Next Article in Journal
Bacterial Microbiota of Ostreobium, the Coral-Isolated Chlorophyte Ectosymbiont, at Contrasted Salinities
Next Article in Special Issue
Microbiome-Related and Infection Control Approaches to Primary and Secondary Prevention of Clostridioides difficile Infections
Previous Article in Journal
Carbapenem-Resistant Pseudomonas aeruginosa Bacteremia, through a Six-Year Infection Control Program in a Hospital
Previous Article in Special Issue
Isolation and Cultivation of Human Gut Microorganisms: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Amino Acid-Derived Bacterial Metabolites in the Colorectal Luminal Fluid: Effects on Microbial Communication, Metabolism, Physiology, and Growth

by
François Blachier
Université Paris-Saclay, AgroParisTech, INRAe, UMR PNCA, 91120 Palaiseau, France
Microorganisms 2023, 11(5), 1317; https://doi.org/10.3390/microorganisms11051317
Submission received: 6 April 2023 / Revised: 14 May 2023 / Accepted: 15 May 2023 / Published: 17 May 2023
(This article belongs to the Special Issue Latest Review Papers in Gut Microbiota 2023)

Abstract

:
Undigested dietary and endogenous proteins, as well as unabsorbed amino acids, can move from the terminal part of the ileum into the large intestine, where they meet a dense microbial population. Exfoliated cells and mucus released from the large intestine epithelium also supply nitrogenous material to this microbial population. The bacteria in the large intestine luminal fluid release amino acids from the available proteins, and amino acids are then used for bacterial protein synthesis, energy production, and in other various catabolic pathways. The resulting metabolic intermediaries and end products can then accumulate in the colorectal fluid, and their concentrations appear to depend on different parameters, including microbiota composition and metabolic activity, substrate availability, and the capacity of absorptive colonocytes to absorb these metabolites. The aim of the present review is to present how amino acid-derived bacterial metabolites can affect microbial communication between both commensal and pathogenic microorganisms, as well as their metabolism, physiology, and growth.

Graphical Abstract

1. Introduction

The process of protein digestion in the small intestine is efficient, with a yield equal to or even above 90% for most alimentary proteins [1]. Following protein digestion in the small intestinal luminal fluid by pancreatic proteases, peptides are further degraded by epithelial peptidases, and released oligopeptides and amino acids are finally absorbed in the portal blood through the small intestine epithelium [2,3,4].
A minor part of undigested or not fully digested proteins, together with other nitrogenous compounds, can move from the terminal part of the ileum to the large intestine. Indeed, it has been estimated in volunteers that between 1.5 and 5 g of nitrogen are transferred every day through the ileo-caecal junction [5,6,7,8,9], with the major parts of this nitrogenous material being proteins and, to a much lesser extent, peptides, while the minor parts of this nitrogenous material are made of amino acids, urea, ammonia, nitrate, and other compounds [8]. Regarding peptides, it has been shown that these compounds are more rapidly absorbed in the small intestine than free amino acids [10,11]. Of note, the nitrogenous material recovered in the distal small intestine is not only originating from the diet but also from endogenous sources, including proteins present in the exocrine secretion, in the fully mature exfoliated enterocytes, and in the mucus layer released in the luminal fluid [12]. It has been determined that roughly 40% of the nitrogenous material transferred to the large intestine is from alimentation, while the remaining 60% is of endogenous origin [9].
By using a conversion factor between nitrogen and protein equal to 6.25 [13], it can be calculated as an approximation that roughly between 4 and 12 g of alimentary proteins escape digestion in the small intestine. This estimation fits rather well with the mean protein consumption in Western countries, which averages 85 g per day [14,15]. Indeed, if we assume a mean protein digestibility efficiency of approximately 90% [16], we can calculate that 8 g of alimentary proteins would be recovered in the large intestine, a value within the 4–12 g range as calculated above.
In addition to the nitrogenous material originating from the small intestine, exfoliation of fully mature colonic epithelial cells and mucus renewal in the large intestine may also contribute to the global amount of nitrogenous compounds available for the microbial population present in the colorectal fluid (Figure 1). In the large intestine luminal fluid, this nitrogenous material meets a dense population of microbes in a context of extended transit [17]. A part of the water in the luminal fluid contained within the large intestine is progressively absorbed by the epithelial cells from the proximal to the distal part, resulting in a median value of water content in the human fecal material equal to 75% of the total mass [18].
The gut microbiota is composed of a vast population of bacteria but also of archaea [19], viruses [20,21], and fungi [22]. Protozoans in the intestine, although classically not included as part of the microbiota itself, represent a heterogeneous group of eukaryotic organisms, with some of them being considered as parasites [23]. The term gut ecosystem usually refers to the biological community of microorganisms living in the exogenous environment of the gut. The metabolism of proteins and amino acids by the intestinal microbes has been mostly studied for the bacterial part of the intestinal microbiota [24,25], and as explained in the following paragraph, these compounds are used in bacteria in both anabolic and catabolic pathways.

2. Bacteria in the Large Intestinal Luminal Fluid Release Amino Acids from Undigested Proteins and Used Them for Protein Synthesis, Energy Production, and Other Catabolic Pathways Which Release Various Bacterial Metabolites

