K.; Sowers, L.C.; Forman, B.M. Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol. Cell 1999, 3, 543–553. [Google Scholar] [CrossRef]
  222. Watanabe, M.; Houten, S.M.; Mataki, C.; Christoffolete, M.A.; Kim, B.W.; Sato, H.; Messaddeq, N.; Harney, J.W.; Ezaki, O.; Kodama, T.; et al. Bile acids induce energy expenditure by promoting intracellular thyroid hormone activation. Nature 2006, 439, 484–489. [Google Scholar] [CrossRef]
  223. Biagioli, M.; Di Giorgio, C.; Massa, C.; Marchianò, S.; Bellini, R.; Bordoni, M.; Urbani, G.; Roselli, R.; Lachi, G.; Morretta, E.; et al. Microbial-derived bile acid reverses inflammation in IBD via GPBAR1 agonism and RORγt inverse agonism. Biomed. Pharmacother. 2024, 181, 117731. [Google Scholar] [CrossRef] [PubMed]
  224. Biagioli, M.; Marchianò, S.; Carino, A.; Di Giorgio, C.; Santucci, L.; Distrutti, E.; Fiorucci, S. Bile Acids Activated Receptors in Inflammatory Bowel Disease. Cells 2021, 10, 1281. [Google Scholar] [CrossRef]
  225. Marchianò, S.; Biagioli, M.; Bordoni, M.; Morretta, E.; Di Giorgio, C.; Vellecco, V.; Roselli, R.; Bellini, R.; Massa, C.; Cari, L.; et al. Defective Bile Acid Signaling Promotes Vascular Dysfunction, Supporting a Role for G-Protein Bile Acid Receptor 1/Farnesoid X Receptor Agonism and Statins in the Treatment of Nonalcoholic Fatty Liver Disease. J. Am. Heart Assoc. 2023, 12, e031241. [Google Scholar] [CrossRef]
  226. Biagioli, M.; Carino, A.; Cipriani, S.; Francisci, D.; Marchianò, S.; Scarpelli, P.; Sorcini, D.; Zampella, A.; Fiorucci, S. The Bile Acid Receptor GPBAR1 Regulates the M1/M2 Phenotype of Intestinal Macrophages and Activation of GPBAR1 Rescues Mice from Murine Colitis. J. Immunol. 2017, 199, 718–733. [Google Scholar] [CrossRef]
  227. Lewis, N.D.; Patnaude, L.A.; Pelletier, J.; Souza, D.J.; Lukas, S.M.; King, F.J.; Hill, J.D.; Stefanopoulos, D.E.; Ryan, K.; Desai, S.; et al. A GPBAR1 (TGR5) small molecule agonist shows specific inhibitory effects on myeloid cell activation in vitro and reduces experimental autoimmune encephalitis (EAE) in vivo. PLoS ONE 2014, 9, e100883. [Google Scholar] [CrossRef]
  228. Zhou, H.; Zhou, S.; Shi, Y.; Wang, Q.; Wei, S.; Wang, P.; Cheng, F.; Auwerx, J.; Schoonjans, K.; Lu, L. TGR5/Cathepsin E signaling regulates macrophage innate immune activation in liver ischemia and reperfusion injury. Am. J. Transplant. Off. J. Am. Soc. Transplant. Am. Soc. Transpl. Surg. 2021, 21, 1453–1464. [Google Scholar] [CrossRef] [PubMed]
  229. Högenauer, K.; Arista, L.; Schmiedeberg, N.; Werner, G.; Jaksche, H.; Bouhelal, R.; Nguyen, D.G.; Bhat, B.G.; Raad, L.; Rauld, C.; et al. G-protein-coupled bile acid receptor 1 (GPBAR1, TGR5) agonists reduce the production of proinflammatory cytokines and stabilize the alternative macrophage phenotype. J. Med. Chem. 2014, 57, 10343–10354. [Google Scholar] [CrossRef]
  230. Carino, A.; Marchianò, S.; Biagioli, M.; Bucci, M.; Vellecco, V.; Brancaleone, V.; Fiorucci, C.; Zampella, A.; Monti, M.C.C.; Distrutti, E.; et al. Agonism for the bile acid receptor GPBAR1 reverses liver and vascular damage in a mouse model of steatohepatitis. FASEB J. 2019, 33, 2809–2822. [Google Scholar] [CrossRef] [PubMed]
  231. Biagioli, M.; Carino, A.; Fiorucci, C.; Marchianò, S.; Di Giorgio, C.; Bordoni, M.; Roselli, R.; Baldoni, M.; Distrutti, E.; Zampella, A.; et al. The Bile Acid Receptor GPBAR1 Modulates CCL2/CCR2 Signaling at the Liver Sinusoidal/Macrophage Interface and Reverses Acetaminophen-Induced Liver Toxicity. J. Immunol. 2020, 204, 2535–2551. [Google Scholar] [CrossRef]
  232. Guo, C.; Xie, S.; Chi, Z.; Zhang, J.; Liu, Y.; Zhang, L.; Zheng, M.; Zhang, X.; Xia, D.; Ke, Y.; et al. Bile Acids Control Inflammation and Metabolic Disorder through Inhibition of NLRP3 Inflammasome. Immunity 2016, 45, 944. [Google Scholar] [CrossRef]
  233. Keitel, V.; Donner, M.; Winandy, S.; Kubitz, R.; Häussinger, D. Expression and function of the bile acid receptor TGR5 in Kupffer cells. Biochem. Biophys. Res. Commun. 2008, 372, 78–84. [Google Scholar] [CrossRef] [PubMed]
  234. Hu, J.; Wang, C.; Huang, X.; Yi, S.; Pan, S.; Zhang, Y.; Yuan, G.; Cao, Q.; Ye, X.; Li, H. Gut microbiota-mediated secondary bile acids regulate dendritic cells to attenuate autoimmune uveitis through TGR5 signaling. Cell Rep. 2021, 36, 109726. [Google Scholar] [CrossRef] [PubMed]
  235. Hu, J.; Zhang, Y.; Yi, S.; Wang, C.; Huang, X.; Pan, S.; Yang, J.; Yuan, G.; Tan, S.; Li, H. Lithocholic acid inhibits dendritic cell activation by reducing intracellular glutathione via TGR5 signaling. Int. J. Biol. Sci. 2022, 18, 4545–4559. [Google Scholar] [CrossRef]
  236. Biagioli, M.; Carino, A.; Fiorucci, C.; Marchianò, S.; Di Giorgio, C.; Roselli, R.; Magro, M.; Distrutti, E.; Bereshchenko, O.; Scarpelli, P.; et al. GPBAR1 Functions as Gatekeeper for Liver NKT Cells and provides Counterregulatory Signals in Mouse Models of Immune-Mediated Hepatitis. Cell Mol. Gastroenterol. Hepatol. 2019, 8, 447–473. [Google Scholar] [CrossRef]
