ArticlePDF AvailableLiterature Review

Biological Effects of Indole-3-Propionic Acid, a Gut Microbiota-Derived Metabolite, and Its Precursor Tryptophan in Mammals' Health and Disease

Authors:

Abstract

Actions of symbiotic gut microbiota are in dynamic balance with the host's organism to maintain homeostasis. Many different factors have an impact on this relationship, including bacterial metabolites. Several substrates for their synthesis have been established, including tryptophan, an exogenous amino acid. Many biological processes are influenced by the action of tryptophan and its endogenous metabolites, serotonin, and melatonin. Recent research findings also provide evidence that gut bacteria-derived metabolites of tryptophan share the biological effects of their precursor. Thus, this review aims to investigate the biological actions of indole-3-propionic acid (IPA), a gut microbiota-derived metabolite of tryptophan. We searched PUBMED and Google Scholar databases to identify pre-clinical and clinical studies evaluating the impact of IPA on the health and pathophysiology of the immune, nervous, gastrointestinal and cardiovascular system in mammals. IPA exhibits a similar impact on the energetic balance and cardiovascular system to its precursor, tryptophan. Additionally, IPA has a positive impact on a cellular level, by preventing oxidative stress injury, lipoperoxidation and inhibiting synthesis of proinflammatory cytokines. Its synthesis can be diminished in the presence of different risk factors of atherosclerosis. On the other hand, protective factors, such as the introduction of a Mediterranean diet, tend to increase its plasma concentration. IPA seems to be a promising new target, linking gut health with the cardiovascular system .
Int. J. Mol. Sci. 2022, 23, 1222. https://doi.org/10.3390/ijms23031222 www.mdpi.com/journal/ijms
Review
Biological Effects of Indole-3-Propionic Acid, a Gut
Microbiota-Derived Metabolite, and Its Precursor Tryptophan
in Mammals’ Health and Disease
Piotr Konopelski * and Izabella Mogilnicka
Laboratory of the Centre for Preclinical Research, Department of Experimental Physiology and
Pathophysiology, Medical University of Warsaw, 02-091 Warsaw, Poland; izabellamogilnicka@gmail.com
* Correspondence: piotr.konopelski@wum.edu.pl; Tel.: +48-22-57-20-734
Abstract: Actions of symbiotic gut microbiota are in dynamic balance with the host’s organism to
maintain homeostasis. Many different factors have an impact on this relationship, including bacte-
rial metabolites. Several substrates for their synthesis have been established, including tryptophan,
an exogenous amino acid. Many biological processes are influenced by the action of tryptophan and
its endogenous metabolites, serotonin, and melatonin. Recent research findings also provide evi-
dence that gut bacteria-derived metabolites of tryptophan share the biological effects of their pre-
cursor. Thus, this review aims to investigate the biological actions of indole-3-propionic acid (IPA),
a gut microbiota-derived metabolite of tryptophan. We searched PUBMED and Google Scholar da-
tabases to identify pre-clinical and clinical studies evaluating the impact of IPA on the health and
pathophysiology of the immune, nervous, gastrointestinal and cardiovascular system in mammals.
IPA exhibits a similar impact on the energetic balance and cardiovascular system to its precursor,
tryptophan. Additionally, IPA has a positive impact on a cellular level, by preventing oxidative
stress injury, lipoperoxidation and inhibiting synthesis of proinflammatory cytokines. Its synthesis
can be diminished in the presence of different risk factors of atherosclerosis. On the other hand,
protective factors, such as the introduction of a Mediterranean diet, tend to increase its plasma con-
centration. IPA seems to be a promising new target, linking gut health with the cardiovascular sys-
tem.
Keywords: gut microbiota; indole-3-propionic acid; oxidative stress; cardiovascular system
1. Introduction
Over the past two decades, research interest on the interactions between diet, gut
microbiota and their host organism has grown. The findings bring new information on
the correlation between the activity of symbiotic gut microbiota and the pathophysiology
of lifestyle diseases, including obesity [1], diabetes [2] and hypertension [3,4]. Initially,
research focused mainly on the role of short-chain fatty acids (SCFAs) [3,5] and carnitine-
derived metabolites [6–8]. The new data suggest that tryptophan, the essential amino acid,
can also be metabolized by microbiota, leading to the synthesis of biologically active
group of indoles [9]. So far, research projects on bacterial metabolism of tryptophan fo-
cused mainly on actions of indole [10,11], and its liver metabolite, indoxyl sulfate (IS)
[12,13]. Indole decreases intestinal inflammation and has positive impact on gastrointes-
tinal and liver homeostasis [10,14]. On the other hand, IS can be classified as uremic toxin,
as its concentration increases significantly in chronic kidney disease [9,15]. IS is proposed
to be one of the factors linking kidney dysfunction with an increased risk of developing
cardiovascular disease [12,16,17].
Citation: Konopelski, P.; Mogilnicka,
I. Biological Effects of Indole-3-
Propionic Acid, a Gut
Microbiota-Derived Metabolite, and
Its Precursor Tryptophan in
Mammals’ Health and Disease. Int. J.
M
ol. Sci. 2022, 23, 1222.
https://doi.org/10.3390/ijms23031222
Academic Editor: Burkhard
Poeggeler
Received: 16 December 2021
Accepted: 19 January 2022
Published: 22 January 2022
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional
claims in published maps and institu-
tional affiliations.
Copyright: © 2022 by the authors. Li-
censee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative Commons At-
tribution (CC BY) license (http://crea-
tivecommons.org/licenses/by/4.0/).
Int. J. Mol. Sci. 2022, 23, 1222 2 of 21
Until now, the biological actions of indole-3-propionic acid (IPA), another microbiota
derived metabolite of tryptophan, have not been properly reviewed in the scientific liter-
ature. A review of articles summarizing the impact of tryptophan metabolism on cardio-
vascular system homeostasis has recently been published, however; the main focus of that
paper was kynurenine- and the serotonin pathway [18]. The aim of our study is to provide
a comprehensive review of the physiological roles of IPA, and changes in its synthesis in
neurological, gastrointestinal and cardiovascular diseases. Scientific papers evaluating the
biological effects of this metabolite have multiplied greatly over the past few years [19–
22]. IPA has a protective role on a cellular and tissue level, by limiting inflammation [23],
lipid peroxidation [24], and the formation of free radicals [25]. Additionally, IPA affects
the function of the immune [22], nervous [26], gastrointestinal [27] and cardiovascular
system [19], and its synthesis decreases significantly in several pathogenic states and dis-
eases, including colitis [28], diabetes [29], and obesity [30]. The novelty of this review is
that we assess the available evidence on biological effects of IPA and emphasize its poten-
tial therapeutic applications.
2. Gut Microbiota
The close relationship between host and symbiotic gut microbiota has gained scien-
tific interest over the past few decades. It has been observed that this relationship is bidi-
rectional and symbiotic, with both participants benefiting from this union. The host pro-
vides a specific niche, supplies nutrients and optimal conditions for the bacteria to grow.
The role of gut microbiota in mammalian health goes beyond the synthesis of vitamins
[31] and is associated with many biochemical pathways and the synthesis of specific me-
tabolites that can be absorbed into the circulation [32]. Gut bacteria participate in the
breakdown of haemoglobin through their ability to transform bilirubin into stercobilino-
gen, enabling elimination of the latter in faeces [33]. Additionally, gut bacteria metabolize
primary bile acids, synthesized in the liver to form secondary bile acids, demonstrating a
protective role in gastrointestinal physiology [34].
2.1. Gut-Blood Barrier (GBB) and Microbiota
The intestinal lumen forms a specific environment for symbiotic microbiota, enabling
its growth and physiological function. Certain factors, including oxygen concentration
and composition of nutrients, form specific conditions of this microbial environment. It
has been observed that composition and richness of microbiota differ significantly be-
tween the small and large intestine [35]. Even within the large intestine some researchers
have provided evidence that caecal and colon contents should be analysed separately, due
to the significant differences in conditions for growth of symbiotic bacteria [36,37]. In ro-
dents, cecum is proposed to have critical role in bacterial fermentation, however; its local-
ization and distance from the anal margin disturb its evaluation in live subjects [38]. In-
tercaecal administration of investigated compounds and the collection of samples requires
surgical preparation [39,40] and potential use of antibiotics that might significantly affect
microbial data. Due to this fact, the colon is more frequently used for the evaluation of
microbial communities and their impact on health in mammals [41,42]. The gut lumen is
separated from the circulation by layers of tissues, which altogether form the gut-blood
barrier (GBB). This barrier is not limited to the epithelium and mucous layers. Gut-asso-
ciated lymphoid tissue (GALT), symbiotic bacteria and secretory proteins, including lac-
toferrin, prevent the transmission of pathogenic bacteria and their toxins into circulation
[43]. Disruption of this protective action of GBB has been observed in several diseases,
including diabetes, hypertension, and inflammatory bowel diseases [44,45], leading to in-
creased permeability to gut bacteria-derived metabolites, including trimethylamine
(TMA) [45]. Jaworska et al. observed increased permeability of GBB to SCFAs in an acetic
acid-induced rat model of colitis, and paediatric patients with inflammatory bowel dis-
ease [44].