In the luminal fluid of the large intestine, proteins and peptides are degraded by the bacterial proteases and peptidases, then released as amino acids [26,27,28]. The proteolytic activities in the large intestine have been mainly attributed to the genera Bacteroides, Clostridium, Propionibacterium, Fusobacterium, Streptococcus, and Lactobacillus [29]. Some bacteria, such as lactic acid bacteria, have developed proteolytic systems, presumably to compensate for their reduced or even absent capabilities to synthesize amino acids [30]. These proteolytic systems include extracellular proteases that degrade proteins into oligopeptides, and this degradation is followed by the entry of oligopeptides within bacteria via dedicated transporters. Furthermore, intracellular peptidases degrade the peptides into amino acids [31,32]. Amino acids and their corresponding metabolites can be imported and exported from the bacteria via transmembrane proteins [33,34], allowing regulation of the concentration of these compounds in the bacterial cells and allowing exchanges of amino acids and related metabolites between bacterial species. Efflux systems for some amino acids such as lysine, arginine, threonine, cysteine, leucine, isoleucine, and valine have been studied in the bacteria E. coli and Corynebacterium glutamicum [35,36].
In the large intestine, amino acids resulting from the transfer of amino acids from the small to the large intestine, together with amino acids released from proteins and peptides that have not been digested in the small intestine and amino acids released from endogenous proteins and peptides in the large intestine, are used by the intestinal microbiota for their metabolism. Amino acid availability for microbial metabolism in the large intestine is presumably little decreased by amino acid absorption through the colonocytes since, except in the neonatal period [37,38], the colonic mucosa does not absorb amino acids to any significant extent [39]. However, several amino acid transporters, as well as the peptide transporter PepT1 have been identified in the large intestine epithelium [40], suggesting that small amounts of amino acids would be transported inside and/or through colonocytes [39].
Since approximately half of the bacterial biomass in the colon is lost every day by defecation [41], the rapid growth of bacteria within the large intestine luminal fluid requires extensive amounts of substrates, including amino acids, that are mainly provided by the host. This intense bacterial metabolism and associated anabolism are possible only if the required substrates are provided in sufficient quantities [24]. Many metabolic pathways involved in amino acid synthesis in bacteria are conserved in the different bacterial lineages, including those bacteria that inhabit the large intestine [42,43]. However, major differences remain when comparing the metabolic capacities for amino acid synthesis at the level of species and strains. As presented above, some bacteria, such as lactic acid bacteria, possess low or even absent capacities for amino acid biosynthesis. Another illustrative example of such bacteria is represented by Clostridium perfringens, which displays no metabolic capacity for the synthesis of glutamate, arginine, histidine, lysine, methionine, serine, threonine, aromatic, and branched-chain amino acids [44], thus depending on the host for the supply of these amino acids for their metabolic needs. Lactobacillus johnsonii is another gut bacterium that is unable to perform the synthesis of almost all amino acids due to the lack of numerous anabolic pathways required for amino acid synthesis [45]. Other intestinal bacteria, including Enterococcus faecalis and Streptococcus agalactiae, perform the synthesis of only a few specific amino acids in mammals (including humans) [46]. In sharp contrast, other intestinal bacteria such as Clostridium acetobutylicum are equipped with a complete set of genes required for the biosynthesis of all amino acids [47]. However, it is important to keep in mind that the sole presence of genes implicated in one given amino acid synthesis within one given bacterium does not allow for a conclusion on the functionality of the corresponding metabolic pathways. A typical example of such limitation is given by the common gut bacterium Lactococcus lactis, for which the genes allowing the synthesis of the 20 amino acids have been identified but which requires supplementation with isoleucine, valine, leucine, histidine, methionine, and glutamate for growth since genes involved in the biosynthetic pathways corresponding to these amino acids have been demonstrated to be non-functional due to point mutations [48,49].
Bacteria may incorporate the available amino acids into proteins or may use them as energy substrates and in various catabolic pathways [50] (Figure 1). Most gut bacteria utilize amino acids and ammonia as preferred nitrogen sources, and amino acids such as glutamine, glutamate, asparagine, aspartate, lysine, serine, threonine, arginine, glycine, histidine, and branched-chain amino acids are preferred substrates for degradation by gut bacteria, with numerous products being formed, notably ammonia, short-chain fatty acids, branched-chain fatty acids, phenols, indoles, organic acids, amines, and various gaseous compounds [51].
In healthy adults, the distal part of the intestinal tract harbors large communities of obligate anaerobes [52,53]. In the virtual absence of oxygen or other suitable electron acceptors, only strict or facultative anaerobic bacteria, such as Clostridia and Fusobacteria, can utilize amino acids as energy sources, thus fermenting amino acids and producing mainly branched-chain fatty acids and ammonia [54], as well as short-chain fatty acids, hydrogen (H2), carbon dioxide (CO2), hydrogen sulfide (H2S), and various organic compounds [55,56] as listed above. Several mechanisms for amino acid degradation have been described in anaerobic bacteria, including the Strickland reaction that is operative in several proteolytic Clostridia. This latter reaction involves the coupled oxidation and reduction of amino acids to organic acids. Other fermentation pathways described in various Clostridia as well as in Fusobacterium spp. and Acidaminococcus spp. involve amino acids that act as electron donors or acceptors [57,58]. The genus Clostridium contains specific amino acid degradation pathways, such as B1-dependent aminomutases, selenium-containing oxidoreductases, and oxygen-sensitive 2-hydroxyacyl-CoA dehydratases [59]. Amino acids can also be decarboxylated, ultimately yielding biogenic amines and polyamines. Of note, luminal parameters such as pH can modulate the catalytic activity of different bacterial enzymes such as deaminases and decarboxylases, thus affecting the production of specific metabolic end products [55].
Some amino acids, depending on the bacterial species considered, may be utilized for specific metabolic pathways. For instance, Clostridium stricklandii preferentially uses threonine, arginine, and serine as carbon sources and for energy production, but little utilizes glutamate, aspartate, and aromatic amino acids for such purposes and uses lysine as fuel only in the stationary growth phase [59]. The ratio of available carbohydrates over proteins is central for fixing substrate utilization by the gut microbiota in different contexts of substrate availability [54], and in humans, higher availability of complex carbohydrates (like plant fibers and resistant starch) lowers the process of protein fermentation by the intestinal bacteria, as determined by measuring several amino acid-derived bacterial metabolites in feces [60,61,62]. In addition, when fermentable carbohydrates are abundant for intestinal bacteria, amino acids are mostly used for anabolic metabolism but little for energy production [63]. On the other hand, in a context of low availability of fermentable carbohydrates, several amino acids are used by intestinal bacteria for energy production, thus supporting bacteria growth [42,50,64]. Due to high carbohydrate fermentation in the proximal colon, there is a progressive decrease in carbohydrate availability in the more median and distal parts of the colon, thus resulting in higher protein degradation and amino acid utilization [65]. Differential use of substrates when comparing different bacterial species has been documented. For instance, most genera in the phylum Firmicutes preferentially use proteins among the available substrates for their metabolism [66].
Relatively few nutritional intervention studies in volunteers have examined the effects of increasing the amounts of proteins in the diet on the gut microbiota composition and metabolic activity. In several of these studies, the dietary protein intake between the groups of volunteers was not the sole parameter modified, since modifications of energy and/or fiber were noticed in these studies. These two latter parameters are known to affect the gut microbiota composition [67,68,69,70,71,72], rendering correct interpretation of the results obtained from these studies difficult. Three studies have used high-protein diets without modification of dietary fiber or energy intake [73,74,75]. In these latter studies, the high-protein diets were made isocaloric by decreasing the amount of carbohydrates in the diet. Under such conditions, these three studies found little change in the fecal and rectal-associated bacterial composition after several weeks of consumption of the high-protein diet. In contrast, in the three studies, marked changes in the amino acid-derived bacterial metabolites were recorded both in feces and urine, reinforcing the view that luminal substrate availability is central to fixing bacterial metabolite concentrations in the luminal fluids.
However, substrate availability is not the sole parameter that influences the concentrations of amino acid-derived bacterial metabolites in biological fluids. Microbial composition and its overall metabolic activity, as well as the absorption of bacterial metabolites through colonocytes, are also believed to influence the luminal concentrations of these compounds [24]. Incidentally, approximately half of the metabolic pathways used by the intestinal bacteria do not occur in the cells of the host, and the largest group of such metabolic pathways involves pathways related to amino acid metabolism [76]. Other physiological parameters, such as transit time, may also influence the concentrations of amino acid-derived metabolites in the colon. Indeed, longer transit times are associated with higher levels of protein fermentation [77,78].
Finally, as explained above, the utilization of the different amino acids by the intestinal bacteria produces numerous amino acid-derived metabolites that are detected in the different biological fluids [24], and several of them, as detailed in the next paragraph, have been shown to play a role in the biology of intestinal microbes.
Although the rate of amino acid catabolism by the bacteria of the large intestine represents a major parameter to determine the concentrations of several amino acid-derived bacterial metabolites in the colorectal fluid, the rate of bacterial amino acid synthesis also likely plays a role in fixing such concentrations. In fact, intestinal bacteria use bacterial metabolites that are produced during amino acid catabolism (such as ammonia and hydrogen sulfide) [42] in the process of amino acid synthesis (Figure 1).

3. Amino Acid-Derived Bacterial Metabolites Are Involved in the Biology of Intestinal Microbes

Several amino acid-derived bacterial metabolites have been shown to be implicated in microbial communication and in microbial metabolism, physiology, and growth. Of note, as presented below, these effects involve both commensal and pathogenic intestinal microorganisms.

3.1. Lactate, Formate, Succinate and Oxaloacetate

During the catabolism of amino acids, the intestinal bacteria produce numerous organic acids, including lactate, formate, succinate, and oxaloacetate [29,50]. Indeed, high-protein diet consumption increases the amounts of organic acids in the large intestine luminal content [79]. Of note, these organic acids are not exclusively produced by the intestinal microbiota from amino acids but also from carbohydrates [80,81,82]. Among these organic acids, lactate is used as carbon and an energy source by indigenous bacteria, including Salmonella and Campylobacter [83]. Formate, which is secreted by the pathogenic bacteria Shigella flexneri, has been shown to promote the expression of genes involved in their virulence [84]. Oxaloacetate, when produced by Escherichia coli helps the parasite Entamoeba histolytica to survive in the large intestine [85] (Figure 2).
This latter result is of notable importance given that this parasite can trigger a strong inflammatory response upon invasion of the colonic mucosa. In addition, this study shows that communication between bacterial species and parasites through dedicated metabolites may occur in the large intestine. Lastly, succinate produced by the gut microbiota has been shown to promote infection by Clostridium difficile after antibiotic treatment [86]. Furthermore, Clostridium butyricum appears to be able to prevent Clostridium difficile proliferation by diminishing the succinate concentration in the large intestine luminal content [87]. Interestingly, this latter decreased succinate concentration is apparently the net result of overall modifications of the microbiota’s metabolic activity [87].

3.2. p-Cresol

The metabolite p-cresol (4-methylphenol) is produced by the intestinal microbiota of the large intestine from the amino acid tyrosine [88]. From in vitro analysis, it has been shown that among the bacteria present in the human gut, specific families of bacteria, namely Fusobacteriaceae, Enterobacteriaceae, Clostridium, and Coriobacteriaceae are strong p-cresol producers [89]. As expected, by increasing the protein content in the diet of mammals, an increased p-cresol concentration is measured in the feces [90]. On the contrary, the fecal excretion of p-cresol is diminished when the diet is enriched with undigestible polysaccharides [60].
Of major interest, the production of p-cresol by Clostridium difficile gives this bacterium a competitive advantage over other gut bacteria such as Escherichia coli, Klebsiella oxytoca, and Bacteroides thetaiomicron [91,92]. By using a model of Clostridium difficile infection in rodents, it has been demonstrated that excessive p-cresol production affects the gut microbiota diversity [91]. Furthermore, by removing the capacity of Clostridium difficile to produce p-cresol, the capacity of this bacterium to recolonize the intestine after an initial infection is diminished [91] (Figure 2). These results are carrying potential applications as Clostridium difficile is a major cause of diarrhea and inflammation in patients following long-term antibiotic treatment [93].

3.3. Indole

Indole is produced from tryptophan by various Gram-positive and Gram-negative species, including Escherichia coli, Proteus vulgaris, Clostridium spp. and Bacteroides spp. [94,95,96]. Indole diminishes cell motility and aggregation in L. monocytogenes [97] (Figure 2). Indole diminishes the virulence of Pseudomonas aeroginosa [98], of Salmonella enterica [99], and the virulence and growth of the fungal species Candida albicans [100]. In another study, indole was found to be bacteriostatic against lactic acid bacteria while affecting their survival [101]. Of note, indole mitigates cytotoxicity by Klebsellia spp. by suppressing toxin production by this bacterium [102]. In addition, indole impairs the ability of enteropathogenic Escherichia coli to enhance its virulence activity in response to Vibrio cholerae [103]. Lastly, Clostridium difficile, which itself is not considered an indole producer, increases indole production by other gut bacteria [104]. By doing so, Clostridium difficile increases the concentration of indole within the intestinal fluid, then limiting growth of indole-sensitive bacteria, and finally adversely improving Clostridium difficile growth and persistence within the intestinal luminal content [104].
Thus, if there is little doubt that indole is involved in intestinal microbe physiology and growth, the beneficial versus deleterious effects of this bacterial metabolite for gut health appears to depend on the overall colorectal ecosystem characteristics.