  237. Vavassori, P.; Mencarelli, A.; Renga, B.; Distrutti, E.; Fiorucci, S. The bile acid receptor FXR is a modulator of intestinal innate immunity. J. Immunol. 2009, 183, 6251–6261. [Google Scholar] [CrossRef]
  238. Renga, B.; D’Amore, C.; Cipriani, S.; Mencarelli, A.; Carino, A.; Sepe, V.; Zampella, A.; Distrutti, E.; Fiorucci, S. FXR mediates a chromatin looping in the GR promoter thus promoting the resolution of colitis in rodents. Pharmacol. Res. 2013, 77, 1–10. [Google Scholar] [CrossRef]
  239. Liu, H.; Pathak, P.; Boehme, S.; Chiang, J.Y. Cholesterol 7α-hydroxylase protects the liver from inflammation and fibrosis by maintaining cholesterol homeostasis. J. Lipid Res. 2016, 57, 1831–1844. [Google Scholar] [CrossRef] [PubMed]
  240. Hao, H.; Cao, L.; Jiang, C.; Che, Y.; Zhang, S.; Takahashi, S.; Wang, G.; Gonzalez, F.J. Farnesoid X Receptor Regulation of the NLRP3 Inflammasome Underlies Cholestasis-Associated Sepsis. Cell Metab. 2017, 25, 856–867.e5. [Google Scholar] [CrossRef]
  241. Jaroonwitchawan, T.; Arimochi, H.; Sasaki, Y.; Ishifune, C.; Kondo, H.; Otsuka, K.; Tsukumo, S.-I.; Yasutomo, K. Stimulation of the farnesoid X receptor promotes M2 macrophage polarization. Front. Immunol. 2023, 14, 1065790. [Google Scholar] [CrossRef]
  242. Jin, D.; Lu, T.; Ni, M.; Wang, H.; Zhang, J.; Zhong, C.; Shen, C.; Hao, J.; Busuttil, R.W.; Kupiec-Weglinski, J.W.; et al. Farnesoid X Receptor Activation Protects Liver from Ischemia/Reperfusion Injury by Up-Regulating Small Heterodimer Partner in Kupffer Cells. Hepatol. Commun. 2020, 4, 540–554. [Google Scholar] [CrossRef]
  243. Verbeke, L.; Mannaerts, I.; Schierwagen, R.; Govaere, O.; Klein, S.; Vander Elst, I.; Windmolders, P.; Farre, R.; Wenes, M.; Mazzone, M.; et al. FXR agonist obeticholic acid reduces hepatic inflammation and fibrosis in a rat model of toxic cirrhosis. Sci. Rep. 2016, 6, 33453. [Google Scholar] [CrossRef] [PubMed]
  244. Massafra, V.; Ijssennagger, N.; Plantinga, M.; Milona, A.; Ramos Pittol, J.M.; Boes, M.; van Mil, S.W. Splenic dendritic cell involvement in FXR-mediated amelioration of DSS colitis. Biochim. Biophys. Acta 2016, 1862, 166–173. [Google Scholar] [CrossRef] [PubMed]
  245. Fu, T.; Li, Y.; Oh, T.G.; Cayabyab, F.; He, N.; Tang, Q.; Coulter, S.; Truitt, M.; Medina, P.; He, M.; et al. FXR mediates ILC-intrinsic responses to intestinal inflammation. Proc. Natl. Acad. Sci. USA 2022, 119, e2213041119. [Google Scholar] [CrossRef]
  246. Mencarelli, A.; Renga, B.; Migliorati, M.; Cipriani, S.; Distrutti, E.; Santucci, L.; Fiorucci, S. The bile acid sensor farnesoid X receptor is a modulator of liver immunity in a rodent model of acute hepatitis. J. Immunol. 2009, 183, 6657–6666. [Google Scholar] [CrossRef] [PubMed]
  247. Campbell, C.; Marchildon, F.; Michaels, A.J.; Takemoto, N.; van der Veeken, J.; Schizas, M.; Pritykin, Y.; Leslie, C.S.; Intlekofer, A.M.; Cohen, P.; et al. FXR mediates T cell-intrinsic responses to reduced feeding during infection. Proc. Natl. Acad. Sci. USA 2020, 117, 33446–33454. [Google Scholar] [CrossRef]
  248. Qiu, Z.; Cervantes, J.L.; Cicek, B.B.; Mukherjee, S.; Venkatesh, M.; Maher, L.A.; Salazar, J.C.; Mani, S.; Khanna, K.M. Pregnane X Receptor Regulates Pathogen-Induced Inflammation and Host Defense Against an Intracellular Bacterial Infection through Toll-like Receptor 4. Sci. Rep. 2016, 6, 31936. [Google Scholar] [CrossRef]
  249. Hudson, G.; Flannigan, K.L.; Venu, V.K.P.; Alston, L.; Sandall, C.F.; MacDonald, J.A.; Muruve, D.A.; Chang, T.K.H.; Mani, S.; Hirota, S.A. Pregnane X Receptor Activation Triggers Rapid ATP Release in Primed Macrophages That Mediates NLRP3 Inflammasome Activation. J. Pharmacol. Exp. Ther. 2019, 370, 44–53. [Google Scholar] [CrossRef]
  250. Dubrac, S.; Elentner, A.; Ebner, S.; Horejs-Hoeck, J.; Schmuth, M. Modulation of T lymphocyte function by the pregnane X receptor. J. Immunol. 2010, 184, 2949–2957. [Google Scholar] [CrossRef]
  251. Casey, S.C.; Blumberg, B. The steroid and xenobiotic receptor negatively regulates B-1 cell development in the fetal liver. Mol. Endocrinol. 2012, 26, 916–925. [Google Scholar] [CrossRef]
  252. Zhang, Y.; Leung, D.Y.M.; Richers, B.N.; Liu, Y.; Remigio, L.K.; Riches, D.W.; Goleva, E. Vitamin D inhibits monocyte/macrophage proinflammatory cytokine production by targeting MAPK phosphatase-1. J. Immunol. 2012, 188, 2127–2135. [Google Scholar] [CrossRef]
  253. Dong, B.; Zhou, Y.; Wang, W.; Scott, J.; Kim, K.; Sun, Z.; Guo, Q.; Lu, Y.; Gonzales, N.M.; Wu, H.; et al. Vitamin D Receptor Activation in Liver Macrophages Ameliorates Hepatic Inflammation, Steatosis, and Insulin Resistance in Mice. Hepatology 2020, 71, 1559–1574. [Google Scholar] [CrossRef] [PubMed]