Int. J. Mol. Sci. 2022, 23, 1222 3 of 21
2.2. Gut Microbiota-Derived Metabolites
Several dietary nutrients were proven to undergo both endogenous and bacterial me-
tabolism. For example, dietary choline can be absorbed in the intestines and used for the
synthesis of phospholipids, vital components of cells’ membranes and lipoproteins, taking
part in the transport of triglycerides and cholesterol between liver and peripheral organs
[46]. Additionally, choline is used for the synthesis of acetylcholine, an important neuro-
transmitter of the parasympathetic nervous system regulating vital physiological func-
tions at a resting state [47]. Furthermore, choline can also be transformed by gut microbi-
ota, leading to the synthesis of TMA [48], which has been revealed to have a negative
impact on the cardiovascular and nervous system in mammals by increasing blood pres-
sure [49], promoting formation of atherosclerotic plaques [50] and taking part in the path-
ophysiology of epilepsy and behavioural disorders [51]. TMA can be then further oxidized
to trimethylamine N-oxide (TMAO) by hepatic flavin monooxygenases [52]. TMAO syn-
thesis increases after phosphatidylcholine- and L-carnitine-rich meals [53], and depends
on the metabolism by gut microbiota, which was demonstrated in humans [54]. TMAO
has multiple known effects on the host, from lowering blood pressure and increasing di-
uresis [55], to acting as an osmolyte in order to protect mammalian cells from high hydro-
static and osmotic stress [56].
Dietary substrates, e.g., sulfates, sulfites and cysteine are also transformed into hy-
drogen sulfide (H2S) by microbes such as E. coli and Desulfovibrio or Enterobacter species
[57]. The vasodilatory and hypotensive properties of this gaseous transmitter have been
known for decades [58–60]. Recently, many studies have supported the hypothesis that
high and low concentrations of H2S have different biological effects on mammalian
health. Namely, low levels of hydrogen sulfide seem to protect the cellular bioenergetics
as well as intestinal epithelium integrity, while its high concentrations exert toxic effects
[61,62].
Diet rich in fibre is associated with the growth of specific bacteria producing SCFAs,
including acetic, propionic, butyric, and valeric acids [53,63]. SCFAs have a significant
impact on health in mammals by showing anti-inflammatory [64], hypotensive [5,65,66]
and hypolipidemic [67] effects, as well as by improving endothelial dysfunction induced
by angiotensin II [68]. Moreover, SCFAs have vasodilatory properties, which were inves-
tigated in coronary and colonic circulation, among others [69–72]. Additionally, microbi-
ota-derived indoles, metabolites of exogenous tryptophan, were proven to affect the nerv-
ous, immune, gastrointestinal, and cardiovascular systems in mammals [9,73], which will
be further discussed in this review.
3. Tryptophan Metabolism
Tryptophan is an essential amino acid vital for maintaining health and homeostasis
afforded by its complex metabolism (Figure 1) and biological actions.
Int. J. Mol. Sci. 2022, 23, 1222 4 of 21
Figure 1. Simplified representation of metabolic pathways of tryptophan in mammals. Tryptophan
can be metabolized by host’s own cells (endogenous pathways, grey arrows) and by symbiotic gut
microbiota (bacterial pathways, black arrow). TpH–tryptophan hydroxylase; TDO–tryptophan 2,3-
dioxygenase; IDO–indoleamine 2,3-dioxygenase; TAA–tryptophan aminotransferase; ArAT–aro-
matic amino acid aminotransferase.
3.1. Kynurenine Pathway of Tryptophan Metabolism
In the kynurenine pathway (KP), which accounts for around 95% of tryptophan ca-
tabolism, tryptophan is oxidized to N-formylkynurenine (NFK) mainly by tryptophan 2,3-
dioxygenase (TDO) located in the liver [18,74]. It is the first and rate-limiting step in this
pathway and its activity is regulated by steroids, including cortisol, and systemic levels of
tryptophan [75]. TDO is highly selective when it comes to substrates and works specifi-
cally with tryptophan [76]. Other enzymes, including indoleamine 2, 3-dioxygenase 1
(IDO) and indoleamine 2, 3-dioxygenase 2 (IDO2) contribute to tryptophan breakdown in
extrahepatic tissues and accept other substrates as well. Under normal circumstances,
those enzymes are significantly less active than TDO and thus, a great part of KP takes
place in the liver. However, it has been reported that inflammation might increase the
significance of extrahepatic kynurenine (Kyn) formation [75]. Furthermore, formidase
transforms NFK to Kyn. It is further metabolized by numerous enzymes into its deriva-
tives such as anthranilic acid, kynurenic acid and quinolinic acid. The latter is converted
into nicotinamide adenine dinucleotide (NAD) in a final step of KP [18,74].
3.2. Serotonergic Pathway of Tryptophan Metabolism
Serotonergic pathway degrades only a small fraction (1–2%) of ingested tryptophan.
Two essential enzymes involved in these processes are tryptophan hydroxylase 1 and 2
(TPH1 and TPH2). They produce an active metabolic intermediate, 5-hydroxytryptamine
(serotonin, 5-HT), in the gut (TPH1) and in the brain (TPH2) [74], which is further trans-
formed into melatonin. Serotonin not only works as a neurotransmitter in the central nerv-
ous system, but also controls several physiological functions from the motility of the gas-
trointestinal tract to glucose homeostasis [77]. Serotonin produced in the gut is released
into the blood stream, where platelets use it as a signalling molecule in clot formation [78].
Int. J. Mol. Sci. 2022, 23, 1222 5 of 21
Its metabolites also play important physiological roles and can be used for diagnostic pur-
poses. One of them, melatonin, regulates circadian rhythm and has anti-inflammatory
properties [79]. Finally, measurement of urine levels of 5-HIAA (5-hydroxy indoleacetic
acid, a waste product of serotonin breakdown), is used to estimate serotonin levels in pa-
tients with serotonin-secreting neuroendocrine tumours [80,81].
3.3. Bacterial Metabolism of Tryptophan
Ingested tryptophan is, in large part, absorbed in the intestines to be further metab-
olized by host’s cells. Fractions of this metabolite remaining in intestinal lumen can be
absorbed by symbiotic microbiota, enabling bacterial growth and function. Gut bacteria
use this amino acid for their own needs, simultaneously producing biologically active me-
tabolites that can influence the host’s homeostasis. Symbiotic microorganisms directly
convert tryptophan to indole, skatole, indole-3-acetic acid (IAA), IPA, and indole-3-alde-
hyde (IAld) [9,18,73].
3.3.1. Formation of IPA by Gut Microbiota
Bacteria taking part in the gut formation of IPA include Lactobacillus reuteri [82], Ak-
kermansia and Clostiridum genus [83,84], including species Clostridium sporogenes [85–88]
and Clostridium caloritolerans [88], as well as some Peptostreptococci [89]. Microbial path-
way of IPA production is primarily controlled by tryptophan aminotransferase (TAA, ar-
omatic amino acid aminotransferase, ArAT) [73,89]. Additionally, it has been also estab-
lished that bacterial tryptophanase enables synthesis of IPA in the gut [18].
3.3.2. Formation of Other Indoles by Gut Microbiota
Multiple genera and species participate in the synthesis of specific indoles from tryp-
tophan [89]. According to Roager et al. Escherichia coli, Clostridium spp. and Bacteroides spp.
catabolize tryptophan to indole using the enzyme tryptophanase [89]. Furthermore, main
bacteria producing IAA are Bacteroides such as Bacteroides ovatus, B. eggerthii, B. thetaiota-
omicron, and B. fragilis, as well as some representants of the Clostridium, Bifidobacterium
and Eubacterium genus [21,90]. Decarboxylase and tryptophanase take part in IAA for-
mation [18]. Finally, several Lactobacilli can also synthesize IAld using aromatic amino
acid aminotransferase [21,90].
4. Biological Effects of Tryptophan and IPA
4.1. Tryptophan and Immune System
The relationship between tryptophan and the immune system is bidirectional. On
one hand, tryptophan and its metabolites have an impact on the expression of interleu-
kins. On the other hand, it has been observed that, in the presence of inflammatory and
autoimmune diseases, tryptophan metabolism shifts, leading to increased synthesis of
kynurenines [91,92]. Tryptophan breakdown by IDO is associated with immune system
function, since metabolites of the KP reveal immunomodulatory activity, by reducing Th-
17 cells formation and promoting formation of regulatory T cells [93]. These effects justify
increased IDO expression in pregnancy, as a factor enabling pregnancy tolerance in mam-
mals [94]. Furthermore, IDO expression increases in viral [93], bacterial [95], and parasitic
[96] infections and states associated with increased expression of tumor necrosis factor α
(TNF-α) and interferon-γ (IFN-γ) [97]. Additionally, the expression of this enzyme is also
increased in autoimmune and neurodegenerative diseases, including rheumatoid arthri-
tis, multiple sclerosis, and Alzheimer’s disease [91]. Moreover, IDO expression increases
in carcinogenesis [98] and enhanced tryptophan breakdown via KP is associated with
poorer outcome and development of complications, including anaemia and fatigue, in
cancer patients [99]. IDO, as a first enzyme of the KP, promotes formations of Kyn and its
further metabolites, simultaneously decreasing concentration of their precursor, trypto-
phan. Hence, Kyn/Trp (kynurenine/tryptophan) ratio was proposed as one of the markers
Int. J. Mol. Sci. 2022, 23, 1222 6 of 21
of increased inflammatory response [74,93]. Interestingly, tryptophan itself reveals strong
antioxidative activity [100], and reduces LPS-induced lipoperoxidation [101]. Endogenous
metabolites of tryptophan, including melatonin, can also act as free radicals’ scavengers
[102]. Finally, decreased consumption of tryptophan in diet is associated with increased
serum levels of pro-inflammatory cytokines, including IL-1alpha [103].
4.2. Tryptophan and Body Mass Regulation
Tryptophan is an essential amino acid and component of a balanced diet. Its presence
in the diet is vital for protein synthesis, metabolism and other functions maintaining ho-
meostasis. Research shows that excessive tryptophan consumption and tryptophan defi-
ciency in the diet can affect body mass regulation in mammals. Rats consuming both tryp-
tophan-low and tryptophan-free chow experienced a significant reduction in body weight
gain [104–106]. Interestingly, supplementing tryptophan in the diet reduces food intake
and weight gain in rats [105,107]. Similar effects were observed by intragastric administra-
tion of this amino acid in mice [108].