3.4. Skatole

Skatole is produced by the intestinal microbiota from tryptophan [54]. This bacterial metabolite has been shown to display a marked inhibitory effect on enterohemorrhagic Escherichia coli biofilm formation [105] (Figure 2). Briefly, biofilms are structures formed by the colon microbiota that are in contact with the mucosal surface and that play a role in the modulation of epithelial barrier function [106].

3.5. Hydrogen Sulfide

Hydrogen sulfide (H2S) is produced by numerous bacterial species in the large intestine from both dietary and endogenous S-containing substrates [107]. These substrates include cysteine, but also sulfate, taurine, and sulfomucins [108,109,110,111]. By using in vitro tests with human fecal material, it has been found that, for instance, production of H2S is greatly enhanced by sulfur-containing amino acids but much more modestly by inorganic sulfate [112]. In addition, H2S production is diminished by fermentable fibers [112], and there is a positive relationship between dietary protein intake and H2S production by the intestinal microbiota [113,114].
The protective role of H2S in bacteria was suggested more than six decades ago. Indeed, H2S produced by Desulfovibrio desulfuricans was demonstrated to be the diffusible factor responsible for protecting Pseudomonas aeruginosa from heavy metal toxicity (for instance, mercury) [115]. Similarly, H2S produced by Escherichia coli contributes to the protection of Staphilococcus aureus against mercuric chloride toxicity [116]. Much more recently, and in the context of the study of resistance to antibiotics, H2S has been shown to be a protective agent in bacteria such as Pseudomonas aeruginosa (found in human fecal samples [117]), Staphylococcus aureus (found in human intestine [118]), and Escherichia coli against the action of numerous antibiotics [119,120] (Figure 2). Sequestration of Fe2+ ions by H2S counteracts oxidative stress triggered by antibiotics in Escherichia coli [121]. Furthermore, H2S has been shown to be involved in the maintenance of redox homeostasis in bacteria and to protect bacteria against the oxidative stress triggered by the antibiotic ampicillin [122]. Also of major importance, cystathionine-Ɣ-lyase has been identified as the primary enzymatic activity that generates H2S in Staphylococcus aureus and Pseudomonas aeruginosa, and inhibition of this activity potentiates the efficiency of bactericidal antibiotics against both pathogens in in vitro and in vivo models of infection [123].
However, H2S is apparently not a bacterial metabolite that limits the efficiency of antibiotics against all pathogenic bacteria. Indeed, for instance, the pathogenic bacteria Acinetobacter baumannii, which is found in the intestinal tract [124] and does not produce H2S, is sensitive to exogenous H2S since, in these bacteria, this compound reinforces the effects of several classes of antibiotics [125]. Thus, the H2S-mediated protection or sensitization of intestinal bacteria to the bactericidal effects of several antibiotics apparently depends on the bacterial species examined.
Lastly, in another context, the implication of H2S in the resistance to infection by pathogenic bacteria has been recently suggested [126]. Indeed, in this latter study, the production of sulfide from taurine appears to be involved in the enhanced capacity of commensal bacteria to counteract pathogen infection.

3.6. Polyamines

Polyamines are small aliphatic amines that are produced by the intestinal microbiota from amino acids. The precursors for putrescine, spermidine, and spermine synthesis in bacteria are ornithine, arginine, and methionine, respectively, while agmatine is produced from arginine and cadaverine is produced from lysine [127]. The polyamines putrescine, spermidine, and spermine play a central role in bacterial growth [128] (Figure 2).
Regarding cadaverine, experimental works suggest that this polyamine plays a major role in the pathogenesis of Shigella infections [129]. In addition, cadaverine can prevent the escape of Shigella flexneri from the phagolysosome, and such an effect likely represents a way to connect the control of bacterial dissemination and neutrophil transepithelial signaling [130].
The polyamines agmatine and spermidine have been shown to be implicated in biofilm formation [131], while spermine inhibits Vibrio cholerae biofilm formation [132]. Spermidine reinforces the production of the bacterial genotoxin colibactin [133]. Lastly, one study reported that putrescine, cadaverine, spermidine, and spermine can modulate Vibrio cholerae virulence properties [134].

3.7. Gamma-Amino Butyric Acid, Norepinephrine, and Serotonin

Several bacterial species present in the mammalian gut, including the human gut, produce metabolites such as gamma-amino butyric acid, norepinephrine, and serotonin from different amino acids. These bacterial metabolites are well known to also be produced by the host with neurotransmitter functions [24]. There are emerging data that indicate that these metabolites are involved in the adaptation of bacteria to changes in their environment as well as in the modulation of bacterial physiology and growth. This is an exciting area of research, as it strongly suggests a very long evolutionary history for the functions of these compounds in the living world.
Gamma-amino butyric acid (GABA) is produced from glutamate by some intestinal bacteria, including strains of Lactobacillus and Bifidobacterium, as well as Bacteroides spp. [135,136,137]. GABA has been shown to be one component involved in bacterial acid tolerance in the context of changing luminal pH through the maintenance of intracellular pH [136,138,139] (Figure 2).
Regarding norepinephrine (also called noradrenaline), this bacterial metabolite is produced by the intestinal microbiota from tyrosine by several bacterial species, such as Bacillus subtilis, Escherichia coli, and Proteus vulgaris [140]. This bacterial metabolite affects the growth, either positively or negatively, depending on the bacterial species considered, of some anaerobic bacteria such as Fusobacterium nucleatum, Prevotella spp., Klebsiella pneumoniae, Pseudomonas aeruginosa, Enterobacter clocae, Shigella sonnei, and Staphylococcus aureus. Norepinephrine increases the virulence of several anaerobic bacteria, such as Clostridium perfringens [141,142,143].
Serotonin (5-hydroxy trypyamine) is produced from tryptophan by a vast number of bacterial species, among which Escherichia coli, Bacteroides, Streptococcus, Bifidobacterium, Lactococcus, Lactobacillus, and Propionibacterium [140]. Serotonin has been shown to promote the colonization of Turicibacter sanguinis in the human gut [144].

3.8. 4-Hydroxyphenylacetate

The bacterial metabolite 4-hydroxyphenylacetate (HPA) is a metabolic intermediary that is produced from tyrosine by phenol and p-cresol producers [89]. This compound inhibits the growth of the foodborne pathogen Listeria monocytogenes (Figure 2), an effect associated with alteration of the morphology of the bacteria and decreased expression of several virulence genes [145].

4. Conclusions and Perspectives

As presented in the present review, the results from experimental studies clearly indicate that numerous metabolites produced from amino acids of dietary and endogenous origin by the intestinal microbiota, which are found in stools and in the luminal fluid of the large intestine, are biologically active on microbes. The effects recorded are related to communication between microbes and to the physiology, metabolism, and growth of intestinal microorganisms. From the recorded effects of these bacterial metabolites on intestinal microbe biology, it is tempting to propose that the production of these compounds is devoted in the very first place to the dialogue between microorganisms coexisting in the mammalian intestinal colorectal luminal fluid.
However, the data obtained from experimental works are often characterized by several limitations that need to be taken into consideration for correct interpretation and future work. Firstly, the concentrations of the bacterial metabolites tested are not necessarily within the range of concentrations that are present in the vicinity of microbes in the colorectal fluid. Since this parameter is experimentally difficult to measure, the bacterial concentrations used in experiments often refer to the concentrations measured in feces, which represents an approximation of the concentrations within the rectal fluid but is likely different from those in the different segments of the colon [24]. Secondly, the intestinal luminal fluid contains a complex mixture of bacterial metabolites that can exert presumably additive, synergistic, or opposite effects on the intestinal microbial population. In fact, in most experimental works, the bacterial metabolites are tested individually, making it difficult to extrapolate from experimental works to “real-life situations”.
With these reservations in mind, the emerging experimental data indicate that intestinal bacteria produce several metabolites from amino acids that are involved in signaling between them and other microbial species (either commensal or pathogenic), as well as in the physiology and metabolism of these microorganisms, thus regulating their respective growth and biological activities. These data encourage clinical work in volunteers to test in different situations the relevance of dietary (or pharmacological) interventions for eventually controlling the colorectal microbial population and its metabolic activities in a way that would be beneficial for the host intestinal health from a preventive or curative perspective [24].

Funding

This research received no specific funding.

Informed Consent Statement

Not applicable.