  254. Zhou, Y.; Dong, B.; Kim, K.H.; Choi, S.; Sun, Z.; Wu, N.; Wu, Y.; Scott, J.; Moore, D.D. Vitamin D Receptor Activation in Liver Macrophages Protects Against Hepatic Endoplasmic Reticulum Stress in Mice. Hepatology 2020, 71, 1453–1466. [Google Scholar] [CrossRef] [PubMed]
  255. Adorini, L.; Penna, G.; Giarratana, N.; Roncari, A.; Amuchastegui, S.; Daniel, K.C.; Uskokovic, M. Dendritic cells as key targets for immunomodulation by Vitamin D receptor ligands. J. Steroid Biochem. Mol. Biol. 2004, 89–90, 437–441. [Google Scholar] [CrossRef]
  256. Català-Moll, F.; Ferreté-Bonastre, A.G.; Godoy-Tena, G.; Morante-Palacios, O.; Ciudad, L.; Barberà, L.; Fondelli, F.; Martínez-Cáceres, E.M.; Rodríguez-Ubreva, J.; Li, T.; et al. Vitamin D receptor, STAT3, and TET2 cooperate to establish tolerogenesis. Cell Rep. 2022, 38, 110244. [Google Scholar] [CrossRef]
  257. Griffin, M.D.; Lutz, W.; Phan, V.A.; Bachman, L.A.; McKean, D.J.; Kumar, R. Dendritic cell modulation by 1alpha,25 dihydroxyvitamin D3 and its analogs: A vitamin D receptor-dependent pathway that promotes a persistent state of immaturity in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2001, 98, 6800–6805. [Google Scholar] [CrossRef]
  258. Liu, Z.-Q.; Li, M.-G.; Geng, X.-R.; Liu, J.; Yang, G.; Qiu, S.-Q.; Liu, Z.-G.; Yang, P.-C. Vitamin D regulates immunoglobulin mucin domain molecule-4 expression in dendritic cells. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 2017, 47, 656–664. [Google Scholar] [CrossRef] [PubMed]
  259. Yu, S.; Cantorna, M.T. The vitamin D receptor is required for iNKT cell development. Proc. Natl. Acad. Sci. USA 2008, 105, 5207–5212. [Google Scholar] [CrossRef]
  260. Chen, J.; Bruce, D.; Cantorna, M.T. Vitamin D receptor expression controls proliferation of naïve CD8+ T cells and development of CD8 mediated gastrointestinal inflammation. BMC Immunol. 2014, 15, 6. [Google Scholar] [CrossRef]
  261. Cantorna, M.T.; Waddell, A. The vitamin D receptor turns off chronically activated T cells. Ann. N. Y. Acad. Sci. 2014, 1317, 70–75. [Google Scholar] [CrossRef]
  262. Lu, D.; Lan, B.; Din, Z.; Chen, H.; Chen, G. A vitamin D receptor agonist converts CD4+ T cells to Foxp3+ regulatory T cells in patients with ulcerative colitis. Oncotarget 2017, 8, 53552–53562. [Google Scholar] [CrossRef]
  263. Cording, S.; Medvedovic, J.; Cherrier, M.; Eberl, G. Development and regulation of RORγt+ innate lymphoid cells. FEBS Lett. 2014, 588, 4176–4181. [Google Scholar] [CrossRef] [PubMed]
  264. Withers, D.R.; Hepworth, M.R.; Wang, X.; Mackley, E.C.; Halford, E.E.; Dutton, E.E.; Marriott, C.L.; Brucklacher-Waldert, V.; Veldhoen, M.; Kelsen, J.; et al. Transient inhibition of ROR-γt therapeutically limits intestinal inflammation by reducing TH17 cells and preserving group 3 innate lymphoid cells. Nat. Med. 2016, 22, 319–323. [Google Scholar] [CrossRef] [PubMed]
  265. Ivanov, I.I.; McKenzie, B.S.; Zhou, L.; Tadokoro, C.E.; Lepelley, A.; Lafaille, J.J.; Cua, D.J.; Littman, D.R. The orphan nuclear receptor RORgammat directs the differentiation program of proinflammatory IL-17+ T helper cells. Cell 2006, 126, 1121–1133. [Google Scholar] [CrossRef] [PubMed]
  266. Sun, M.; He, C.; Chen, L.; Yang, W.; Wu, W.; Chen, F.; Cao, A.T.; Yao, S.; Dann, S.M.; Dhar, T.G.M.; et al. RORγt Represses IL-10 Production in Th17 Cells To Maintain Their Pathogenicity in Inducing Intestinal Inflammation. J. Immunol. 2019, 202, 79–92. [Google Scholar] [CrossRef]
  267. Kawamata, Y.; Fujii, R.; Hosoya, M.; Harada, M.; Yoshida, H.; Miwa, M.; Fukusumi, S.; Habata, Y.; Itoh, T.; Shintani, Y.; et al. A G protein-coupled receptor responsive to bile acids. J. Biol. Chem. 2003, 278, 9435–9440. [Google Scholar] [CrossRef]
  268. Oeckinghaus, A.; Hayden, M.S.; Ghosh, S. Crosstalk in NF-κB signaling pathways. Nat. Immunol. 2011, 12, 695–708. [Google Scholar] [CrossRef]
  269. Wang, Y.D.; Chen, W.D.; Yu, D.; Forman, B.M.; Huang, W. The G-protein-coupled bile acid receptor, Gpbar1 (TGR5), negatively regulates hepatic inflammatory response through antagonizing nuclear factor κ light-chain enhancer of activated B cells (NF-κB) in mice. Hepatology 2011, 54, 1421–1432. [Google Scholar] [CrossRef]
  270. Lieu, T.; Jayaweera, G.; Bunnett, N.W. GPBA: A GPCR for bile acids and an emerging therapeutic target for disorders of digestion and sensation. Br. J. Pharmacol. 2014, 171, 1156–1166. [Google Scholar] [CrossRef]
  271. Shin, D.J.; Wang, L. Bile Acid-Activated Receptors: A Review on FXR and Other Nuclear Receptors. In Handbook of Experimental Pharmacology; Springer: Cham, Switzerland, 2019; Volume 256, pp. 51–72. [Google Scholar] [CrossRef]
  272. Sepe, V.; Ummarino, R.; D’Auria, M.V.; Mencarelli, A.; D’Amore, C.; Renga, B.; Zampella, A.; Fiorucci, S. Total synthesis and pharmacological characterization of solomonsterol A, a potent marine pregnane-X-receptor agonist endowed with anti-inflammatory activity. J. Med. Chem. 2011, 54, 4590–4599. [Google Scholar] [CrossRef]