4.3. Tryptophan and Cardiovascular System Regulation
Regulation of the cardiovascular system involves action of many specific tissues and
hormones to adapt to rapid changes in blood pressure and other hemodynamic parame-
ters. Small molecules, such as noradrenaline and adrenaline, which are metabolites of
amino acid tyrosine, are well-known regulators of cardiovascular system function [109].
Tryptophan also affects hemodynamic parameters. With tyrosine and histidine, trypto-
phan belongs to a group of sensitizers of β-adrenergic receptors (ESBAR) [110]. Addition-
ally, tryptophan increases the contractility of human myocardial cells, demonstrating an
inotropic property [111]. Oral administration of this amino acid increases portal blood
pressure and produces a trend towards higher mean arterial blood pressure in rats [11].
Parenteral tryptophan infusion increases blood pressure in normotensive rats [112]. Ad-
ditionally, tryptophan given orally and parenterally reduces sodium excretion in the kid-
neys revealing an antidiuretic effect [113,114]. On the other hand, parenteral infusion of
tryptophan decreases blood pressure in spontaneously hypertensive rats [112]. The hypo-
tensive effect was also observed after oral administration to patients with essential hyper-
tension [115]. Moreover, IDO activity is also associated with blood pressure regulation.
Increased IDO activity in mice infected with malarial parasite was associated with a de-
crease in systolic blood pressure. Interestingly, inhibition of IDO significantly increased
blood pressure in infected mice [96], a pattern similar to effects of tryptophan administra-
tion in normotensive rats [112]. These observations show complex biological effects of
tryptophan that might be explained, at least partially, by the action of its endogenous and
microbiota-derived metabolites, including IPA.
5. Biological Effects of IPA and Its Impact on Health in Mammals
Knowledge on the biological action of IPA has increased significantly over the past
several years, giving new evidence on the positive and protective effects of this metabolite
(Figure 2).
Int. J. Mol. Sci. 2022, 23, 1222 7 of 21
Figure 2. Protective effects of indole-3-propionic acid (IPA) on cellular and tissue level. GBB- gut-
blood barrier; ROS–reactive oxygen species.
5.1. IPA Improves Gut-Blood Barrier Function
Studies on the GBB demonstrate that indole [10] and IPA [20,116,117], two trypto-
phan metabolites, improve barrier properties by increasing the expression of claudins and
other tight junction proteins. Additionally, IPA increases secretion of mucins in in vitro
human colonic culture [20] and increases the number of goblet cells and mucosa thickness
in rats [118]. Moreover, IPA acts as a ligand of the aryl hydrocarbon receptor (AhR) pre-
sent in colonic epithelial cells, activation of which is associated with anti-inflammatory
and anticancer effects [119] Interestingly, it has been observed that in patients suffering
from colitis serum, IPA concentration decreases significantly [28]. Apart from being an
AhR ligand, IPA can also produce its biological action by activating the pregnane X recep-
tor (PXR) present in colonic, liver endothelial and muscle cells [23,119,120]. Complex la-
boratory techniques allow testing the function of GBB by measuring its permeability to
specific substances, including Fluorescein Isothiocyanate-Dextran (FITC-Dextran). Three
independent research projects revealed that IPA decreases FITC-dextran-dependent gut
permeability in mice [121–123]. Venkatesh et al. proved that IPA decreases gut permea-
bility by interacting with PXR [121].
5.2. IPA Protects against Oxidative Stress and Attenuates Inflammation
Physiological activity of the immune system is important in the regulation of mam-
malian health and represents a dualistic approach. Its activation limits the invasion of
pathogens and is a key mechanism protecting from infections. On the other hand, regula-
tory mechanisms moderating the immune response must be preserved to prevent the ex-
cessive activation of immune cells associated with autoimmune diseases. Bacterial metab-
olites of tryptophan have a complex impact on the immune system, revealing both pro-
and anti-inflammatory properties. IS, the liver metabolite of microbiota-derived indole,
promotes the production of reactive oxygen species (ROS) [124] and induces expression
of proinflammatory cytokines, including IL-1β [125], TNF-α [125] and MCP-1 [125–127].
Interestingly, IPA protects cells from ROS [25,128,129], oxidative damage and lipid perox-
idation caused by potassium bromate [130–132], potassium iodate [24], iron (II) sulphate
[133,134], iron (III) chloride [135], chromium (III) chloride [136], and 2,2-Azino-bis(3-
ethylbenzothiazoline-6-sulfonic acid (ABTS) [137]. Additionally, IPA decreases the ex-
pression of proinflammatory cytokines, including TNF-α [22,27,121], IL-1β [23,27], IL-6
[27], IL-12 [22], IL-13 [22] and MCP-1 [22,82]. Interestingly, IPA also shows antimicrobial
activity by inhibiting the growth of Legionella pneumophila [138,139], and Salmonella typhi-
murium [140]. Apart from that, in patients with type II diabetes, serum concentration of
Int. J. Mol. Sci. 2022, 23, 1222 8 of 21
IPA is negatively correlated with high-sensitivity C-reactive protein (hsCRP) [141]. More-
over, Nyström et al. observed that in immunodeficient patients with HIV infection syn-
thesis of IPA is significantly diminished [142].
5.3. IPA Protects against Carcinogens and Has an Antitumor Potential
Carcinogenesis is a complex and complicated process of modification of healthy cells
into autonomous and self-sufficient pathogenic conglomerate of cells. Many factors en-
hancing carcinogenesis have been established, including ionizing radiation, ultraviolet
(UV) radiation, oxidative stress, DNA damage, viral infections, and lifestyle factors, such
as tobacco smoke, ethanol consumption, and ingestion of nutritional carcinogens
[143,144]. Increased formation of ROS has a negative impact on lipid barriers, affects DNA
structure, and takes part in carcinogenesis; therefore antioxidants, both nutritional and
pharmacological, have been targets of great research interest over the past years [145,146].
Both melatonin and IPA can be classified as free radical’s scavengers, and their antitumor
potential have been investigated in the scientific literature [136,147]. IPA prevents DNA
damage in hamsters’ kidneys exposed to oestradiol [147], rats’ brains exposed to chlorpyr-
ifos [148], and calf thymus samples exposed to chromium (III) chloride [136]. Addition-
ally, IPA decreases fluidity of rats’ liver microsomal membranes incubated with chro-
mium (III) chloride [149]. Due to its positive cellular effects, IPA has become a promising
particle used for the chemical modification of well-known antineoplastic drugs [150] and
compounds [151,152]. Addition of an IPA particle as a ligand to a cisplatin structure
caused significant cytotoxic effect, associated with increased ROS formation [150]. Cad-
mium-IPA complexes possess antiproliferative and proteasome-inhibitory activity in in
vitro breast cancer cells [151]. Interestingly, IPA alleviates hematopoietic and gastrointes-
tinal side effects of radiotherapy in the mice model [153]. Further research projects should
focus on testing the impact of IPA on neoplastic cells’ survival and evaluate its role in both
cancer prevention and therapy.
5.4. IPA has a protective role in neurodegenerative disease models
The pathophysiology of neurodegenerative diseases is complex, and many possible
pathways have been proposed as therapeutic targets to slow down disease progression
and preserve neurological functions of the affected individuals. In general, the develop-
ment of Alzheimer’s, Parkinson’s and Huntington’s diseases is associated with patholog-
ical accumulation of specific proteins in vital brain centres responsible for cognition,
memory, motor activity and other essential neurological processes [154,155].
5.4.1. IPA in Alzheimer’s Disease
Abnormal folding of amyloid β-proteins and their deposition in amyloid plaques re-
veals neurodegenerative properties due to the activation of ROS[156].
Melatonin has antioxidative action, and also increases clearance of amyloid β-protein
in the mice model of Alzheimer’s disease amyloidosis [157]. IPA shares protective biolog-
ical effects against oxidative stress with melatonin [24,133] and decreases aggregation of
amyloid β-protein in in vitro experiments [158]. The addition of IPA to media of the pri-
mary neurons and neuroblastoma cells exposed to amyloid β-protein significantly re-
duced toxicity and prevented cell death [159]. Dragicevic et al. observed that both IPA and
melatonin restore mitochondrial function in the in vitro model of Alzheimer’s disease
[160]. Interestingly, IPA also protects brain tissue from oxidative stress, lipid peroxidation
and oxidative DNA damage in hippocampal region after acute brain ischemia [26]. Addi-
tionally, IPA acts synergistically with glutathione to prevent ABTS-related formation of
free radicals in the rat brain and reduces associated lipid peroxidation [128]. Surprisingly,
Huangs et al. observed a trend towards higher plasma IPA levels in patients with pro-
gressive mild cognitive impairment (MCI) and Alzheimer’s disease, compared to patients
Int. J. Mol. Sci. 2022, 23, 1222 9 of 21
with stable MCI. However, these changes were not statistically significant [161]. Addi-
tional research data are needed to fully evaluate role of IPA as a protective or predictive
factor in Alzheimer’s disease.
5.4.2. IPA and other Neurodegenerative Diseases
IPA exerts a chemical chaperone-like activity and inhibits abnormal aggregation of
the regular [162] and denatured [163] proteins. In two cell culture models of Parkinson’s
disease associated with overexpression of Parkin-associated endothelin receptor-like re-
ceptor (Pael-R) and α-synuclein, IPA significantly reduced ROS-associated cell death [163]
Additionally, patients with Huntington’s disease have lower plasma concentration of IPA,
which might be associated with system-wide decreased ability to protect against ROS for-
mation [87]. On the contrary, increased IPA formation was observed in mice with experi-
mental autoimmune encephalitis (EAE) [164], animal model of multiple sclerosis (MS),
and patients with relapsing-remitting MS [165].
5.5. IPA has a Positive Impact on Cardiovascular Disease Risk Factors
Initially endogenous metabolites of tryptophan were suspected to be accountable for
tryptophan’s cardiovascular actions [166]. Recently, it has been established that IPA itself
can also affect cardiovascular system [19,167]. Additionally, its synthesis can change in
presence of cardiovascular-related diseases (Table 1).