Acknowledgments

The author wish to thank Université Paris Saclay, INRAe, and AgroParisTech for their constant support.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Bandyopadhyay, S.; Kashyap, S.; Calvez, J.; Devi, S.; Azzout-Marniche, D.; Tomé, D.; Kurpad, A.V.; Gaudichon, C. Evaluation of protein quality in humans and insights on stable isotope approaches to measure digestibility—A Review. Adv. Nutr. 2022, 13, 1131–1143. [Google Scholar] [CrossRef]
  2. Lister, N.; Sykes, A.P.; Bailey, P.D.; Boyd, C.A.; Bronk, J.R. Dipeptide transport and hydrolysis in isolated loops of rat small intestine: Effects of stereospecificity. J. Physiol. 1995, 484, 173–182. [Google Scholar] [CrossRef]
  3. Buddington, R.K.; Elnif, J.; Puchal-Gardiner, A.A.; Sangild, P.T. Intestinal apical amino acid absorption during development of the pig. Am. J. Physiol. 2001, 280, R241–R247. [Google Scholar] [CrossRef]
  4. Broër, S.; Fairweather, S.J. Amino acid transport across the mammalian intestine. Compr. Physiol. 2018, 9, 343–373. [Google Scholar]
  5. Gibson, J.A.; Sladen, G.E.; Dawson, A.M. Protein absorption and ammonia production: The effects of dietary protein and removal of the colon. Br. J. Nutr. 1976, 35, 61–65. [Google Scholar] [CrossRef]
  6. Kramer, P. The effect of varying sodium loads on the ileal excreta of human ileostomized subjects. J. Clin. Investig. 1966, 45, 1710–1718. [Google Scholar] [CrossRef]
  7. Smiddy, F.G.; Gregory, S.D.; Smith, I.B.; Goligher, J.C. Faecal loss of fluid, electrolytes, and nitrogen in colitis before and after ileostomy. Lancet 1960, 1, 14–19. [Google Scholar] [CrossRef]
  8. Chacko, A.; Cummings, J.H. Nitrogen losses from the human small bowel: Obligatory losses and the effect of physical form of food. Gut 1988, 29, 809–815. [Google Scholar] [CrossRef]
  9. Gaudichon, C.; Bos, C.; Morens, C.; Petzke, K.J.; Mariotti, F.; Everwand, J.; Benamouzig, R.; Daré, S.; Tomé, D.; Metges, C.C. Ileal losses of nitrogen and amino acids in humans and their importance to the assessment of amino acid requirements. Gastroenterology 2002, 123, 50–59. [Google Scholar] [CrossRef]
  10. Webb, K.E., Jr.; Matthews, J.C.; DiRienzo, D.B. Peptide absorption: A review of current concepts and future perspectives. J. Anim. Sci. 1992, 70, 3248–3257. [Google Scholar] [CrossRef]
  11. Webb, K.E., Jr.; DiRienzo, D.B.; Matthews, J.C. Recent developments in gastrointestinal absorption and tissue utilization of peptides: A review. J. Dairy Sci. 1993, 76, 351–361. [Google Scholar] [CrossRef]
  12. Dave, L.A.; Montoya, C.A.; Rutherfurd, S.M.; Moughan, P.J. Gastrointestinal endogenous proteins as a source of bioactive peptides--an in silico study. PLoS ONE 2014, 9, e98922. [Google Scholar] [CrossRef]
  13. Mariotti, F.; Tomé, D.; Mirand, P.P. Converting nitrogen into protein--beyond 6.25 and Jones’ factors. Crit. Rev. Food. Sci. Nutr. 2008, 48, 177–184. [Google Scholar] [CrossRef]
  14. Dubuisson, C.; Lioret, S.; Touvier, M.; Dufour, A.; Calamassi-Tran, G.; Volatier, J.L.; Lafay, L. Trends in food and nutritional intakes of French adults from 1999 to 2007: Results from the INCA surveys. Br. J. Nutr. 2010, 103, 1035–1048. [Google Scholar] [CrossRef]
  15. Pasiakos, S.M.; Agarwal, S.; Lieberman, H.R.; Fulgoni, V.L., 3rd. Sources and amounts of animal, dairy, and plant protein intake of US adults in 2007–2010. Nutrients 2015, 7, 7058–7069. [Google Scholar] [CrossRef]
  16. Kashyap, S.; Shivakumar, N.; Varkey, A.; Duraisamy, R.; Thomas, T.; Preston, T.; Devi, S.; Kurpad, A.V. Ileal digestibility of intrinsically labeled hen’s egg and meat protein determined with the dual stable isotope tracer method in Indian adults. Am. J. Clin. Nutr. 2018, 108, 980–987. [Google Scholar] [CrossRef]
  17. Bharucha, A.E.; Anderson, B.; Bouchoucha, M. More movement with evaluating colonic transit in humans. Neurogastroenterol. Motil. 2019, 31, e13541. [Google Scholar] [CrossRef]
  18. Rose, C.; Parker, A.; Jefferson, B.; Cartmell, E. The characterization of feces and urine: A review of the literature to inform advanced treatment technology. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1827–1879. [Google Scholar] [CrossRef]
  19. Raymann, K.; Moeller, A.H.; Goodman, A.L.; Ochman, H. Unexplored archaeal diversity in the great ape Gut Microbiome. mSphere 2017, 2, e00026-17. [Google Scholar] [CrossRef]
  20. Shkoporov, A.N.; Clooney, A.G.; Sutton, T.D.S.; Ryan, F.J.; Daly, K.M.; Nolan, J.A.; McDonnell, S.A.; Khokhlova, E.V.; Draper, L.A.; Forde, A.; et al. The human gut virome Is highly diverse, stable, and individual specific. Cell Host Microbe 2019, 26, 527–541.e5. [Google Scholar] [CrossRef]
  21. Carding, S.R.; Davis, N.; Hoyles, L. Review article: The human intestine virome in health and disease. Aliment. Pharmacol. Ther. 2017, 46, 800–815. [Google Scholar] [CrossRef] [PubMed]
  22. Paterson, M.J.; Oh, S.; Underhill, D.M. Host-microbe interactions: Commensal fungi in the gut. Curr. Opin. Microbiol. 2017, 40, 131–137. [Google Scholar] [CrossRef] [PubMed]
  23. Burgess, S.L.; Gilchrist, C.A.; Lynn, T.C.; Petri, W.A., Jr. Parasitic protozoan and interactions with the host intestinal microbiota. Infect. Immunol. 2017, 85, e00101-17. [Google Scholar] [CrossRef] [PubMed]
  24. Blachier, F. Metabolism of Alimentary Compounds by the Intestinal Microbiota and Health; Springer: Wien, Austria, 2023. [Google Scholar]
  25. Dodd, D.; Spitzer, M.H.; Van Treuren, W.; Merrill, B.D.; Hryckowian, A.J.; Higginbottom, S.K.; Le, A.; Cowan, T.M.; Nolan, G.P.; Fischbach, M.A.; et al. A gut bacterial pathway metabolizes aromatic amino acids into nine circulating metabolites. Nature 2017, 551, 648–652. [Google Scholar] [CrossRef]
  26. Kaman, W.E.; Hays, J.P.; Endtz, H.P.; Bikker, F.J. Bacterial proteases: Targets for diagnostics and therapy. Eur. J. Clin.Microbiol. Infect. Dis. 2014, 33, 1081–1087. [Google Scholar] [CrossRef]
  27. Cristofori, F.; Francavilla, R.; Capobianco, D.; Dargenio, V.N.; Filardo, S.; Mastromarino, P. Bacterial-based strategies to hydrolyze gluten peptides and protect intestinal mucosa. Front. Immunol. 2020, 11, 567801. [Google Scholar] [CrossRef]
  28. Macfarlane, G.T.; Macfarlane, S. Bacteria, colonic fermentation, and gastrointestinal health. J. AOAC Int. 2012, 95, 50–60. [Google Scholar] [CrossRef]
  29. Macfarlane, G.T.; Cummings, J.H. The Colonic flora, fermentation, and large bowel digestive function. In The Large Intestine; Phillips, S.F., Pemberton, J.H., Shorter, R.G., Eds.; Raven Press: New York, NY, USA, 1991; pp. 51–92. [Google Scholar]
  30. Pessione, E. Lactic acid bacteria contribution to gut microbiota complexity: Lights and shadows. Front. Cell Infect. Microbiol. 2012, 2, 86. [Google Scholar] [CrossRef]
  31. Liu, M.; Bayjanov, J.R.; Renckens, B.; Nauta, A.; Siezen, R.J. The proteolytic system of lactic acid bacteria revisited: A genomic comparison. BMC Genom. 2010, 11, 36. [Google Scholar] [CrossRef]
  32. Steiner, H.Y.; Naider, F.; Becker, J.M. The PTR family: A new group of peptide transporters. Mol. Microbiol. 1995, 16, 825–834. [Google Scholar] [CrossRef]
  33. Davies, J.S.; Currie, M.J.; Wright, J.D.; Newton-Vesty, M.C.; North, R.A.; Mace, P.D.; Allison, J.R.; Dobson, R.C.J. Selective nutrient transport in bacteria: Multicomponent transporter systems reign supreme. Front. Mol. Biosci. 2021, 8, 699222. [Google Scholar] [CrossRef]
  34. Garai, P.; Chandra, K.; Chakravortty, D. Bacterial peptide transporters: Messengers of nutrition to virulence. Virulence 2017, 8, 297–309. [Google Scholar] [CrossRef]