  273. Chen, M.L.; Huang, X.; Wang, H.; Hegner, C.; Liu, Y.; Shang, J.; Eliason, A.; Diao, H.; Park, H.; Frey, B.; et al. CAR directs T cell adaptation to bile acids in the small intestine. Nature 2021, 593, 147–151. [Google Scholar] [CrossRef]
  274. Hang, S.; Paik, D.; Yao, L.; Kim, E.; Trinath, J.; Lu, J.; Ha, S.; Nelson, B.N.; Kelly, S.P.; Wu, L.; et al. Bile acid metabolites control TH17 and Treg cell differentiation. Nature 2019, 576, 143–148. [Google Scholar] [CrossRef]
  275. Cook, D.N.; Kang, H.S.; Jetten, A.M. Retinoic Acid-Related Orphan Receptors (RORs): Regulatory Functions in Immunity, Development, Circadian Rhythm, and Metabolism. Nucl. Recept. Res. 2015, 2, 101185. [Google Scholar] [CrossRef]
  276. Tanoue, T.; Atarashi, K.; Honda, K. Development and maintenance of intestinal regulatory T cells. Nat. Rev. Immunol. 2016, 16, 295–309. [Google Scholar] [CrossRef] [PubMed]
  277. Campbell, C.; McKenney, P.T.; Konstantinovsky, D.; Isaeva, O.I.; Schizas, M.; Verter, J.; Mai, C.; Jin, W.-B.; Guo, C.-J.; Violante, S.; et al. Bacterial metabolism of bile acids promotes generation of peripheral regulatory T cells. Nature 2020, 581, 475–479. [Google Scholar] [CrossRef] [PubMed]
  278. Inagaki, T.; Moschetta, A.; Lee, Y.K.; Peng, L.; Zhao, G.; Downes, M.; Yu, R.T.; Shelton, J.M.; Richardson, J.A.; Repa, J.J.; et al. Regulation of antibacterial defense in the small intestine by the nuclear bile acid receptor. Proc. Natl. Acad. Sci. USA 2006, 103, 3920–3925. [Google Scholar] [CrossRef] [PubMed]
  279. Katsuma, S.; Hirasawa, A.; Tsujimoto, G. Bile acids promote glucagon-like peptide-1 secretion through TGR5 in a murine enteroendocrine cell line STC-1. Biochem. Biophys. Res. Commun. 2005, 329, 386–390. [Google Scholar] [CrossRef]
  280. Lee, J.M.; Wagner, M.; Xiao, R.; Kim, K.H.; Feng, D.; Lazar, M.A.; Moore, D.D. Nutrient-sensing nuclear receptors coordinate autophagy. Nature 2014, 516, 112–115. [Google Scholar] [CrossRef]
  281. Fang, S.; Suh, J.M.; Reilly, S.M.; Yu, E.; Osborn, O.; Lackey, D.; Yoshihara, E.; Perino, A.; Jacinto, S.; Lukasheva, Y.; et al. Intestinal FXR agonism promotes adipose tissue browning and reduces obesity and insulin resistance. Nat. Med. 2015, 21, 159–165. [Google Scholar] [CrossRef]
  282. Sepe, V.; Distrutti, E.; Fiorucci, S.; Zampella, A. Farnesoid X receptor modulators 2014-present: A patent review. Expert Opin. Ther. Pat. 2018, 28, 351–364. [Google Scholar] [CrossRef]
  283. Nogueiras, R.; Nauck, M.A.; Tschöp, M.H. Gut hormone co-agonists for the treatment of obesity: From bench to bedside. Nat. Metab. 2023, 5, 933–944. [Google Scholar] [CrossRef]
  284. Byun, S.; Kim, Y.C.; Zhang, Y.; Kong, B.; Guo, G.; Sadoshima, J.; Ma, J.; Kemper, B.; Kemper, J.K. A postprandial FGF19-SHP-LSD1 regulatory axis mediates epigenetic repression of hepatic autophagy. EMBO J. 2017, 36, 1755–1769. [Google Scholar] [CrossRef]
  285. Carino, A.; Marchianò, S.; Biagioli, M.; Scarpelli, P.; Bordoni, M.; Di Giorgio, C.; Roselli, R.; Fiorucci, C.; Monti, M.C.C.; Distrutti, E.; et al. The bile acid activated receptors GPBAR1 and FXR exert antagonistic effects on autophagy. FASEB J. 2021, 35, e21271. [Google Scholar] [CrossRef]
  286. Fiorucci, S.; Urbani, G.; Di Giorgio, C.; Biagioli, M.; Distrutti, E. Current Landscape and Evolving Therapies for Primary Biliary Cholangitis. Cells 2024, 13, 1580. [Google Scholar] [CrossRef]
  287. Festa, C.; Renga, B.; D’Amore, C.; Sepe, V.; Finamore, C.; De Marino, S.; Carino, A.; Cipriani, S.; Monti, M.C.C.; Zampella, A.; et al. Exploitation of cholane scaffold for the discovery of potent and selective farnesoid X receptor (FXR) and G-protein coupled bile acid receptor 1 (GP-BAR1) ligands. J. Med. Chem. 2014, 57, 8477–8495. [Google Scholar] [CrossRef] [PubMed]
  288. Sepe, V.; Renga, B.; Festa, C.; D’Amore, C.; Masullo, D.; Cipriani, S.; Di Leva, F.S.S.; Monti, M.C.C.; Novellino, E.; Limongelli, V.; et al. Modification on ursodeoxycholic acid (UDCA) scaffold. discovery of bile acid derivatives as selective agonists of cell-surface G-protein coupled bile acid receptor 1 (GP-BAR1). J. Med. Chem. 2014, 57, 7687–7701. [Google Scholar] [CrossRef] [PubMed]
  289. Jiang, C.; Xie, C.; Lv, Y.; Li, J.; Krausz, K.W.; Shi, J.; Brocker, C.N.; Desai, D.; Amin, S.G.; Bisson, W.H.; et al. Intestine-selective farnesoid X receptor inhibition improves obesity-related metabolic dysfunction. Nat. Commun. 2015, 6, 10166. [Google Scholar] [CrossRef] [PubMed]
  290. Renga, B.; Mencarelli, A.; D’Amore, C.; Cipriani, S.; D’Auria, M.V.V.; Sepe, V.; Chini, M.G.G.; Monti, M.C.C.; Bifulco, G.; Zampella, A.; et al. Discovery that theonellasterol a marine sponge sterol is a highly selective FXR antagonist that protects against liver injury in cholestasis. PLoS ONE 2012, 7, e30443. [Google Scholar] [CrossRef]