Table 1. Comparison of tendencies in the synthesis of indole-3-propionic acid (IPA) associated with
protective and harmful factors of cardiovascular diseases.
Impact of a Factor on
Cardiovascular Health
Factor Affecting
Cardiovascular Health
Change in the
Synthesis of IPA References
Positive
Mediterranean diet Increase [168]
Increased composition of fibre
in the diet Increase [29,141]
Increased mulberry consumption Increase [83]
Negative
Diabetes Decrease [29]
Dyslipidaemia Decrease [169]
Obesity Decrease [30]
Atherosclerosis Decrease [170]
5.5.1. Diet
The composition of specific nutrients in diet can affect mammalian health and disrupt
lipid and carbohydrate homeostasis. Diets rich in red meat and saturated fatty acids are
long-established cardiovascular disease risk factors [171]. Additionally, diets with a high
content of red meat increase the synthesis of TMA, a bacterial metabolite formed mainly
from carnitine [6,172]. Complex nutritional interventions stimulate the formation of other
microbiota-derived metabolites, including IPA. The Mediterranean diet, known for its
beneficial and cardio-protective role, significantly increases the plasma concentration of
IPA in humans [168]. Moreover, diets rich in fibre [29,141] and inulin [173] also promote
the synthesis of this metabolite. On the contrary, a fast-food diet, a known risk factor for
cardiovascular diseases, significantly reduces the plasma concentration of IPA in humans
[168]. Interestingly, the addition of IPA to a high-fat diet in mice restores bone minerali-
zation and osteoblasts’ function diminished in mice ingesting a high-fat diet alone [174].
5.5.2. Dyslipidaemia
Disturbances in lipid metabolism, including the rise in total cholesterol levels, have
been linked to an increased risk of cardiovascular disease-related mortality [175]. Associ-
ation of high concentration of triglycerides and low HDL cholesterol level is classified as
atherogenic dyslipidaemia and is linked to the progression of atherosclerosis [176]. Long-
Int. J. Mol. Sci. 2022, 23, 1222 10 of 21
lasting disturbances in lipid metabolism lead to obesity, metabolic syndrome and non-
alcoholic fatty liver disease [177]. Plasma concentration of IPA is significantly reduced in
patients with dyslipidaemia [169]. Hypolipidemic interventions, including the admin-
istration of mulberry leaf extract and 1-deoxynojirimycin, are associated with increased
IPA concentration in stools [83]. Additionally, IPA plasma concentration correlates nega-
tively with lipid parameters, including triglycerides and LDL-C plasma levels [173]. More-
over, IPA reduces hepatic steatosis and hepatocyte dysfunction in a rat model of high-fat
diet-induced steatohepatitis [27]. In patients with hepatic lobular inflammation and liver
fibrosis significant decrease in circulating IPA levels in serum can be observed [178]. Sur-
prisingly, IPA enhances liver damage in mice with carbon tetrachloride-induced liver fi-
brosis, without affecting the liver functions of healthy controls [179]. Previous observa-
tions need to be further investigated to establish whether IPA has a positive or negative
effect on liver function. There is only one research paper, in which IPA failed to reveal
protective metabolic effects in mice fed Western diet, while simultaneously improving
intestinal functions [122]. Taken together, more research data is needed to fully under-
stand regulatory role of IPA in lipid, metabolic, and liver homeostasis. It is possible that
specific metabolic changes occur when the concentration of this metabolite in peripheral
blood increases above a certain threshold that needs to be established in further experi-
ments.
5.5.3. Obesity
The number of patients with excess body fat increases drastically each year, and data
extrapolations suggest this trend will continue [180]. Increased body weight and obesity
are associated with a broad group of metabolic disturbances due to excessive adipose tis-
sue accumulation and increased cytokine synthesis [181]. Previously, we discussed the
positive role of tryptophan in the reduction in body mass. Recent results from our labor-
atory showed that parenteral administration of IPA significantly recuses weight gain in
rats [105]. Additionally, human studies showed that in obese female patients, a significant
decrease in the concentration of IPA in serum and follicular fluid could be observed com-
pared to women with normal weight [30]. Apart from that, in an animal model of gluco-
corticoid withdrawal syndrome, IPA formation is significantly diminished, showing that
its synthesis might also by affected by changes in the function of the adrenal glands [182].
5.5.4. Hyperglycaemia
Increased plasma glucose concentration is associated with severe complications, in-
cluding endothelial dysfunction, atherosclerosis progression, and lipid homeostasis
changes [183]. Diabetes and prediabetic states are considered significant risk factors for
cardiovascular disease, the prime cause of death in this patient population [184]. Despite
multiple therapeutic options and new innovative treatments, many patients struggle to
control their glycaemia adequately. New research findings suggest that carbohydrate ho-
meostasis might be affected by metabolites produced by gut bacteria, including indoles
[29,185]. Abildgaard et al. observed that introducing an IPA-enriched diet for 6 weeks
significantly reduces fasting blood glucose concentration and plasma insulin level in rats
[186]. Furthermore, in a human study of the Finnish population, patients who developed
type II diabetes had significantly lower serum concentration of IPA [29]. Additionally, the
concentration of this metabolite has been inversely correlated with the incidence of type
II diabetes and tended to be positively correlated with insulin secretion [141].
5.5.5. Hypertension
Blood pressure control depends on two main components: mechanical action of the
myocardium and peripheral vasculature [187]. Elevated blood pressure is an important
cardiovascular risk factor, as it takes part in the pathophysiology of atherosclerosis and
Int. J. Mol. Sci. 2022, 23, 1222 11 of 21
causes microcirculatory dysfunction and progressive tissue damage [188]. Bacterial me-
tabolites of tryptophan reveal similar hemodynamic patterns as their precursor. Results
from our laboratory show that both IS and indole increase blood pressure in normotensive
rats, with and without a concomitant increase in heart rate [13]. The scientific literature
indicates that IPA is also involved in blood pressure regulation in mammals [19]. Using
the Langendorff heart model in mice, an IPA-dependent increase in myocardial contrac-
tility was demonstrated [167]. Additionally, this metabolite causes vasoconstriction of the
endothelium-denuded mesenteric resistance arteries [19] and diminishes vasodilatation
associated with pre-treatment with sodium nitroprusside [189] and acetylcholine [19]. The
vascular effect of IPA might be mediated by the activation of PXR [189]. So far, it has not
been elucidated whether long-term administration of IPA takes part in the pathophysiol-
ogy of hypertension and hypertension-related cardiovascular diseases. Human studies
revealed that patients with advanced atherosclerosis had significantly reduced plasma
IPA concentrations [170]. Additionally, in this group, IPA levels correlated strongly with
higher ankle-brachial index (ABI) and less severe peripheral arterial disease (PAD) [170].
6. Modulation of IPA Concentration as a Therapeutic Target
6.1. Antibiotics
Knowledge of antibiotics and their potential in treating infections dates back to the
discovery of penicillin by Alexander Flemming in 1928 [190]. Nowadays, antibiotics are
some of the most frequently prescribed medications in everyday medical practice. Grow-
ing evidence suggests that using antibiotics affects pathogenic bacteria and strikes back
against symbiotic microbiota, leading to the development of serious diseases, including
Clostridioides difficile infection and pseudomembranous colitis [191]. Antibiotic groups dif-
fer in their pharmacokinetic properties, which affects their distribution and action site in
the organism. For example, neomycin given orally acts predominantly in the gut due to
its low absorption in the intestines [192]. Behr et al. tested how oral administration of spe-
cific antibiotics affects the synthesis of certain bacterial metabolites, including IPA. A 4-
week treatment with fluoroquinolones, tetracyclines and aminoglycosides significantly
reduced IPA plasma concentration, with the greatest effect on the third group [193]. Data
from our laboratory confirmed the observation, as mentioned above. A 2-week-long oral
administration of neomycin significantly reduced the levels of IPA in the stool, portal, and
peripheral blood of Sprague Dawley rats [105]. Other antibiotics, including ampicillin,
might also affect the bacterial metabolism of tryptophan [82].
6.2. Tryptophan Concentration in Diet
Several factors affect IPA metabolism in mammals. Diet, medications, and intestinal
disturbances can influence the concentration of this metabolite in the gut. The introduc-
tion of a tryptophan-rich diet for 2 weeks significantly increases the concentration of IPA
in colon contents, portal, and peripheral blood [105]. On the other hand, a tryptophan-free
diet, administered for the same period, significantly reduces the synthesis of this metabo-
lite [105]. Tryptophan concentration in the diet is not the only nutritional factor affecting
the synthesis of IPA [83,168].
6.3. Probiotics
Administration of specific bacterial species as probiotics has increased its recognition
as a therapeutic option in gastrointestinal diseases over the past few years [194,195]. Pro-
biotics co-administered with antibiotics seem to decrease the risk of developing severe
adverse effects, including Clostridioides difficile infection [194]. Probiotics restore healthy
gut microbiota composition and its functions, including synthesis of specific metabolites
[174,196–198]. Lactobacillus reuteri can be administered as a probiotic, revealing a positive
impact, by improving infantile colic symptoms [199,200]. Simultaneously it has been ob-
served that this bacterial species is able to synthetize IPA [82]. It still needs to be elucidated
Int. J. Mol. Sci. 2022, 23, 1222 12 of 21
whether the positive impact of L. reuteri might be due to the synthesis of specific metabo-
lites, including IPA.
7. Materials and Methods
We searched PUBMED and Google Scholar databases to identify pre-clinical and
clinical studies on synthesis and biological effects of IPA. The key words included micro-
biota, tryptophan, indoles, indole-3-propionic acid. The search was confined to manu-
scripts that were published from 1961 to December 2021. Specific steps of the review pro-
cess and the evaluation of scientific papers are present in Figure 3. A total of 131 records
on IPA were obtained from databases, and an additional 10 papers were included from
other sources. A total of 141 papers were screened for relevance and 67 records were ex-
cluded from further analysis. Exclusion criteria were language other than English (2 rec-
ords), chemistry-focused papers (21 records), papers evaluating results from experiments
on subjects other than mammals and bacteria (24 records), and papers on indoles other
than IPA and not evaluating effects of IPA (20 records). Finally, 74 papers evaluating the
synthesis and biological effects of IPA were included in this review.