  35. Eggeling, L.; Sahm, H. New ubiquitous translocators: Amino acid export by Corynebacterium glutamicum and Escherichia coli. Arch. Microbiol. 2003, 180, 155–160. [Google Scholar]
  36. Katsube, S.; Ando, T.; Yoneyama, H. L-Alanine exporter, AlaE, of Escherichia coli functions as a safety valve to enhance survival under feast conditions. Int. J. Mol. Sci. 2019, 20, 4942. [Google Scholar] [CrossRef]
  37. James, P.S.; Smith, M.V. Methionine transport by pig colonic mucosa measured during early post-natal development. J. Physiol. 1976, 262, 151–168. [Google Scholar] [CrossRef]
  38. Sepulveda, F.V.; Smith, M.W. Different mechanisms for neutral amino acid uptake by new-born colon. J. Physiol. 1979, 286, 479–490. [Google Scholar] [CrossRef]
  39. van der Wielen, N.; Moughan, P.J.; Mensink, M. Amino acid absorption in the large intestine of human and porcine models. J. Nutr. 2017, 147, 1493–1498. [Google Scholar] [CrossRef]
  40. Wuensch, T.; Schulz, S.; Ullrich, S.; Lill, N.; Stelzl, T.; Rubio-Aliaga, I.; Loh, G.; Chamaillard, M.; Haller, D.; Daniel, H. The peptide transporter PEPT1 is expressed in distal colon in rodents and humans and contributes to water absorption. Am. J. Physiol. 2013, 305, G66–G73. [Google Scholar] [CrossRef]
  41. Stephen, A.M.; Cummings, J.H. The microbial contribution to human fecal mass. J. Med. Microbiol. 1980, 13, 45–56. [Google Scholar] [CrossRef]
  42. Portune, K.J.; Beaumont, M.; Davila, A.M.; Tomé, D.; Blachier, F.; Sanz, Y. Gut microbiota role in dietary protein metabolism and health-related outcomes: The two sides of the coin. Trends Food. Sci. Technol. 2016, 57, 213–232. [Google Scholar] [CrossRef]
  43. Smith, A.B.; Jenior, M.L.; Keenan, O.; Hart, J.L.; Specker, J.; Abbas, A.; Rangel, P.C.; Di, C.; Green, J.; Bustin, K.A.; et al. Enterococci enhance Clostridioides difficile pathogenesis. Nature 2022, 611, 780–786. [Google Scholar] [CrossRef] [PubMed]
  44. Shimizu, T.; Ohtani, K.; Hirakawa, H.; Ohshima, K.; Yamashita, A.; Shiba, T.; Ogasawara, N.; Hattori, M.; Kuhara, S.; Hayashi, S. Complete genome sequence of Clostridium perfringens, an anaerobic flesh-eater. Proc. Natl. Acad. Sci. USA 2002, 99, 996–1001. [Google Scholar] [CrossRef] [PubMed]
  45. Pridmore, R.D.; Berger, B.; Desiere, F.; Vilanova, D.; Barretto, C.; Pittet, A.C.; Zwallen, M.C.; Rouvet, M.; Altermann, E.; Barrangou, R.; et al. The genome sequence of the probiotic intestinal bacterium Lactobacillus johnsonii NCC 533. Proc. Natl. Acad. Sci. USA 2004, 101, 2512–2517. [Google Scholar] [CrossRef] [PubMed]
  46. Yu, X.J.; Walker, D.H.; Liu, Y.; Zhang, L. Amino acid biosynthesis deficiency in bacteria associated with human and animal hosts. Infect. Genet. Evol. 2009, 9, 514–517. [Google Scholar] [CrossRef]
  47. Nolling, J.; Breton, G.; Omeichenko, M.V.; Marakova, K.S.; Zeng, Q.; Gibson, R.; Lee, H.M.; Dubois, D.; Qiu, D.; Hitti, J.; et al. Genome sequence and comparative analysis of the solvent-producing bacterium Clostridium acetobutylicum. J. Bacteriol. 2001, 183, 4823–4838. [Google Scholar] [CrossRef]
  48. Bolotin, A.; Wincker, P.; Mauger, S.; Jaillon, O.; Malarme, K.; Weissenbach, J.; Ehrlich, S.D.; Sorokin, A. The complete genome sequence of the lactic acid bacterium Lactococcus lactis spp. Lactis IL1403. Genome. Res. 2011, 11, 731–753. [Google Scholar] [CrossRef]
  49. Godon, J.J.; Delorme, C.; Bardowski, J.; Chopin, M.C.; Ehrlich, S.D.; Renault, P. Gene inactivation in Lactococcus lactis: Branched-chain amino acid biosynthesis. J. Bacteriol. 1993, 175, 4383–4390. [Google Scholar] [CrossRef]
  50. Dai, Z.L.; Wu, G.; Zhu, W.Y. Amino acid metabolism in intestinal bacteria: Links between gut ecology and host health. Front. Biosci. 2011, 16, 1768–1786. [Google Scholar] [CrossRef]
  51. Barker, H.A. Amino acid degradation by anaerobic bacteria. Annu. Rev. Biochem. 1981, 50, 23–40. [Google Scholar] [CrossRef]
  52. Riggottier-Gois, L. Dysbiosis in inflammatory bowel diseases: The oxygen hypothesis. ISME J. 2013, 7, 1256–1261. [Google Scholar] [CrossRef]
  53. Kim, R.; Attayek, P.J.; Wang, Y.; Furtado, K.L.; Tamayo, R.; Sims, C.E.; Allbritton, N.L. An in vitro intestinal platform with a self-sustaining oxygen gradient to study the human gut/microbiome interface. Biofabrication 2019, 12, 015006. [Google Scholar] [CrossRef] [PubMed]
  54. Smith, E.A.; Macfarlane, G.T. Enumeration of human colonic bacteria producing phenolic and indolic compounds: Effects of pH, carbohydrate availability and retention time on dissimilatory aromatic amino acid metabolism. J. Appl. Bacteriol. 1996, 81, 288–302. [Google Scholar] [CrossRef] [PubMed]
  55. Davila, A.M.; Blachier, F.; Gotteland, M.; Andriamihaja, M.; Benetti, P.H.; Sanz, Y.; Tomé, D. Intestinal luminal nitrogen metabolism: Role of the gut microbiota and consequences for the host. Pharmacol. Res. 2013, 68, 95–107. [Google Scholar] [CrossRef] [PubMed]
  56. Kim, J.; Hetzel, M.; Boiangiu, C.D.; Buckel, W. Dehydration of (R)-2-hydroxyacyl-CoA to enoyl-CoA in the fermentation of alpha-amino acids by anaerobic bacteria. FEMS Microbiol. Rev. 2004, 28, 455–468. [Google Scholar] [CrossRef] [PubMed]
  57. Fischbach, M.A.; Sonnenburg, J.L. Eating for two: How metabolism establishes interspecies interactions in the gut. Cell Host Microbe. 2011, 10, 336–347. [Google Scholar] [CrossRef] [PubMed]
  58. Mei, R.; Nobu, M.K.; Liu, W.T. Identifying anaerobic amino acids degraders through the comparison of short-term and long- term enrichments. Environ. Microbiol. Rep. 2020, 12, 173–184. [Google Scholar] [CrossRef]
  59. Fonknechten, N.; Chaussonnerie, S.; Tricot, S.; Lajus, A.; Andreesen, J.R.; Perchat, N.; Pelletier, E.; Gouyvenoux, M.; Barbe, V.; Salanoubat, M.; et al. Clostridium stricklandii, a specialist in amino acid degradation: Revisiting its metabolism through its genome sequence. BMC Genom. 2010, 11, 555. [Google Scholar] [CrossRef]
  60. Birkett, A.; Muir, J.; Phillips, J.; Jones, G.; O’Dea, K. Resistant starch lowers fecal concentrations of ammonia and phenol in humans. Am. J. Clin. Nutr. 1996, 63, 766–772. [Google Scholar] [CrossRef]
  61. Geboes, K.P.; De Hertogh, G.; De Preter, V.; Luypaerts, A.; Bammens, B.; Evenepoel, P.; Ghoss, Y.; Geboes, K.; Rutgeerts, P.; Verbeke, K. The influence of inulin on the absorption of nitrogen and the production of metabolites of protein fermentation in the colon. Br. J. Nutr. 2006, 96, 1078–1086. [Google Scholar] [CrossRef]
  62. Windey, K.; De Preter, V.; Huys, G.; Broekaert, W.F.; Delcour, J.A.; Louat, T.; Herman, J.; Verbeke, K. Wheat bran extract alters colonic fermentation and microbial composition but does not affect faecal water toxicity: A randomized controlled trial in healthy subjects. Br. J. Nutr. 2015, 113, 225–238. [Google Scholar] [CrossRef]
  63. Cummings, J.H.; Macfarlane, G.T. Role of intestinal bacteria in nutrient metabolism. J. Parenter. Enter. Nutr. 1997, 21, 357–365. [Google Scholar] [CrossRef] [PubMed]
  64. Pavao, A.; Graham, M.; Arrieta-Ortiz, M.L.; Immanuel, S.R.C.; Baliga, N.S.; Bry, L. Reconsidering the in vivo functions of Clostridial Stickland amino acid fermentations. Anaerobe 2022, 76, 102600. [Google Scholar] [CrossRef] [PubMed]
  65. Macfarlane, G.T.; Gibson, G.R.; Cummings, J.H. Comparison of fermentation reactions in different regions of the human colon. J. Appl. Bacteriol. 1992, 72, 57–64. [Google Scholar] [PubMed]