  291. Sepe, V.; Bifulco, G.; Renga, B.; D’Amore, C.; Fiorucci, S.; Zampella, A. Discovery of sulfated sterols from marine invertebrates as a new class of marine natural antagonists of farnesoid-X-receptor. J. Med. Chem. 2011, 54, 1314–1320. [Google Scholar] [CrossRef]
  292. Nadkarni, P.; Chepurny, O.G.; Holz, G.G. Regulation of glucose homeostasis by GLP-1. Prog. Mol. Biol. Transl. Sci. 2014, 121, 23–65. [Google Scholar] [CrossRef]
  293. Renga, B.; Mencarelli, A.; Vavassori, P.; Brancaleone, V.; Fiorucci, S. The bile acid sensor FXR regulates insulin transcription and secretion. Biochim. Biophys. Acta 2010, 1802, 363–372. [Google Scholar] [CrossRef]
  294. Düfer, M.; Hörth, K.; Wagner, R.; Schittenhelm, B.; Prowald, S.; Wagner, T.F.; Oberwinkler, J.; Lukowski, R.; Gonzalez, F.J.; Krippeit-Drews, P.; et al. Bile acids acutely stimulate insulin secretion of mouse β-cells via farnesoid X receptor activation and K(ATP) channel inhibition. Diabetes 2012, 61, 1479–1489. [Google Scholar] [CrossRef] [PubMed]
  295. Rizzo, G.; Disante, M.; Mencarelli, A.; Renga, B.; Gioiello, A.; Pellicciari, R.; Fiorucci, S. The farnesoid X receptor promotes adipocyte differentiation and regulates adipose cell function in vivo. Mol. Pharmacol. 2006, 70, 1164–1173. [Google Scholar] [CrossRef]
  296. Carino, A.; Cipriani, S.; Marchianò, S.; Biagioli, M.; Scarpelli, P.; Zampella, A.; Monti, M.C.C.M.C.M.C.; Fiorucci, S. Gpbar1 agonism promotes a Pgc-1α-dependent browning of white adipose tissue and energy expenditure and reverses diet-induced steatohepatitis in mice. Sci. Rep. 2017, 7, 13689. [Google Scholar] [CrossRef]
  297. Burkewitz, K.; Zhang, Y.; Mair, W.B. AMPK at the nexus of energetics and aging. Cell Metab. 2014, 20, 10–25. [Google Scholar] [CrossRef] [PubMed]
  298. Qu, Q.; Chen, Y.; Wang, Y.; Wang, W.; Long, S.; Yang, H.-Y.; Wu, J.; Li, M.; Tian, X.; Wei, X.; et al. Lithocholic acid binds TULP3 to activate sirtuins and AMPK to slow down ageing. Nature 2024. [Google Scholar] [CrossRef] [PubMed]
  299. Most, J.; Redman, L.M. Impact of calorie restriction on energy metabolism in humans. Exp. Gerontol. 2020, 133, 110875. [Google Scholar] [CrossRef]
  300. Montefusco, L.; D’Addio, F.; Loretelli, C.; Ben Nasr, M.; Garziano, M.; Rossi, A.; Pastore, I.; Plebani, L.; Lunati, M.E.; Bolla, A.M.; et al. Anti-inflammatory effects of diet and caloric restriction in metabolic syndrome. J. Endocrinol. Investig. 2021, 44, 2407–2415. [Google Scholar] [CrossRef]
  301. Kökten, T.; Hansmannel, F.; Ndiaye, N.C.; Heba, A.-C.; Quilliot, D.; Dreumont, N.; Arnone, D.; Peyrin-Biroulet, L. Calorie Restriction as a New Treatment of Inflammatory Diseases. Adv. Nutr. 2021, 12, 1558–1570. [Google Scholar] [CrossRef]
  302. Connell, E.; Le Gall, G.; Pontifex, M.G.; Sami, S.; Cryan, J.F.; Clarke, G.; Müller, M.; Vauzour, D. Microbial-derived metabolites as a risk factor of age-related cognitive decline and dementia. Mol. Neurodegener. 2022, 17, 43. [Google Scholar] [CrossRef]
  303. Yamashita, M.; Honda, A.; Shimoyama, S.; Umemura, M.; Ohta, K.; Chida, T.; Noritake, H.; Kurono, N.; Ichimura-Shimizu, M.; Tsuneyama, K.; et al. Breach of tolerance versus burden of bile acids: Resolving the conundrum in the immunopathogenesis and natural history of primary biliary cholangitis. J. Autoimmun. 2023, 136, 103027. [Google Scholar] [CrossRef]
  304. Bernstein, H.; Bernstein, C.; Payne, C.M.; Dvorakova, K.; Garewal, H. Bile acids as carcinogens in human gastrointestinal cancers. Mutat. Res. 2005, 589, 47–65. [Google Scholar] [CrossRef] [PubMed]
  305. Bernstein, H.; Bernstein, C.; Payne, C.M.; Dvorak, K. Bile acids as endogenous etiologic agents in gastrointestinal cancer. World J. Gastroenterol. 2009, 15, 3329–3340. [Google Scholar] [CrossRef] [PubMed]
  306. Bernstein, H.; Bernstein, C. Bile acids as carcinogens in the colon and at other sites in the gastrointestinal system. Exp. Biol. Med. 2023, 248, 79–89. [Google Scholar] [CrossRef]
  307. Fu, T.; Huan, T.; Rahman, G.; Zhi, H.; Xu, Z.; Oh, T.G.; Guo, J.; Coulter, S.; Tripathi, A.; Martino, C.; et al. Paired microbiome and metabolome analyses associate bile acid changes with colorectal cancer progression. Cell Rep. 2023, 42, 112997. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effects of SCFAs. Intestinal microbiota bacteria, particularly species within the Bacteroidetes and Firmicutes phyla, produce short-chain fatty acids (SCFAs) through the fermentation of partially and non-digestible polysaccharides. SCFAs function as signaling molecules by activating three membrane receptors: FFAR3 (also known as GPR41), FFAR2 (also known as GPR43), and HCAR2 (also known as GPR109A). The activation of these receptors elicits anti-inflammatory effects, preserves intestinal barrier integrity, and reduces fat accumulation by enhancing insulin secretion and energy expenditure.