Figure 3. Schematic representation of study protocol of papers evaluating synthesis and biological
effects of IPA. Black arrows show steps of the review process. IPA–indole-3-propionic acid.
8. Conclusions
Symbiotic gut microbiota is able to produce biologically active metabolites that can
affect functions of the host. IPA belongs to the wide group of indoles, microbiota derived
metabolites of tryptophan. It has been observed that IPA has beneficial impact on host
health by possessing anti-inflammatory and ROS-scavenging activity. Further research
projects should evaluate its possible clinical applications in the treatment of autoimmune,
inflammatory and oncological diseases. Additionally, its synthesis decreases in many
pathological states associated with an increased risk of developing cardiovascular dis-
eases, making it a promising new pharmacotherapeutic target.
Funding: This research was funded by the PRELUDIUM grant of the Polish National Science Centre
(2019/35/N/NZ4/01111) to P.K.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: All relevant data are presented in this paper.
Acknowledgments: No assistance in the preparation of this article is to be declared.
Conflicts of Interest: The authors declare no conflict of interest.
Int. J. Mol. Sci. 2022, 23, 1222 13 of 21
References
1. Davis, C.D. The gut microbiome and its role in obesity. Nutr Today 2016, 51, 167–174.
https://doi.org/10.1097/NT.0000000000000167.
2. Gurung, M.; Li, Z.; You, H.; Rodrigues, R.; Jump, D.B.; Morgun, A.; Shulzhenko, N. Role of gut microbiota in type 2 diabetes
pathophysiology. EBioMedicine 2020, 51, 102590. https://doi.org/10.1016/j.ebiom.2019.11.051.
3. Durgan, D.J.; Ganesh, B.P.; Cope, J.L.; Ajami, N.J.; Phillips, S.C.; Petrosino, J.F.; Hollister, E.B.; Bryan, R.M., Jr. Role of the gut
microbiome in obstructive sleep apnea-induced hypertension. Hypertension 2016, 67, 469–474.
https://doi.org/10.1161/HYPERTENSIONAHA.115.06672.
4. Yang, T.; Santisteban, M.M.; Rodriguez, V.; Li, E.; Ahmari, N.; Carvajal, J.M.; Zadeh, M.; Gong, M.; Qi, Y.; Zubcevic, J.; et al. Gut
dysbiosis is linked to hypertension. Hypertension 2015, 65, 1331–1340. https://doi.org/10.1161/HYPERTENSIONAHA.115.05315.
5. Onyszkiewicz, M.; Gawrys-Kopczynska, M.; Konopelski, P.; Aleksandrowicz, M.; Sawicka, A.; Kozniewska, E.; Samborowska,
E.; Ufnal, M. Butyric acid, a gut bacteria metabolite, lowers arterial blood pressure via colon-vagus nerve signaling and GPR41/43
receptors. Pflugers Arch 2019, 471, 1441–1453. https://doi.org/10.1007/s00424-019-02322-y.
6. Koeth, R.A.; Wang, Z.; Levison, B.S.; Buffa, J.A.; Org, E.; Sheehy, B.T.; Britt, E.B.; Fu, X.; Wu, Y.; Li, L.; et al. Intestinal microbiota
metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat. Med. 2013, 19, 576–585.
https://doi.org/10.1038/nm.3145.
7. Ussher, J.R.; Lopaschuk, G.D.; Arduini, A. Gut microbiota metabolism of L-carnitine and cardiovascular risk. Atherosclerosis 2013,
231, 456–461. https://doi.org/10.1016/j.atherosclerosis.2013.10.013.
8. Li, D.Y.; Tang, W.H.W. Gut microbiota and atherosclerosis. Curr. Atheroscler. Rep. 2017, 19, 39. https://doi.org/10.1007/s11883-
017-0675-9.
9. Konopelski, P.; Ufnal, M. Indoles—Gut bacteria metabolites of tryptophan with pharmacotherapeutic potential. Curr. Drug Metab.
2018, 19, 883–890. https://doi.org/10.2174/1389200219666180427164731.
10. Bansal, T.; Alaniz, R.C.; Wood, T.K.; Jayaraman, A. The bacterial signal indole increases epithelial-cell tight-junction resistance
and attenuates indicators of inflammation. Proc. Natl. Acad. Sci. USA 2010, 107, 228–233. https://doi.org/10.1073/pnas.0906112107.
11. Huc, T.; Konop, M.; Onyszkiewicz, M.; Podsadni, P.; Szczepanska, A.; Turlo, J.; Ufnal, M. Colonic indole, gut bacteria metabolite
of tryptophan, increases portal blood pressure in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2018, 315, R646–R655.
https://doi.org/10.1152/ajpregu.00111.2018.
12. Lekawanvijit, S. Role of gut-derived protein-bound uremic toxins in cardiorenal syndrome and potential treatment modalities.
Circ. J. 2015, 79, 2088–2097. https://doi.org/10.1253/circj.CJ-15-0749.
13. Huc, T.; Nowinski, A.; Drapala, A.; Konopelski, P.; Ufnal, M. Indole and indoxyl sulfate, gut bacteria metabolites of tryptophan,
change arterial blood pressure via peripheral and central mechanisms in rats. Pharmacol. Res. 2018, 130, 172–179.
https://doi.org/10.1016/j.phrs.2017.12.025.
14. Knudsen, C.; Neyrinck, A.M.; Leyrolle, Q.; Baldin, P.; Leclercq, S.; Rodriguez, J.; Beaumont, M.; Cani, P.D.; Bindels, L.B.; Lanthier,
N.; et al. Hepatoprotective effects of indole, a gut microbial metabolite, in leptin-deficient obese mice. J. Nutr. 2021, 151, 1507–
1516. https://doi.org/10.1093/jn/nxab032.
15. Hobby, G.P.; Karaduta, O.; Dusio, G.F.; Singh, M.; Zybailov, B.L.; Arthur, J.M. Chronic kidney disease and the gut microbiome.
Am. J. Physiol. Renal. Physiol. 2019, 316, F1211–F1217. https://doi.org/10.1152/ajprenal.00298.2018.
16. Lekawanvijit, S.; Kompa, A.R.; Manabe, M.; Wang, B.H.; Langham, R.G.; Nishijima, F.; Kelly, D.J.; Krum, H. Chronic kidney
disease-induced cardiac fibrosis is ameliorated by reducing circulating levels of a non-dialysable uremic toxin, indoxyl sulfate.
PLoS ONE 2012, 7, e41281. https://doi.org/10.1371/journal.pone.0041281.
17. Gao, H.; Liu, S. Role of uremic toxin indoxyl sulfate in the progression of cardiovascular disease. Life Sci. 2017, 185, 23–29.
https://doi.org/10.1016/j.lfs.2017.07.027.
18. Melhem, N.J.; Taleb, S. Tryptophan: From diet to cardiovascular diseases. Int. J. Mol. Sci. 2021, 22, 9904.
https://doi.org/10.3390/ijms22189904.
19. Konopelski, P.; Chabowski, D.; Aleksandrowicz, M.; Kozniewska, E.; Podsadni, P.; Szczepanska, A.; Ufnal, M. Indole-3-
propionic acid, a tryptophan-derived bacterial metabolite, increases blood pressure via cardiac and vascular mechanisms in rats.
Am. J. Physiol. Regul. Integr. Comp. Physiol. 2021, 321, R969-R981. https://doi.org/10.1152/ajpregu.00142.2021.
20. Li, J.; Zhang, L.; Wu, T.; Li, Y.; Zhou, X.; Ruan, Z. Indole-3-propionic acid improved the intestinal barrier by enhancing epithelial
barrier and mucus barrier. J. Agric. Food. Chem. 2021, 69, 1487–1495. https://doi.org/10.1021/acs.jafc.0c05205.
21. Pappolla, M.A.; Perry, G.; Fang, X.; Zagorski, M.; Sambamurti, K.; Poeggeler, B. Indoles as essential mediators in the gut-brain
axis. Their role in Alzheimer’s disease. Neurobiol. Dis. 2021, 156, 105403. https://doi.org/10.1016/j.nbd.2021.105403.
22. Garcez, M.L.; Tan, V.X.; Heng, B.; Guillemin, G.J. Sodium Butyrate and Indole-3-propionic acid prevent the increase of cytokines
and kynurenine levels in LPS-induced human primary astrocytes. Int. J. Tryptophan. Res. 2020, 13, 1178646920978404.
https://doi.org/10.1177/1178646920978404.
23. Du, L.; Qi, R.; Wang, J.; Liu, Z.; Wu, Z. Indole-3-Propionic acid, a functional metabolite of clostridium sporogenes, promotes
muscle tissue development and reduces muscle cell inflammation. Int. J. Mol. Sci. 2021, 22, 12435.
https://doi.org/10.3390/ijms222212435.
24. Iwan, P.; Stepniak, J.; Karbownik-Lewinska, M. Cumulative protective effect of melatonin and indole-3-propionic acid against
KIO3-induced lipid peroxidation in porcine thyroid. Toxics 2021, 9, 89. https://doi.org/10.3390/toxics9050089.
Int. J. Mol. Sci. 2022, 23, 1222 14 of 21
25. Zhang, L.S.; Davies, S.S. Microbial metabolism of dietary components to bioactive metabolites: Opportunities for new
therapeutic interventions. Genome Med. 2016, 8, 46. https://doi.org/10.1186/s13073-016-0296-x.