  66. Zeng, X.; Xing, X.; Gupta, M.; Keber, F.C.; Lopez, J.G.; Lee, Y.J.; Roichman, A.; Wang, L.; Neinast, M.D.; Donia, M.S.; et al. Gut bacterial nutrient preferences quantified in vivo. Cell 2022, 185, 3441–3456.e19. [Google Scholar] [CrossRef]
  67. Zheng, X.; Wang, S.; Jia, W. Calorie restriction and its impact on gut microbial composition and global metabolism. Front. Med. 2018, 12, 634–644. [Google Scholar] [CrossRef]
  68. Schmidt, N.S.; Lorentz, A. Dietary restrictions modulate the gut microbiota: Implications for health and disease. Nutr. Res. 2021, 89, 10–22. [Google Scholar] [CrossRef]
  69. Flint, H.J. The impact of nutrition on the human microbiome. Nutr. Rev. 2012, 70, S10–S13. [Google Scholar] [CrossRef]
  70. Sbierski-Kind, J.; Grenkowitz, S.; Schlickeiser, S.; Sandforth, A.; Friedrich, M.; Kunkel, D.; Glauben, R.; Brachs, S.; Mai, K.; Thürmer, A.; et al. Effects of caloric restriction on the gut microbiome are linked with immune senescence. Microbiome 2022, 10, 57. [Google Scholar] [CrossRef]
  71. Kable, M.E.; Chin, E.L.; Storms, D.; Lemay, D.G.; Stephensen, C.B. Tree-based analysis of dietary diversity captures associations between fiber intake and gut microbiota composition in a healthy US adult cohort. J. Nutr. 2022, 152, 779–788. [Google Scholar] [CrossRef]
  72. Tanes, C.; Bittinger, K.; Gao, Y.; Friedman, E.S.; Nessel, L.; Paladhi, U.R.; Chau, L.; Panfen, E.; Fischbach, M.A.; Braun, J.; et al. Role of dietary fiber in the recovery of the human gut microbiome and its metabolome. Cell Host Microbe 2021, 29, 394–407.e5. [Google Scholar] [CrossRef]
  73. Beaumont, M.; Portune, K.J.; Steuer, N.; Lan, A.; Cerrudo, V.; Audebert, M.; Dumont, F.; Mancano, G.; Khodorova, N.; Andriamihaja, M.; et al. Quantity and source of dietary protein influence metabolite production by gut microbiota and rectal mucosa gene expression: A randomized, parallel, double-blind trial in overweight humans. Am. J. Clin. Nutr. 2017, 106, 1005–1019. [Google Scholar] [CrossRef] [PubMed]
  74. Windey, K.; De Preter, V.; Louat, T.; Schuit, F.; Herman, J.; Vansant, G.; Verbeke, K. Modulation of protein fermentation does not affect fecal water toxicity: A randomized cross-over study in healthy subjects. PLoS ONE 2012, 7, e52387. [Google Scholar] [CrossRef] [PubMed]
  75. Bel Lassen, P.; Belda, E.; Prifti, E.; Dao, M.C.; Specque, F.; Henegar, C.; Rinaldi, L.; Wang, X.; Kennedy, S.P.; Zucker, J.D.; et al. Protein supplementation during an energy-restricted diet induces visceral fat loss and gut microbiota amino acid metabolism activation: A randomized trial. Sci. Rep. 2021, 11, 15620. [Google Scholar] [CrossRef] [PubMed]
  76. Sridharan, G.V.; Choi, K.; Klemashevich, C.; Wu, C.; Prabakaran, D.; Pan, L.B.; Steimeyer, S.; Mueller, C.; Yousofshani, M.; Alaniz, R.C.; et al. Prediction and quantification of bioactive microbiota metabolites in the mouse. Nat. Commun. 2014, 5, 5492. [Google Scholar] [CrossRef] [PubMed]
  77. Macfarlane, G.T.; Cummings, J.H.; Macfarlane, S.; Gibson, G.R. Influence of retention time on degradation of pancreatic enzymes by human colonic bacteria grown in a 3-stage continuous system. J. Appl. Bacteriol. 1989, 67, 520–527. [Google Scholar] [CrossRef]
  78. Roager, H.M.; Hansen, L.B.S.; Bahl, M.I.; Frandsen, H.L.; Carvalho, V.; Gobel, R.J.; Dalgaard, M.D.; Plichta, D.R.; Sparholt, M.H.; Vestergaard, H.; et al. Colonic transit time is related to bacterial metabolism and mucosal turnover in the gut. Nat. Microbiol. 2016, 1, 16093. [Google Scholar] [CrossRef]
  79. Liu, X.; Blouin, J.M.; Santacruz, A.; Lan, A.; Andriamihaja, M.; Wilkanowicz, S.; Benetti, P.H.; Tomé, D.; Sanz, Y.; Blachier, F.; et al. High-protein diet modifies colonic microbiota and luminal environment but not colonocyte metabolism: The increased luminal bulk connection. Am. J. Physiol. 2014, 307, G459–G470. [Google Scholar] [CrossRef]
  80. Endo, A.; Nakamura, S.; Konishi, K.; Nakagawa, J.; Tochio, T. Variations in prebiotic oligosaccharide fermentation by intestinal lactic acid bacteria. Int. J. Food Sci. Nutr. 2016, 67, 125–132. [Google Scholar] [CrossRef]
  81. Weber, F.L., Jr. Effects of lactulose on nitrogen metabolism. Scand. J. Gastroenterol. 1997, 222, 83–87. [Google Scholar] [CrossRef]
  82. Liong, M.T.; Shah, N.P. Production of organic acids from fermentation of mannitol, fructooligosaccharide and inulin by a cholesterol removing Lactobacillus acidophilus strain. J. Appl. Microbiol. 2005, 99, 783–793. [Google Scholar] [CrossRef]
  83. Sheridan, P.O.; Louis, P.; Tsompanidou, E.; Shaw, S.; Harmsen, H.J.; Duncan, S.H.; Flint, H.J.; Walker, A.W. Distribution, organization and expression of genes concerned with anaerobic lactate utilization in human intestinal bacteria. Microb. Genom. 2022, 8, 000739. [Google Scholar] [CrossRef] [PubMed]
  84. Koestler, B.J.; Fisher, C.R.; Payne, S.M. Formate promotes Shigella intercellular spread and virulence gene expression. mBio 2018, 9, e01777-18. [Google Scholar] [CrossRef] [PubMed]
  85. Shaulov, Y.; Shimokawa, C.; Trebicz-Geffren, M.; Nagaraja, S.; Methling, K.; Lalk, M.; Weiss-Cerem, L.; Lamm, A.T.; Hisaeda, H.; Ankri, S. Escherichia coli mediated resistance of Entamoeba histolytica to oxidative stress is triggered by oxaloacetate. PLoS Pathog. 2018, 14, e1007295. [Google Scholar] [CrossRef] [PubMed]
  86. Ferreyra, J.A.; Wu, K.J.; Hryckowian, A.J.; Bouley, D.M.; Weiner, B.C.; Sonnenburg, J.L. Gut microbiota-produced succinate promotes C. difficile infection after antibiotic treatment or motility disturbance. Cell Host Microbe 2014, 16, 770–777. [Google Scholar] [CrossRef] [PubMed]
  87. Hagihara, M.; Ariyoshi, T.; Kuroki, Y.; Eguchi, S.; Higashi, S.; Mori, T.; Nonogaki, T.; Iwasaki, K.; Yamashita, M.; Asai, N.; et al. Clostridium butyricum enhances colonization resistance against Clostridioides difficile by metabolic and immune modulation. Sci. Rep. 2021, 11, 15007. [Google Scholar] [CrossRef]
  88. Bone, E.; Tamm, A.; Hill, M. The production of urinary phenols by gut bacteria and their possible role in the causation of large bowel cancer. Am. J. Clin. Nutr. 1976, 29, 1448–1454. [Google Scholar] [CrossRef]
  89. Saito, Y.; Sato, T.; Nomoto, K.; Tsuji, H. Identification of phenol- and p-cresol-producing intestinal bacteria by using media supplemented with tyrosine and its metabolites. FEMS Microbiol. Ecol. 2018, 94, fiy125. [Google Scholar] [CrossRef]
  90. Geypens, B.; Claus, D.; Evenepoel, P.; Hiele, M.; Maes, B.; Peeters, M.; Rutgeers, P.; Ghoos, Y. Influence of dietary protein supplements on the formation of bacterial metabolites in the colon. Gut 1997, 41, 70–76. [Google Scholar] [CrossRef]
  91. Passmore, I.J.; Letertre, M.P.M.; Preston, M.D.; Bianconi, I.; Harrison, M.A.; Nasher, F.; Kaur, H.; Hong, H.A.; Baines, H.D.; Cutting, S.M.; et al. Para-cresol production by Clostridium difficile affects microbial diversity and membrane integrity of Gram-negative bacteria. PLoS Pathog. 2018, 14, e1007191. [Google Scholar] [CrossRef]
  92. Harrison, M.A.; Strahl, H.; Dawson, L.F. Regulation of para-cresol production in Clostridioides difficile. Curr. Opin. Microbiol. 2022, 65, 131–137. [Google Scholar] [CrossRef]
  93. Abt, M.C.; McKenney, P.T.; Pamer, E.G. Clostridium difficile colitis: Pathogenesis and host defence. Nat. Rev. Microbiol. 2016, 14, 609–620. [Google Scholar] [CrossRef] [PubMed]
  94. Keszthelyi, D.; Troost, F.J.; Masclee, A.A. Understanding the role of tryptophan and serotonin metabolism in gastrointestinal function. Neurogastroenterol. Motil. 2009, 21, 1239–1249. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, J.Y.; Wood, T.K.; Lee, J. Roles of indole as an interspecies and interkingdom signaling molecule. Trends Microbiol. 2015, 23, 707–718. [Google Scholar] [CrossRef] [PubMed]