Figure 1. Effects of SCFAs. Intestinal microbiota bacteria, particularly species within the Bacteroidetes and Firmicutes phyla, produce short-chain fatty acids (SCFAs) through the fermentation of partially and non-digestible polysaccharides. SCFAs function as signaling molecules by activating three membrane receptors: FFAR3 (also known as GPR41), FFAR2 (also known as GPR43), and HCAR2 (also known as GPR109A). The activation of these receptors elicits anti-inflammatory effects, preserves intestinal barrier integrity, and reduces fat accumulation by enhancing insulin secretion and energy expenditure.
Cells 14 00595 g001
Figure 2. Bile acid biosynthesis and enterohepatic recirculation. Primary bile acids—cholic acid (CA) and chenodeoxycholic acid (CDCA)—along with their taurine and glycine conjugates, are synthesized in the liver from cholesterol via two distinct metabolic routes. The classical (neutral) pathway is initiated by CYP7A1, whereas the alternative (acidic) pathway begins with CYP27A1. Following a meal, these bile acids are secreted into the small intestine through the common bile duct, where they facilitate lipid emulsification and absorption. In the gut, microbial enzymes further transform bile acids, yielding secondary bile acids such as deoxycholic acid (DCA) and lithocholic acid (LCA) along with their derivatives. In the ileum, bile acids are absorbed by enterocytes via the ASBT transporter; this uptake activates the farnesoid X receptor (FXR), which in turn downregulates ASBT to restrict further reabsorption. Subsequently, bile acids are conveyed from enterocytes to the portal vein through MRP3 and OSTα/β transporters, returning to the liver where hepatocytes reabsorb them via OATPs and NTCP. Elevated intracellular bile acid concentrations in hepatocytes activate the FXR/SHP signaling axis, thereby suppressing further bile acid synthesis through inhibition of CYP7A1. Moreover, hepatocytes export bile acids into the systemic circulation via MRP4 [94], OSTα/β and MRP3, while also exporting them back into bile ducts through MDR2, MRP2 and BSEP. Abbreviations: CYP7A1, cytochrome P450 7A1; CYP27A1, cytochrome P450 27A1; CYP8B1, cytochrome P450 8B1; CA, cholic acid; CDCA, chenodeoxycholic acid; BAAT, bile acid-CoA:amino acid N-acyltransferase; BA, bile acid; BSH, bile salt hydrolases; HSDH, 7α-hydroxysteroid dehydrogenase; ASBT, apical sodium-dependent bile acid transporter; MRP2, multidrug resistance-associated protein 2; FXR, farnesoid X receptor; MRP3, multidrug resistance-associated protein 3; NTCP, sodium/taurocholate co-transporting polypeptide; OATPs, organic anion-transporting polypeptides; MDR2, multidrug resistance protein 2; BSEP, bile salt export pump; MRP4, multidrug resistance-associated protein 4.
Figure 2. Bile acid biosynthesis and enterohepatic recirculation. Primary bile acids—cholic acid (CA) and chenodeoxycholic acid (CDCA)—along with their taurine and glycine conjugates, are synthesized in the liver from cholesterol via two distinct metabolic routes. The classical (neutral) pathway is initiated by CYP7A1, whereas the alternative (acidic) pathway begins with CYP27A1. Following a meal, these bile acids are secreted into the small intestine through the common bile duct, where they facilitate lipid emulsification and absorption. In the gut, microbial enzymes further transform bile acids, yielding secondary bile acids such as deoxycholic acid (DCA) and lithocholic acid (LCA) along with their derivatives. In the ileum, bile acids are absorbed by enterocytes via the ASBT transporter; this uptake activates the farnesoid X receptor (FXR), which in turn downregulates ASBT to restrict further reabsorption. Subsequently, bile acids are conveyed from enterocytes to the portal vein through MRP3 and OSTα/β transporters, returning to the liver where hepatocytes reabsorb them via OATPs and NTCP. Elevated intracellular bile acid concentrations in hepatocytes activate the FXR/SHP signaling axis, thereby suppressing further bile acid synthesis through inhibition of CYP7A1. Moreover, hepatocytes export bile acids into the systemic circulation via MRP4 [94], OSTα/β and MRP3, while also exporting them back into bile ducts through MDR2, MRP2 and BSEP. Abbreviations: CYP7A1, cytochrome P450 7A1; CYP27A1, cytochrome P450 27A1; CYP8B1, cytochrome P450 8B1; CA, cholic acid; CDCA, chenodeoxycholic acid; BAAT, bile acid-CoA:amino acid N-acyltransferase; BA, bile acid; BSH, bile salt hydrolases; HSDH, 7α-hydroxysteroid dehydrogenase; ASBT, apical sodium-dependent bile acid transporter; MRP2, multidrug resistance-associated protein 2; FXR, farnesoid X receptor; MRP3, multidrug resistance-associated protein 3; NTCP, sodium/taurocholate co-transporting polypeptide; OATPs, organic anion-transporting polypeptides; MDR2, multidrug resistance protein 2; BSEP, bile salt export pump; MRP4, multidrug resistance-associated protein 4.
Cells 14 00595 g002
Figure 3. Systemic effects of secondary bile acids. Within the gastrointestinal tract, the resident microbiota converts primary bile acids into their secondary counterparts. Nearly 90% of these bile acids are reabsorbed in the ileum by enterocyte and returned to the liver through the portal circulation, a process facilitated by OATP and NTCP transporters. In the colon, secondary bile acids engage GPBAR1 receptors on enteroendocrine L-cells, thereby triggering GLP-1 secretion. GLP-1 subsequently enhances satiety, decelerates gastric emptying and augments insulin release via activation of GLP-1R on neuronal networks, gastric tissues and pancreatic β-cells. Moreover, systemic activation of GPBAR1 in adipose tissue, skeletal muscle, blood vessels and immune cells by secondary bile acids promotes, respectively, thermogenic activity through T4-to-T3 conversion and vasodilation and confers anti-inflammatory effects.