26. Hwang, I.K.; Yoo, K.Y.; Li, H.; Park, O.K.; Lee, C.H.; Choi, J.H.; Jeong, Y.G.; Lee, Y.L.; Kim, Y.M.; Kwon, Y.G.; et al. Indole-3-
propionic acid attenuates neuronal damage and oxidative stress in the ischemic hippocampus. J. Neurosci. Res. 2009, 87, 2126–
2137. https://doi.org/10.1002/jnr.22030.
27. Zhao, Z.H.; Xin, F.Z.; Xue, Y.; Hu, Z.; Han, Y.; Ma, F.; Zhou, D.; Liu, X.L.; Cui, A.; Liu, Z.; et al. Indole-3-propionic acid inhibits
gut dysbiosis and endotoxin leakage to attenuate steatohepatitis in rats. Exp. Mol. Med. 2019, 51, 1–14.
https://doi.org/10.1038/s12276-019-0304-5.
28. Alexeev, E.E.; Lanis, J.M.; Kao, D.J.; Campbell, E.L.; Kelly, C.J.; Battista, K.D.; Gerich, M.E.; Jenkins, B.R.; Walk, S.T.; Kominsky,
D.J.; et al. Microbiota-derived indole metabolites promote human and murine intestinal homeostasis through regulation of
interleukin-10 receptor. Am. J. Pathol. 2018, 188, 1183–1194. https://doi.org/10.1016/j.ajpath.2018.01.011.
29. de Mello, V.D.; Paananen, J.; Lindstrom, J.; Lankinen, M.A.; Shi, L.; Kuusisto, J.; Pihlajamaki, J.; Auriola, S.; Lehtonen, M.;
Rolandsson, O.; et al. Indolepropionic acid and novel lipid metabolites are associated with a lower risk of type 2 diabetes in the
Finnish Diabetes Prevention Study. Sci. Rep. 2017, 7, 46337. https://doi.org/10.1038/srep46337.
30. Ruebel, M.L.; Piccolo, B.D.; Mercer, K.E.; Pack, L.; Moutos, D.; Shankar, K.; Andres, A. Obesity leads to distinct metabolomic
signatures in follicular fluid of women undergoing in vitro fertilization. Am. J. Physiol. Endocrinol. Metab. 2019, 316, E383–E396.
https://doi.org/10.1152/ajpendo.00401.2018.
31. Magnusdottir, S.; Ravcheev, D.; de Crecy-Lagard, V.; Thiele, I. Systematic genome assessment of B-vitamin biosynthesis suggests
co-operation among gut microbes. Front. Genet. 2015, 6, 148. https://doi.org/10.3389/fgene.2015.00148.
32. Wang, Z.; Zhao, Y. Gut microbiota derived metabolites in cardiovascular health and disease. Protein Cell 2018, 9, 416–431.
https://doi.org/10.1007/s13238-018-0549-0.
33. Vitek, L.; Majer, F.; Muchova, L.; Zelenka, J.; Jiraskova, A.; Branny, P.; Malina, J.; Ubik, K. Identification of bilirubin reduction
products formed by Clostridium perfringens isolated from human neonatal fecal flora. J. Chromatogr. B Analyt. Technol. Biomed.
Life Sci. 2006, 833, 149–157. https://doi.org/10.1016/j.jchromb.2006.01.032.
34. Ramirez-Perez, O.; Cruz-Ramon, V.; Chinchilla-Lopez, P.; Mendez-Sanchez, N. The Role of the Gut Microbiota in Bile Acid
Metabolism. Ann. Hepatol. 2017, 16, s15–s20. https://doi.org/10.5604/01.3001.0010.5494.
35. Adak, A.; Khan, M.R. An insight into gut microbiota and its functionalities. Cell. Mol. Life. Sci. 2019, 76, 473–493.
https://doi.org/10.1007/s00018-018-2943-4.
36. Wang, B.; Kong, Q.; Li, X.; Zhao, J.; Zhang, H.; Chen, W.; Wang, G. A high-fat diet increases gut microbiota biodiversity and
energy expenditure due to nutrient difference. Nutrients 2020, 12, 3197. https://doi.org/10.3390/nu12103197.
37. James, K.R.; Gomes, T.; Elmentaite, R.; Kumar, N.; Gulliver, E.L.; King, H.W.; Stares, M.D.; Bareham, B.R.; Ferdinand, J.R.;
Petrova, V.N.; et al. Distinct microbial and immune niches of the human colon. Nat. Immunol. 2020, 21, 343–353.
https://doi.org/10.1038/s41590-020-0602-z.
38. Bartoli, R.; Boix, J.; Odena, G.; De la Ossa, N.D.; de Vega, V.M.; Lorenzo-Zuniga, V. Colonoscopy in rats: An endoscopic,
histological and tomographic study. World J. Gastrointest. Endosc. 2013, 5, 226–230. https://doi.org/10.4253/wjge.v5.i5.226.
39. Koruda, M.J.; Rolandelli, R.H.; Bliss, D.Z.; Hastings, J.; Rombeau, J.L.; Settle, R.G. Parenteral nutrition supplemented with short-
chain fatty acids: Effect on the small-bowel mucosa in normal rats. Am. J. Clin. Nutr. 1990, 51, 685–689.
https://doi.org/10.1093/ajcn/51.4.685.
40. Medina, V.; Afonso, J.J.; Alvarez-Arguelles, H.; Hernandez, C.; Gonzalez, F. Sodium butyrate inhibits carcinoma development
in a 1,2-dimethylhydrazine-induced rat colon cancer. JPEN J. Parenter Enteral. Nutr. 1998, 22, 14–17.
https://doi.org/10.1177/014860719802200114.
41. Konopelski, P.; Konop, M.; Perlejewski, K.; Bukowska-Osko, I.; Radkowski, M.; Onyszkiewicz, M.; Jaworska, K.; Mogilnicka, I.;
Samborowska, E.; Ufnal, M. Genetically determined hypertensive phenotype affects gut microbiota composition, but not vice
versa. J. Hypertens. 2021, 39, 1790–1799. https://doi.org/10.1097/HJH.0000000000002864.
42. Morris, G.P.; Beck, P.L.; Herridge, M.S.; Depew, W.T.; Szewczuk, M.R.; Wallace, J.L. Hapten-induced model of chronic
inflammation and ulceration in the rat colon. Gastroenterology 1989, 96, 795–803.
43. Vega-Bautista, A.; de la Garza, M.; Carrero, J.C.; Campos-Rodriguez, R.; Godinez-Victoria, M.; Drago-Serrano, M.E. The Impact
of Lactoferrin on the Growth of Intestinal Inhabitant Bacteria. Int. J. Mol. Sci. 2019, 20, 4707. https://doi.org/10.3390/ijms20194707.
44. Jaworska, K.; Konop, M.; Bielinska, K.; Hutsch, T.; Dziekiewicz, M.; Banaszkiewicz, A.; Ufnal, M. Inflammatory bowel disease is
associated with increased gut-to-blood penetration of short-chain fatty acids: A new, non-invasive marker of a functional
intestinal lesion. Exp. Physiol. 2019, 104, 1226–1236. https://doi.org/10.1113/EP087773.
45. Jaworska, K.; Huc, T.; Samborowska, E.; Dobrowolski, L.; Bielinska, K.; Gawlak, M.; Ufnal, M. Hypertension in rats is associated
with an increased permeability of the colon to TMA, a gut bacteria metabolite. PLoS ONE 2017, 12, e0189310.
https://doi.org/10.1371/journal.pone.0189310.
46. Ueland, P.M. Choline and betaine in health and disease. J. Inherit. Metab. Dis. 2011, 34, 3–15. https://doi.org/10.1007/s10545-010-
9088-4.
47. Wiedeman, A.M.; Barr, S.I.; Green, T.J.; Xu, Z.; Innis, S.M.; Kitts, D.D. Dietary choline intake: Current state of knowledge across
the life cycle. Nutrients 2018, 10, 1513. https://doi.org/10.3390/nu10101513.
48. Zeisel, S.H.; Warrier, M. Trimethylamine N-Oxide, the microbiome, and heart and kidney disease. Annu. Rev. Nutr. 2017, 37,
157–181. https://doi.org/10.1146/annurev-nutr-071816-064732.
Int. J. Mol. Sci. 2022, 23, 1222 15 of 21
49. Jaworska, K.; Bielinska, K.; Gawrys-Kopczynska, M.; Ufnal, M. TMA (trimethylamine), but not its oxide TMAO (trimethylamine-
oxide), exerts haemodynamic effects: Implications for interpretation of cardiovascular actions of gut microbiome. Cardiovasc. Res.
2019, 115, 1948–1949. https://doi.org/10.1093/cvr/cvz231.
50. Srinivasa, S.; Fitch, K.V.; Lo, J.; Kadar, H.; Knight, R.; Wong, K.; Abbara, S.; Gauguier, D.; Capeau, J.; Boccara, F.; et al. Plaque
burden in HIV-infected patients is associated with serum intestinal microbiota-generated trimethylamine. AIDS 2015, 29, 443–
452. https://doi.org/10.1097/QAD.0000000000000565.
51. McConnell, H.W.; Mitchell, S.C.; Smith, R.L.; Brewster, M. Trimethylaminuria associated with seizures and behavioural
disturbance: A case report. Seizure 1997, 6, 317–321. https://doi.org/10.1016/s1059-1311(97)80080-5.
52. Agus, A.; Clement, K.; Sokol, H. Gut microbiota-derived metabolites as central regulators in metabolic disorders. Gut 2021, 70,
1174–1182. https://doi.org/10.1136/gutjnl-2020-323071.
53. Nowinski, A.; Ufnal, M. Gut bacteria-derived molecules as mediators and markers in cardiovascular diseases. The role of the
gut-blood barrier. Kardiol. Pol. 2018, 76, 320–327. https://doi.org/10.5603/KP.a2017.0204.