  96. Roager, H.M.; Licht, T.R. Microbial tryptophan catabolites in health and disease. Nat. Commun. 2018, 9, 3294. [Google Scholar] [CrossRef]
  97. Rattanaphan, P.; Mittraparp-Arthorn, P.; Srinoun, K.; Vuddhakul, V.; Tansila, N. Indole signaling decreases biofilm formation and related virulence of Listeria monocytogenes. FEMS Microbiol. Lett. 2020, 367, fnaa116. [Google Scholar] [CrossRef]
  98. Lee, J.; Attila, C.; Cirillo, S.L.; Cirillo, J.D.; Wood, T.K. Indole and 7-hydroxyindole diminish Pseudomonas aeruginosa virulence. Microb. Biotechnol. 2009, 2, 75–90. [Google Scholar] [CrossRef]
  99. Nikaido, E.; Giraud, E.; Baucheron, S.; Yamasaki, S.; Wiedemann, A.; Okamoto, K.; Takagi, T.; Yamaguchi, A.; Cloeckaert, A.; Nishino, K. Effects of indole on drug resistance and virulence of Salmonella enterica serovar Typhimurium revealed by genome-wide analyses. Gut Pathog. 2012, 4, 5. [Google Scholar] [CrossRef]
  100. Oh, S.; Go, G.W.; Mylonakis, E.; Kim, Y. The bacterial signalling molecule indole attenuates the virulence of the fungal pathogen Candida albicans. J. Appl. Microbiol. 2012, 113, 622–628. [Google Scholar] [CrossRef]
  101. Nowak, A.; Libudzisz, Z. Influence of phenol, p-cresol and indole on growth and survival of intestinal lactic acid bacteria. Anaerobe 2006, 12, 80–84. [Google Scholar] [CrossRef]
  102. Ledala, N.; Malik, M.; Rezaul, K.; Paveglio, S.; Provatas, A.; Kiel, A.; Caimano, M.; Zhou, Y.; Lindgren, J.; Krasulova, K.; et al. Bacterial indole as a multifunctional regulator of Klebsella oxytoca complex enterotoxicity. mBio 2022, 13, e0375221. [Google Scholar] [CrossRef]
  103. Gorelik, O.; Rogad, A.; Holoidovsky, L.; Meijler, M.M.; Sal-Man, N. Indole intercepts the communication between enteropathogenic E. coli and Vibrio cholerae. Gut Microbes 2022, 14, 2138677. [Google Scholar] [CrossRef] [PubMed]
  104. Darkoh, C.; Plants-Paris, K.; Bishoff, D.; DuPont, H.L. Clostridium difficile modulates the gut microbiota by inducing the production of indole, an interkingdom signaling and antimicrobial molecule. mSystems 2019, 4, e00346-18. [Google Scholar] [CrossRef] [PubMed]
  105. Choi, S.H.; Kim, Y.; Oh, S.; Oh, S.; Chun, T.; Kim, T.; Kim, S.H. Inhibitory effect of skatole (3-methylindole) on enterohemorrhagic Escherichia coli O157:H7 ATCC 43894 biofilm formation mediated by elevated endogenous oxidative stress. Lett. Appl. Microbiol. 2014, 58, 454–461. [Google Scholar] [CrossRef] [PubMed]
  106. Probert, H.M.; Gibson, G.R. Bacterial biofilms in the human gastrointestinal tract. Curr. Issues Intest. Microbiol. 2002, 3, 23–27. [Google Scholar]
  107. Blachier, F.; Davila, A.M.; Mimoun, S.; Benetti, P.H.; Atanasiu, C.; Andriamihaja, M.; Benamouzig, R.; Bouillaud, F.; Tomé, D. Luminal sulfide and large intestine mucosa: Friend or foe? Amino Acids 2010, 39, 335–347. [Google Scholar] [CrossRef]
  108. Rowan, F.E.; Docherty, N.G.; Coffey, J.C.; O’Connell, P.R. Sulphate-reducing bacteria and hydrogen sulphide in the aetiology of ulcerative colitis. Br. J. Surg. 2009, 96, 151–158. [Google Scholar] [CrossRef]
  109. Laue, H.; Denger, K.; Cook, A.M. Taurine reduction in anaerobic respiration of Bilophila wadsworthia RZATAU. Appl. Environ. Microbiol. 1997, 63, 2016–2021. [Google Scholar] [CrossRef]
  110. Laue, H.; Friedrich, M.; Ruff, J.; Cook, A.M. Dissimilatory sulfite reductase (desulfoviridin) of the taurine-degrading, non-sulfate-reducing bacterium Bilophila wadsworthia RZATAU contains a fused DsrB-DsrD subunit. J. Bacteriol. 2001, 183, 1727–1733. [Google Scholar] [CrossRef]
  111. Carbonero, F.; Benefiel, A.C.; Alizadeh-Ghamsari, A.H.; Gaskins, R. Microbial pathways in colonic sulfur metabolism and links with health and disease. Front. Physiol. 2012, 3, 448. [Google Scholar] [CrossRef]
  112. Yao, C.K.; Rotbart, A.; Ou, J.Z.; Kalantar-Zadeh, K.; Muir, J.G.; Gibson, P.R. Modulation of colonic hydrogen sulfide production by diet and mesalazine utilizing a novel gas-profiling technology. Gut Microbes 2018, 9, 510–522. [Google Scholar] [CrossRef]
  113. Teigen, L.; Biruete, A.; Khoruts, A. Impact of diet on hydrogen sulfide production: Implications for gut health. Curr. Opin. Clin. Nutr. Metab. Care 2023, 26, 55–58. [Google Scholar] [CrossRef] [PubMed]
  114. Magee, E.A.; Richardson, C.J.; Hughes, R.; Cummings, J.H. Contribution of dietary protein to sulfide production in the large intestine: An in vitro and a controlled feeding study in humans. Am. J. Clin. Nutr. 2000, 72, 1488–1494. [Google Scholar] [CrossRef] [PubMed]
  115. Bachenheimer, A.G.; Bennett, E.O. The sensitivity of mixed population of bacteria to inhibitors. I. The mechanism by which desulfovibrio desulfuricans protects Ps. aeruginosa from the toxicity of mercurials. Antonie Van Leeuwenhoek 1961, 27, 180–188. [Google Scholar] [CrossRef]
  116. Stutzenberger, F.J.; Bennett, E.O. Sensitivity of mixed populations of Staphylococcus aureus and Escherichia coli to mercurials. Appl. Microbiol. 1965, 13, 570–574. [Google Scholar] [CrossRef] [PubMed]
  117. Ruiz-Roldan, L.; Bellés, A.; Bueno, J.; Azcona-Guttierrez, J.M.; Rojo-Bezares, B.; Torres, C.; Castillo, F.J.; Saenz, Y.; Seral, C. Pseudomonas aeruginosa Isolates from Spanish Children: Occurrence in Faecal Samples, Antimicrobial Resistance, Virulence, and Molecular Typing. Biomed. Res. Int. 2018, 2018, 8060178. [Google Scholar] [CrossRef] [PubMed]
  118. Piewngam, P.; Otto, M. Probiotics to prevent Staphylococcus aureus disease? Gut Microbes 2020, 11, 94–101. [Google Scholar] [CrossRef]
  119. Pal, V.K.; Bandyopadhyay, P.; Singh, A. Hydrogen sulfide in physiology and pathogenesis of bacteria and viruses. IUBMB Life 2018, 70, 393–410. [Google Scholar] [CrossRef]
  120. Shatalin, K.; Shatalina, E.; Mironov, A.; Nudler, E. H2S: A universal defense against antibiotics in bacteria. Science 2011, 334, 986–990. [Google Scholar] [CrossRef]
  121. Mironov, A.; Seregina, T.; Nagornykh, M.; Luhachack, L.G.; Korolkova, N.; Lopes, L.E.; Kotova, V.; Zavilgelsky, G.; Shakulov, R.; Shatalin, K.; et al. Mechanism of H2S-mediated protection against oxidative stress in Escherichia coli. Proc. Natl. Acad. Sci. USA 2017, 114, 6022–6027. [Google Scholar] [CrossRef]
  122. Shukla, P.; Khodade, V.S.; SharathChandra, M.; Chauhan, P.; Mishra, S.; Siddaramappa, S.; Pradeep, B.E.; Singh, A.; Chakrapani, H. “On demand” redox buffering by H2S contributes to antibiotic resistance revealed by a bacteria-specific H2S donor. Chem. Sci. 2017, 8, 4967–4972. [Google Scholar] [CrossRef]
  123. Shatalin, K.; Nuthanakanti, A.; Kaushik, A.; Shishov, D.; Peselis, A.; Shamovsky, I.; Pani, B.; Lechpammer, M.; Vasilyev, N.; Shatalina, E.; et al. Inhibitors of bacterial H2S biogenesis targeting antibiotic resistance and tolerance. Science 2021, 372, 1169–1175. [Google Scholar] [CrossRef]
  124. Ketter, P.M.; Yu, J.J.; Guentzel, M.N.; May, H.C.; Gupta, R.; Eppinger, M.; Klose, K.E.; Seshu, J.; Chambers, J.P.; Cap, A.P.; et al. Acinetobacter baumannii Gastrointestinal Colonization Is Facilitated by Secretory IgA Which Is Reductively Dissociated by Bacterial Thioredoxin A. mBio 2018, 9, e1298-18. [Google Scholar] [CrossRef] [PubMed]
  125. Ng, S.Y.; Ong, K.X.; Surendran, S.T.; Sinha, A.; Lai, J.J.H.; Chen, J.; Liang, J.; Tay, L.K.S.; Cui, L.; Loo, H.L.; et al. Hydrogen sulfide sensitizes Acinetobacter baumannii to killing by antibiotics. Front. Microbiol. 2020, 11, 1875. [Google Scholar] [CrossRef] [PubMed]