Figure 3. Systemic effects of secondary bile acids. Within the gastrointestinal tract, the resident microbiota converts primary bile acids into their secondary counterparts. Nearly 90% of these bile acids are reabsorbed in the ileum by enterocyte and returned to the liver through the portal circulation, a process facilitated by OATP and NTCP transporters. In the colon, secondary bile acids engage GPBAR1 receptors on enteroendocrine L-cells, thereby triggering GLP-1 secretion. GLP-1 subsequently enhances satiety, decelerates gastric emptying and augments insulin release via activation of GLP-1R on neuronal networks, gastric tissues and pancreatic β-cells. Moreover, systemic activation of GPBAR1 in adipose tissue, skeletal muscle, blood vessels and immune cells by secondary bile acids promotes, respectively, thermogenic activity through T4-to-T3 conversion and vasodilation and confers anti-inflammatory effects.
Cells 14 00595 g003
Table 1. Definitions of postbiotics before the 2019 ISAPP consensus.
Table 1. Definitions of postbiotics before the 2019 ISAPP consensus.
Definition of PostbioticsMicrobial Cells IncludedReference
Any factor resulting from the metabolic activity of a probiotic or any released molecule capable of conferring beneficial effects to the host in a direct or indirect wayNo[2]
Soluble factors (products or metabolic byproducts), secreted by live bacteria, or released after bacterial lysis, such as enzymes, peptides, teichoic acids, peptidoglycan-derived muropeptides, polysaccharides, cell surface proteins and organic acidsNo[3]
Compounds produced by microorganisms, released from food components or microbial constituents, including non-viable cells that, when administered in adequate amounts, promote health and well-beingYes[4]
Non-viable metabolites produced by probiotics that exert biological effects on the hostsNo[5]
Non-viable bacterial products or metabolic byproducts from probiotic microorganisms that have positive effects on the host or microbiotaYes[6]
Functional bioactive compounds, generated in a matrix during fermentation, which may be used to promote healthYes[7]
Table 2. Postbiotics overview.
Table 2. Postbiotics overview.
CompoundsMain Function(s)
Short-Chain Fatty Acids (SCFAs)
Acetate, propionate, butyrate
Anti-inflammatory and immunomodulatory effects; glucose and lipid metabolism regulation; energy production contribution
Lactic AcidEssential for glycolysis in hypoxia or anaerobic conditions; healthy vaginal environment maintenance
BacteriocinsImmunomodulatory functions; antimicrobial activity
Secondary Bile Acids
Deoxycholic Acid (DCA), Lithocholic Acid (LCA)
Dietary lipid adsorption facilitation; anti-inflammatory and immunomodulatory activity, apoptosis signaling modulation [15]; anti-aging benefits [16]; tight junctions stability maintenance [17]
Bacterial Cell Wall Components
Exopolysaccharides (EPS), Peptidoglycan (PG), Lipoteichoic Acid (LTA)
Anti-inflammatory activity; SCFA synthesis upregulation; wound healing and tissue regeneration promotion; ion homeostasis maintenance; immune system response enhancement
Plasmalogens (Pls)Anti-inflammatory and antioxidant activities; adipogenesis regulation; cognitive function improvement
Intestinal bacteria-derived vitamins
Water-soluble (B1, B2, B5, B7, B9, B12), fat-soluble (K)
Immunomodulatory and antioxidant activity; Calcium homeostasis and bone health maintenance; blood clotting regulation
Tryptophan metabolitesNeurotransmitters level regulation and brain health support; anti-inflammatory and antimicrobial activity; intestinal epithelial barrier function enhancement; insulin resistance and lipid metabolism promotion
Conjugated Linoleic Acids (CLAs)Anti-breast cancer, anti-inflammatory and immunomodulatory activity; lipolysis potentiation; atherosclerosis inhibition; osteoporosis prevention
PolyaminesIntestinal epithelial barrier integrity and epithelial renewal maintenance; immunomodulatory activity; B cell senescence reversion
Phenolic compoundsAnti-inflammatory and antioxidant properties; neuroprotective and cardioprotective effects
Hydrogen Peroxide (H2O2)Antimicrobial activity (colonization resistance)
Organic Acids (OAs)Antimicrobial activity (colonization resistance)
Glutathione (GSH)Antioxidant, anti-inflammatory and immunomodulatory activity; bone health maintenance
Microbial enzymes
Proteases, lipases
Food digestion and adsorption facilitation; anti-inflammatory, immunomodulatory, antimicrobial and anticancer activity; lipid metabolism modulation
Table 3. SCFAs’ receptors and main functions.
Table 3. SCFAs’ receptors and main functions.
ReceptorCellular ExpressionFunction(s)
FFAR3 (GPR41)ColonocytesSensor for luminal SCFAs [42]
Pancreatic β-cellGSIS regulation [43,44]
Sympathetic gangliaHeart rate and energy expenditure increase [45]
DCsQuiescent/tolerogenic DC induction [41,46,47]
EnterocytesTJ enhancement and intestinal barrier integrity maintenance [48], NF-κB pathway inhibition [38]
Goblet cellsMucus secretion [49]
FFAR2 (GPR43)Enteroendocrine L cellsGLP-1 release [50]
WATReduction of lipolysis and fat accumulation [51]
Pancreatic β-cellGSIS regulation [43]
NeutrophilsChemotactic effect and neutrophil activation (phagocytic activity and ROS formation) [52,53]
Treg cellsMaintenance of intestinal immune homeostasis [54]
DCsQuiescent/tolerogenic DC induction [41,46,47]
EnterocytesTJ enhancement and intestinal barrier integrity maintenance [48], NF-κB pathway inhibition [38]
Goblet cellsMucus secretion [49]
HCAR2 (GPR109A)EnterocytesTJ enhancement and intestinal barrier integrity maintenance [48], NF-κB pathway inhibition [38]
AdipocytesMetabolic sensor for lipolysis suppression during starvation [55]
NeutrophilsApoptosis induction [56]
DCsQuiescence/tolerogenic DC induction [41,46,47], IL-10 secretion [57]
ColonocytesTumor suppressor [58]
DCs: dendritic cells; FFAR: free fatty acid receptor; GLP-1: glucagon-like peptide 1; GPR: G protein-coupled receptor; GSIS: glucose-stimulated insulin secretion; HCAR: hydroxycarboxylic receptor; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; ROS: reactive oxygen species; TJs: tight junctions; WAT: white adipose tissue.