54. Tang, W.H.; Wang, Z.; Levison, B.S.; Koeth, R.A.; Britt, E.B.; Fu, X.; Wu, Y.; Hazen, S.L. Intestinal microbial metabolism of
phosphatidylcholine and cardiovascular risk. N. Engl. J. Med. 2013, 368, 1575–1584. https://doi.org/10.1056/NEJMoa1109400.
55. Gawrys-Kopczynska, M.; Konop, M.; Maksymiuk, K.; Kraszewska, K.; Derzsi, L.; Sozanski, K.; Holyst, R.; Pilz, M.; Samborowska,
E.; Dobrowolski, L.; et al. TMAO, a seafood-derived molecule, produces diuresis and reduces mortality in heart failure rats. Elife
2020, 9, 57028. https://doi.org/10.7554/eLife.57028.
56. Yancey, P.H.; Siebenaller, J.F. Co-evolution of proteins and solutions: Protein adaptation versus cytoprotective micromolecules
and their roles in marine organisms. J. Exp. Biol. 2015, 218, 1880–1896. https://doi.org/10.1242/jeb.114355.
57. Tomasova, L.; Grman, M.; Ondrias, K.; Ufnal, M. The impact of gut microbiota metabolites on cellular bioenergetics and
cardiometabolic health. Nutr. Metab. 2021, 18, 72. https://doi.org/10.1186/s12986-021-00598-5.
58. Tomasova, L.; Dobrowolski, L.; Jurkowska, H.; Wrobel, M.; Huc, T.; Ondrias, K.; Ostaszewski, R.; Ufnal, M. Intracolonic
hydrogen sulfide lowers blood pressure in rats. Nitric Oxide 2016, 60, 50–58. https://doi.org/10.1016/j.niox.2016.09.007.
59. Al-Magableh, M.R.; Kemp-Harper, B.K.; Hart, J.L. Hydrogen sulfide treatment reduces blood pressure and oxidative stress in
angiotensin II-induced hypertensive mice. Hypertens. Res. 2015, 38, 13–20. https://doi.org/10.1038/hr.2014.125.
60. Zoccali, C.; Catalano, C.; Rastelli, S. Blood pressure control: Hydrogen sulfide, a new gasotransmitter, takes stage. Nephrol. Dial.
Transplant. 2009, 24, 1394–1396. https://doi.org/10.1093/ndt/gfp053.
61. Blachier, F.; Beaumont, M.; Kim, E. Cysteine-derived hydrogen sulfide and gut health: A matter of endogenous or bacterial origin.
Curr. Opin. Clin. Nutr. Metab. Care 2019, 22, 68–75. https://doi.org/10.1097/MCO.0000000000000526.
62. Modis, K.; Wolanska, K.; Vozdek, R. Hydrogen sulfide in cell signaling, signal transduction, cellular bioenergetics and
physiology in C. elegans. Gen. Physiol. Biophys. 2013, 32, 1–22. https://doi.org/10.4149/gpb_2013001.
63. Koh, A.; De Vadder, F.; Kovatcheva-Datchary, P.; Backhed, F. From dietary fiber to host physiology: short-chain fatty acids as
key bacterial metabolites. Cell 2016, 165, 1332–1345. https://doi.org/10.1016/j.cell.2016.05.041.
64. Al-Lahham, S.; Rezaee, F. Propionic acid counteracts the inflammation of human subcutaneous adipose tissue: A new avenue
for drug development. Daru 2019, 27, 645–652. https://doi.org/10.1007/s40199-019-00294-z.
65. Kondo, S.; Tayama, K.; Tsukamoto, Y.; Ikeda, K.; Yamori, Y. Antihypertensive effects of acetic acid and vinegar on spontaneously
hypertensive rats. Biosci. Biotechnol. Biochem. 2001, 65, 2690–2694. https://doi.org/10.1271/bbb.65.2690.
66. Natarajan, N.; Hori, D.; Flavahan, S.; Steppan, J.; Flavahan, N.A.; Berkowitz, D.E.; Pluznick, J.L. Microbial short chain fatty acid
metabolites lower blood pressure via endothelial G protein-coupled receptor 41. Physiol. Genom. 2016, 48, 826–834.
https://doi.org/10.1152/physiolgenomics.00089.2016.
67. Demigne, C.; Morand, C.; Levrat, M.A.; Besson, C.; Moundras, C.; Remesy, C. Effect of propionate on fatty acid and cholesterol
synthesis and on acetate metabolism in isolated rat hepatocytes. Br. J. Nutr. 1995, 74, 209–219. https://doi.org/10.1079/bjn19950124.
68. Robles-Vera, I.; Toral, M.; de la Visitacion, N.; Aguilera-Sanchez, N.; Redondo, J.M.; Duarte, J. Protective effects of short-chain
fatty acids on endothelial dysfunction induced by angiotensin II. Front. Physiol. 2020, 11, 277.
https://doi.org/10.3389/fphys.2020.00277.
69. Knock, G.; Psaroudakis, D.; Abbot, S.; Aaronson, P.I. Propionate-induced relaxation in rat mesenteric arteries: A role for
endothelium-derived hyperpolarising factor. J. Physiol. 2002, 538, 879–890. https://doi.org/10.1113/jphysiol.2001.013105.
70. Mortensen, F.V.; Nielsen, H.; Mulvany, M.J.; Hessov, I. Short chain fatty acids dilate isolated human colonic resistance arteries.
Gut 1990, 31, 1391–1394. https://doi.org/10.1136/gut.31.12.1391.
71. Hulsmann, W.C. Coronary vasodilation by fatty acids. Basic Res. Cardiol. 1976, 71, 179–191. https://doi.org/10.1007/BF01927870.
72. Nutting, C.W.; Islam, S.; Daugirdas, J.T. Vasorelaxant effects of short chain fatty acid salts in rat caudal artery. Am. J. Physiol.
1991, 261, H561–H567. https://doi.org/10.1152/ajpheart.1991.261.2.H561.
73. Agus, A.; Planchais, J.; Sokol, H. Gut Microbiota Regulation of Tryptophan Metabolism in Health and Disease. Cell Host Microbe
2018, 23, 716–724. https://doi.org/10.1016/j.chom.2018.05.003.
74. Kanova, M.; Kohout, P. Tryptophan: A Unique Role in the Critically Ill. Int. J. Mol. Sci. 2021, 22, 11714.
https://doi.org/10.3390/ijms222111714.
75. Badawy, A.A. Kynurenine pathway of tryptophan metabolism: Regulatory and functional aspects. Int. J. Tryptophan Res. 2017,
10, 1178646917691938. https://doi.org/10.1177/1178646917691938.
76. Ball, H.J.; Jusof, F.F.; Bakmiwewa, S.M.; Hunt, N.H.; Yuasa, H.J. Tryptophan-catabolizing enzymes-party of three. Front. Immunol.
2014, 5, 485. https://doi.org/10.3389/fimmu.2014.00485.
Int. J. Mol. Sci. 2022, 23, 1222 16 of 21
77. Jones, L.A.; Sun, E.W.; Martin, A.M.; Keating, D.J. The ever-changing roles of serotonin. Int. J. Biochem. Cell Biol. 2020, 125, 105776.
https://doi.org/10.1016/j.biocel.2020.105776.
78. Fernstrom, J.D. A Perspective on the safety of supplemental tryptophan based on its metabolic fates. J. Nutr. 2016, 146, 2601S–
2608S. https://doi.org/10.3945/jn.115.228643.
79. Zhao, D.; Yu, Y.; Shen, Y.; Liu, Q.; Zhao, Z.; Sharma, R.; Reiter, R.J. Melatonin synthesis and function: Evolutionary history in
animals and plants. Front. Endocrinol. 2019, 10, 249. https://doi.org/10.3389/fendo.2019.00249.
80. Deacon, A.C. The measurement of 5-hydroxyindoleacetic acid in urine. Ann. Clin. Biochem. 1994, 31, 215–232.
https://doi.org/10.1177/000456329403100302.
81. Wedin, M.; Mehta, S.; Angeras-Kraftling, J.; Wallin, G.; Daskalakis, K. The Role of Serum 5-HIAA as a Predictor of progression
and an alternative to 24-h urine 5-HIAA in well-differentiated neuroendocrine neoplasms. Biology 2021, 10, 76.
https://doi.org/10.3390/biology10020076.
82. Rothhammer, V.; Mascanfroni, I.D.; Bunse, L.; Takenaka, M.C.; Kenison, J.E.; Mayo, L.; Chao, C.C.; Patel, B.; Yan, R.; Blain, M.;
et al. Type I interferons and microbial metabolites of tryptophan modulate astrocyte activity and central nervous system
inflammation via the aryl hydrocarbon receptor. Nat. Med. 2016, 22, 586–597. https://doi.org/10.1038/nm.4106.
83. Li, Y.; Xu, W.; Zhang, F.; Zhong, S.; Sun, Y.; Huo, J.; Zhu, J.; Wu, C. The Gut Microbiota-Produced Indole-3-Propionic Acid
Confers the antihyperlipidemic effect of mulberry-derived 1-Deoxynojirimycin. mSystems 2020, 5, e00313-20.
https://doi.org/10.1128/mSystems.00313-20.
84. Mercer, K.E.; Yeruva, L.; Pack, L.; Graham, J.L.; Stanhope, K.L.; Chintapalli, S.V.; Wankhade, U.D.; Shankar, K.; Havel, P.J.;
Adams, S.H.; et al. Xenometabolite signatures in the UC Davis type 2 diabetes mellitus rat model revealed using a metabolomics
platform enriched with microbe-derived metabolites. Am. J. Physiol. Gastrointest. Liver. Physiol. 2020, 319, G157–G169.
https://doi.org/10.1152/ajpgi.00105.2020.
85. Wikoff, W.R.; Anfora, A.T.; Liu, J.; Schultz, P.G.; Lesley, S.A.; Peters, E.C.; Siuzdak, G. Metabolomics analysis reveals large effects
of gut microflora on mammalian blood metabolites. Proc. Natl. Acad. Sci. USA 2009, 106, 3698–3703.
https://doi.org/10.1073/pnas.0812874106.