  126. Stacy, A.; Andrade-Oliveira, V.; McCulloch, J.A.; Hild, B.; Oh, J.H.; Perez-Chaparro, P.J.; Sim, C.K.; Lim, A.I.; Link, V.M.; Enamorado, M.; et al. Infection trains the host for microbiota-enhanced resistance to pathogens. Cell 2021, 184, 615–627.e17. [Google Scholar] [CrossRef] [PubMed]
  127. Shah, P.; Swiatlo, E. A multifaceted role for polyamines in bacterial pathogens. Mol. Microbiol. 2008, 68, 4–16. [Google Scholar] [CrossRef] [PubMed]
  128. Igarashi, K.; Kashiwagi, K. Modulation of cellular function by polyamines. Int. J. Biochem. Cell Biol. 2010, 42, 39–51. [Google Scholar] [CrossRef]
  129. Maurelli, A.T.; Fernandez, R.E.; Bloch, C.A.; Rode, C.K.; Fasano, A. “Black holes” and bacterial pathogenicity: A large genomic deletion that enhances the virulence of Shigella spp. and enteroinvasive Escherichia coli. Proc. Natl. Acad. Sci. USA 1998, 95, 3943–3948. [Google Scholar] [CrossRef]
  130. Fernandez, I.M.; Silva, M.; Schuch, R.; Walker, W.A.; Siber, A.M.; Maurelli, A.T.; McCormick, B.A. Cadaverine prevents the escape of Shigella flexneri from the phagolysosome: A connection between bacterial dissemination and neutrophil transepithelial signaling. J. Infect. Dis. 2001, 184, 743–753. [Google Scholar] [CrossRef]
  131. Burrell, M.; Hanfrey, C.C.; Murray, E.J.; Stanley-Wall, N.R.; Michael, A.J. Evolution and multiplicity of arginine decarboxylases in polyamine biosynthesis and essential role in Bacillus subtilis biofilm formation. J. Biol. Chem. 2010, 285, 39224–39238. [Google Scholar] [CrossRef]
  132. Sobe, R.C.; Bond, W.G.; Wotanis, C.K.; Zayner, J.P.; Burriss, M.A.; Fernandez, N.; Bruger, E.L.; Waters, C.M.; Neufeld, H.S.; Karatan, E. Spermine inhibits Vibrio cholerae biofilm formation through the NspS-MbaA polyamine signaling system. J. Biol. Chem. 2017, 292, 17025–17036. [Google Scholar] [CrossRef]
  133. Chagneau, C.V.; Garcie, C.; Bossuet-Greif, N.; Tronnet, S.; Brachmann, A.O.; Piel, J.; Nougayrède, J.P.; Martin, P.; Oswald, E. The polyamine spermidine modulates the production of the bacterial genotoxin colibactin. mSphere 2019, 4, e00414-19. [Google Scholar] [CrossRef] [PubMed]
  134. Goforth, J.B.; Walter, N.E.; Karatan, E. Effects of polyamines on Vibrio cholerae virulence properties. PLoS ONE 2013, 8, e60765. [Google Scholar] [CrossRef] [PubMed]
  135. Barrett, E.; Ross, R.P.; O’Toole, P.W.; Fitzgerald, G.F.; Stanton, C. γ-Aminobutyric acid production by culturable bacteria from the human intestine. J. Appl. Microbiol. 2012, 113, 411–417. [Google Scholar] [CrossRef] [PubMed]
  136. Otaru, N.; Ye, K.; Mujezinovic, D.; Berchtold, L.; Constancias, F.; Cornejo, F.A.; Krzystek, A.; de Wouters, T.; Braegger, C.; Lacroix, C.; et al. GABA Production by human intestinal Bacteroides spp.: Prevalence, regulation, and role in acid stress tolerance. Front. Microbiol. 2021, 12, 656895. [Google Scholar] [CrossRef]
  137. Strandwitz, P.; Kim, K.H.; Terekhova, D.; Liu, J.K.; Sharma, A.; Levering, J.; McDonald, D.; Dietrich, D.; Ramadhar, T.R.; Lekbua, A.; et al. GABA-modulating bacteria of the human gut microbiota. Nat. Microbiol. 2019, 4, 396–403. [Google Scholar] [CrossRef] [PubMed]
  138. Feehily, C.; Karatzas, K.A.G. Role of glutamate metabolism in bacterial responses towards acid and other stresses. J. Appl. Microbiol. 2013, 114, 11–24. [Google Scholar] [CrossRef]
  139. Cotter, P.D.; Hill, C. Surviving the acid test: Responses of gram-positive bacteria to low pH. Microbiol. Mol. Biol. Rev. 2003, 67, 429–453. [Google Scholar] [CrossRef]
  140. Dicks, L.M.T. Gut Bacteria and Neurotransmitters. Microorganisms 2022, 10, 1838. [Google Scholar] [CrossRef]
  141. Boyanova, L. Stress hormone epinephrine (adrenaline) and norepinephrine (noradrenaline) effects on the anaerobic bacteria. Anaerobe 2017, 44, 13–19. [Google Scholar] [CrossRef]
  142. Lustri, B.C.; Sperandio, V.; Moreira, C.G. Bacterial chat: Intestinal metabolites and signals in host-microbiota-pathogen interactions. Infect. Immun. 2017, 85, e00476-17. [Google Scholar] [CrossRef]
  143. O’Donnell, P.M.; Aviles, H.; Lyte, M.; Sonnenfeld, G. Enhancement of in vitro growth of pathogenic bacteria by norepinephrine: Importance of inoculum density and role of transferrin. Appl. Environ. Microbiol. 2006, 72, 5097–5099. [Google Scholar] [CrossRef] [PubMed]
  144. Fung, T.C.; Vuong, H.E.; Luna, C.D.G.; Pronovost, G.N.; Aleksandrova, A.A.; Riley, N.G.; Vavilina, A.; McGinn, J.; Rendon, T.; Forrest, L.R.; et al. Intestinal serotonin and fluoxetine exposure modulate bacterial colonization in the gut. Nat. Microbiol. 2019, 4, 2064–2073. [Google Scholar] [CrossRef] [PubMed]
  145. Liu, Y.; Shi, C.; Zhang, G.; Zhan, H.; Liu, B.; Li, C.; Wang, L.; Wang, H.; Wang, J. Antimicrobial mechanism of 4-hydroxyphenylacetic acid on Listeria monocytogenes membrane and virulence. Biochem. Biophys. Res. Commun. 2021, 572, 145–150. [Google Scholar]
Figure 1. Schematic representation of the metabolic fate of undigested endogenous/dietary proteins and free amino acids in the large intestine luminal fluid. Undigested proteins are substrates for bacterial proteases and peptidases which release amino acids. Amino acids are little absorbed by the colonic epithelium, but are likely mostly used for bacterial protein synthesis, energy metabolism and other metabolic pathways, then releasing numerous bacterial metabolites. Several among these amino acid-derived metabolites have been shown to be active on the biology of microorganisms in terms of signaling, metabolism, physiology, and growth.
Figure 1. Schematic representation of the metabolic fate of undigested endogenous/dietary proteins and free amino acids in the large intestine luminal fluid. Undigested proteins are substrates for bacterial proteases and peptidases which release amino acids. Amino acids are little absorbed by the colonic epithelium, but are likely mostly used for bacterial protein synthesis, energy metabolism and other metabolic pathways, then releasing numerous bacterial metabolites. Several among these amino acid-derived metabolites have been shown to be active on the biology of microorganisms in terms of signaling, metabolism, physiology, and growth.
Microorganisms 11 01317 g001
Figure 2. Schematic representation of the production of bacterial metabolites from amino acid precursors, and effects of these metabolites on microorganism biology.
Figure 2. Schematic representation of the production of bacterial metabolites from amino acid precursors, and effects of these metabolites on microorganism biology.
Microorganisms 11 01317 g002
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

Blachier, F. Amino Acid-Derived Bacterial Metabolites in the Colorectal Luminal Fluid: Effects on Microbial Communication, Metabolism, Physiology, and Growth. Microorganisms 2023, 11, 1317. https://doi.org/10.3390/microorganisms11051317

AMA Style

Blachier F. Amino Acid-Derived Bacterial Metabolites in the Colorectal Luminal Fluid: Effects on Microbial Communication, Metabolism, Physiology, and Growth. Microorganisms. 2023; 11(5):1317. https://doi.org/10.3390/microorganisms11051317

Chicago/Turabian Style

Blachier, François. 2023. "Amino Acid-Derived Bacterial Metabolites in the Colorectal Luminal Fluid: Effects on Microbial Communication, Metabolism, Physiology, and Growth" Microorganisms 11, no. 5: 1317. https://doi.org/10.3390/microorganisms11051317

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