Table 4. Some of the most common bacteriocins secreted by Gram-positive (GPB) and Gram-negative (GNB) bacteria.
Table 4. Some of the most common bacteriocins secreted by Gram-positive (GPB) and Gram-negative (GNB) bacteria.
BacteriocinProducerMechanism(s) of Action
Nisin ALactococcus lactisMurein synthesis inhibition [78]; induction of preferential apoptosis and cell cycle arrest, reduction of cell proliferation [79]
EpiderminStaphylococcus epidermidisMurein and WTA synthesis inhibition [80]
GalliderminStaphylococcus gallinarumStaphylococci growth inhibition and biofilm formation prevention [81]
MersacidinBacillus sp.Peptidoglycan synthesis inhibition [82]
SublacinBacillus subtilisDNA, RNA and protein synthesis inhibition [83]
LysostaphinStaphylococcus simulansCell wall lytic enzyme (endopeptidase activity) [84]
Enterocin AEnterococcus faeciumTarget cell membrane pore formation [85]
ThiazomycinAmycolatopsis fastidiosaProtein synthesis inhibitor [86]
Microcin LEscherichia coliDNA/RNA folding and/or synthesis inhibition [87]
Microcin E429Klebsiella pneoumoniaeTarget cell membrane pore formation [88]
WTA: wall teichoic acid.
Table 5. Distribution and function of secondary-bile-acid-activated receptors in immune cells.
Table 5. Distribution and function of secondary-bile-acid-activated receptors in immune cells.
ReceptorImmune Cell
Distribution
Function(s)Reference
GPBAR1Monocytes/macrophagesDownregulation of inflammatory cytokines (TNFα, IFNγ, IL-6, IL-1β) and upregulation of anti-inflammatory ones (IL-10); downregulation of CCL2, subsequent suppression of macrophages migration and polarization facilitation toward the M2 anti-inflammatory phenotype; inhibition of NLRP3 inflammasome activation[226,227,228,229,230,231,232]
KCsInhibition of LPS-induced cytokine expression via cAMP-dependent pathways[231,233]
DCsInhibition of NF-κB pro-inflammatory cytokines; induction of a tolerogenic/quiescent state; apoptosis promotion[234,235]
NKT cellsRegulation of type I and II NKT polarization and induction of a tolerogenic phenotype; upregulation of anti-inflammatory cytokines (IL-10)[236]
FXRMonocytes/macrophagesDownregulation of inflammatory cytokines (IL-1β, TNFα); inhibition of NLRP3 inflammasome activation; polarization facilitation toward the M2 anti-inflammatory phenotype.[237,238,239,240,241]
KCsDownregulation of inflammatory cytokines (TNFα, IL-6, IL-1β) and upregulation of anti-inflammatory ones (IL-10); downregulation of CCL2.[242,243]
DCsDownregulation of MdCAM-1 in the inflamed site with subsequent retention of DCs in the spleen.[244]
ILCsRegulation of ILC commitment towards functional active ILC2 and ILC3 subtypes.[245]
NKT cellsDownregulation of OPN.[246]
CD8+ T lymphocytesRegulation of the immune response based on the nutritional status via imitation of the metabolic flexibility of CD8+ effector T cells.[247]
PXRMonocytes/macrophagesDownregulation of TLR4 signaling; upregulation of inflammatory cytokines (IL-1β) via caspase-1 activation.[248,249]
Th lymphocytesDownregulation of NF-κB and IFNγ.[250]
B lymphocytesPossible attenuation of B1 cell production.[251]
VDRMonocytes/macrophagesUpregulation of MKP-1 and subsequent downregulation of inflammatory cytokines (IL-6, TNFα).[252]
KCsAnti-inflammatory effects in liver steatosis; protection against hepatic endoplasmic reticulum stress.[253,254]
DCsInduction of tolerogenic DCs; inhibition of mature DCs; inhibition of TIM4 gene expression.[255,256,257,258]
NKT cellsRegulation of iNKT cell development and function.[259]
CD8+ T lymphocytesPrevention of CD8+ T cell proliferation.[260]
Th lymphocytesInhibition of Th1 cell response.[261,262]
RORγtILCsPromotion of ILC3 differentiation and function.[263,264]
Th lymphocytesPromotion of Th17 differentiation; inhibition of Treg cell differentiation.[265,266]
CCL2: C-C motif chemokine ligand 2; DCs: dendritic cells; IFNγ: interferon γ; IL: interleukin; ILCs: innate lymphoid cells; KCs: Kupffer cells; LPS: lipopolysaccharide; MAdCAM-1: mucosal vascular addressin cell adhesion molecule 1; MKP-1: MAP kinase phosphatase 1; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; NKT: natural killer T cells; NLRP3: NLR family pyrin domain containing 3; OPN: osteopontin; Th: T helper cells; TLR: Toll-like receptor; TNFα: tumor necrosis factor α.
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

Urbani, G.; Rondini, E.; Distrutti, E.; Marchianò, S.; Biagioli, M.; Fiorucci, S. Phenotyping the Chemical Communications of the Intestinal Microbiota and the Host: Secondary Bile Acids as Postbiotics. Cells 2025, 14, 595. https://doi.org/10.3390/cells14080595

AMA Style

Urbani G, Rondini E, Distrutti E, Marchianò S, Biagioli M, Fiorucci S. Phenotyping the Chemical Communications of the Intestinal Microbiota and the Host: Secondary Bile Acids as Postbiotics. Cells. 2025; 14(8):595. https://doi.org/10.3390/cells14080595

Chicago/Turabian Style

Urbani, Ginevra, Elena Rondini, Eleonora Distrutti, Silvia Marchianò, Michele Biagioli, and Stefano Fiorucci. 2025. "Phenotyping the Chemical Communications of the Intestinal Microbiota and the Host: Secondary Bile Acids as Postbiotics" Cells 14, no. 8: 595. https://doi.org/10.3390/cells14080595

APA Style

Urbani, G., Rondini, E., Distrutti, E., Marchianò, S., Biagioli, M., & Fiorucci, S. (2025). Phenotyping the Chemical Communications of the Intestinal Microbiota and the Host: Secondary Bile Acids as Postbiotics. Cells, 14(8), 595. https://doi.org/10.3390/cells14080595

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