86. 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. https://doi.org/10.1038/nature24661.
87. Rosas, H.D.; Doros, G.; Bhasin, S.; Thomas, B.; Gevorkian, S.; Malarick, K.; Matson, W.; Hersch, S.M. A systems-level
‘’misunderstanding’’: The plasma metabolome in Huntington’s disease. Ann. Clin. Transl. Neurol. 2015, 2, 756–768.
https://doi.org/10.1002/acn3.214.
88. Elsden, S.R.; Hilton, M.G.; Waller, J.M. The end products of the metabolism of aromatic amino acids by Clostridia. Arch. Microbiol.
1976, 107, 283–288. https://doi.org/10.1007/BF00425340.
89. Roager, H.M.; Licht, T.R. Microbial tryptophan catabolites in health and disease. Nat. Commun. 2018, 9, 3294.
https://doi.org/10.1038/s41467-018-05470-4.
90. Parthasarathy, A.; Cross, P.J.; Dobson, R.C.J.; Adams, L.E.; Savka, M.A.; Hudson, A.O. A Three-Ring Circus: Metabolism of the
three proteogenic aromatic amino acids and their role in the health of plants and animals. Front. Mol. Biosci. 2018, 5, 29.
https://doi.org/10.3389/fmolb.2018.00029.
91. Huang, Y.S.; Ogbechi, J.; Clanchy, F.I.; Williams, R.O.; Stone, T.W. IDO and Kynurenine Metabolites in Peripheral and CNS
Disorders. Front. Immunol. 2020, 11, 388. https://doi.org/10.3389/fimmu.2020.00388.
92. Biernacki, T.; Sandi, D.; Bencsik, K.; Vecsei, L. Kynurenines in the Pathogenesis of Multiple Sclerosis: Therapeutic Perspectives.
Cells 2020, 9, 1564. https://doi.org/10.3390/cells9061564.
93. Barth, H.; Raghuraman, S. Persistent infectious diseases say-IDO. Role of indoleamine-2,3-dioxygenase in disease pathogenesis
and implications for therapy. Crit. Rev. Microbiol. 2014, 40, 360–368. https://doi.org/10.3109/1040841X.2012.742037.
94. Li, X.; Zhou, J.; Fang, M.; Yu, B. Pregnancy immune tolerance at the maternal-fetal interface. Int. Rev. Immunol. 2020, 39, 247–263.
https://doi.org/10.1080/08830185.2020.1777292.
95. Suzuki, Y.; Suda, T.; Asada, K.; Miwa, S.; Suzuki, M.; Fujie, M.; Furuhashi, K.; Nakamura, Y.; Inui, N.; Shirai, T.; et al. Serum
indoleamine 2,3-dioxygenase activity predicts prognosis of pulmonary tuberculosis. Clin. Vaccine Immunol. 2012, 19, 436–442.
https://doi.org/10.1128/CVI.05402-11.
96. Wang, Y.; Liu, H.; McKenzie, G.; Witting, P.K.; Stasch, J.P.; Hahn, M.; Changsirivathanathamrong, D.; Wu, B.J.; Ball, H.J.; Thomas,
S.R.; et al. Kynurenine is an endothelium-derived relaxing factor produced during inflammation. Nat. Med. 2010, 16, 279–285.
https://doi.org/10.1038/nm.2092.
97. Cervenka, I.; Agudelo, L.Z.; Ruas, J.L. Kynurenines: Tryptophan’s metabolites in exercise, inflammation, and mental health.
Science 2017, 357, eaaf9749. https://doi.org/10.1126/science.aaf9794.
98. Munn, D.H.; Mellor, A.L. IDO in the tumor microenvironment: inflammation, counter-regulation, and tolerance. Trends Immunol.
2016, 37, 193–207. https://doi.org/10.1016/j.it.2016.01.002.
99. Lanser, L.; Kink, P.; Egger, E.M.; Willenbacher, W.; Fuchs, D.; Weiss, G.; Kurz, K. Inflammation-induced tryptophan breakdown
is related with anemia, fatigue, and depression in cancer. Front. Immunol. 2020, 11, 249. https://doi.org/10.3389/fimmu.2020.00249.
100. Tsopmo, A.; Diehl-Jones, B.W.; Aluko, R.E.; Kitts, D.D.; Elisia, I.; Friel, J.K. Tryptophan released from mother’s milk has
antioxidant properties. Pediatr. Res. 2009, 66, 614–618. https://doi.org/10.1203/PDR.0b013e3181be9e7e.
Int. J. Mol. Sci. 2022, 23, 1222 17 of 21
101. Bitzer-Quintero, O.K.; Davalos-Marin, A.J.; Ortiz, G.G.; Meza, A.R.; Torres-Mendoza, B.M.; Robles, R.G.; Huerta, V.C.; Beas-
Zarate, C. Antioxidant activity of tryptophan in rats under experimental endotoxic shock. Biomed. Pharmacother. 2010, 64, 77–81.
https://doi.org/10.1016/j.biopha.2009.07.002.
102. Sanchez, A.; Calpena, A.C.; Clares, B. Evaluating the Oxidative Stress in Inflammation: Role of Melatonin. Int. J. Mol. Sci. 2015,
16, 16981–17004. https://doi.org/10.3390/ijms160816981.
103. Yusufu, I.; Ding, K.; Smith, K.; Wankhade, U.D.; Sahay, B.; Patterson, G.T.; Pacholczyk, R.; Adusumilli, S.; Hamrick, M.W.; Hill,
W.D.; et al. A Tryptophan-Deficient Diet Induces Gut Microbiota Dysbiosis and Increases Systemic Inflammation in Aged Mice.
Int. J. Mol. Sci. 2021, 22, 5005. https://doi.org/10.3390/ijms22095005.
104. Bortolato, M.; Frau, R.; Orru, M.; Collu, M.; Mereu, G.; Carta, M.; Fadda, F.; Stancampiano, R. Effects of tryptophan deficiency
on prepulse inhibition of the acoustic startle in rats. Psychopharmacology 2008, 198, 191–200. https://doi.org/10.1007/s00213-008-
1116-9.
105. Konopelski, P.; Konop, M.; Gawrys-Kopczynska, M.; Podsadni, P.; Szczepanska, A.; Ufnal, M. Indole-3-propionic acid, a
tryptophan-derived bacterial metabolite, reduces weight gain in rats. Nutrients 2019, 11, 591. https://doi.org/10.3390/nu11030591.
106. Franklin, M.; Bermudez, I.; Murck, H.; Singewald, N.; Gaburro, S. Sub-chronic dietary tryptophan depletion—An animal model
of depression with improved face and good construct validity. J. Psychiatr. Res. 2012, 46, 239–247.
https://doi.org/10.1016/j.jpsychires.2011.10.003.
107. Ayaso, R.; Ghattas, H.; Abiad, M.; Obeid, O. Meal pattern of male rats maintained on amino acid supplemented diets: The effect
of tryptophan, lysine, arginine, proline and threonine. Nutrients 2014, 6, 2509–2522. https://doi.org/10.3390/nu6072509.
108. Gartner, S.N.; Aidney, F.; Klockars, A.; Prosser, C.; Carpenter, E.A.; Isgrove, K.; Levine, A.S.; Olszewski, P.K. Intragastric
preloads of l-tryptophan reduce ingestive behavior via oxytocinergic neural mechanisms in male mice. Appetite 2018, 125, 278–
286. https://doi.org/10.1016/j.appet.2018.02.015.
109. Ufnal, M.; Skrzypecki, J. Blood borne hormones in a cross-talk between peripheral and brain mechanisms regulating blood
pressure, the role of circumventricular organs. Neuropeptides 2014, 48, 65–73. https://doi.org/10.1016/j.npep.2014.01.003.
110. Toropov, A.L.; Tsirkin, V.I.; Kostyaev, A.A. Combined effects of blood serum as a source of endogenous beta-adrenoceptor-
sensitizing agent and its analogues histidine, tryptophan, tyrosine, mildronat, and preductal. Bull. Exp. Biol. Med. 2011, 151, 84–
87. https://doi.org/10.1007/s10517-011-1265-4.
111. Korotaeva, K.N.; Tsirkin, V.I.; Vyaznikov, V.A. Positive inotropic effect of tyrosine, histidine, and tryptophan in experiments on
isolated human myocardium. Bull. Exp. Biol. Med. 2012, 153, 51–53. https://doi.org/10.1007/s10517-012-1640-9.
112. Wolf, W.A.; Kuhn, D.M. Effects of L-tryptophan on blood pressure in normotensive and hypertensive rats. J. Pharmacol. Exp.
Ther. 1984, 230, 324–329.
113. Bertaccini, G.; Nobili, M.B. Effect of L-tryptophan on diuresis and 5-hydroxyindoleacetic acid excretion in the rat. Br. J. Pharmacol.
Chemother. 1961, 17, 519–525. https://doi.org/10.1111/j.1476-5381.1961.tb01138.x.
114. Reuther, E.; Weber, H.J.; Herken, H. Studies on sodium ion retention and antidiuretic effects after administration of L-tryptophan
to rats. Naunyn. Schmiedebergs Arch. Pharmacol. 1977, 297, 213–217. https://doi.org/10.1007/BF00509263.
115. Feltkamp, H.; Meurer, K.A.; Godehardt, E. Tryptophan-induced lowering of blood pressure and changes of serotonin uptake by
platelets in patients with essential hypertension. Klin. Wochenschr. 1984, 62, 1115–1119. https://doi.org/10.1007/BF01782468.
116. Liu, Q.; Yu, Z.; Tian, F.; Zhao, J.; Zhang, H.; Zhai, Q.; Chen, W. Surface components and metabolites of probiotics for regulation
of intestinal epithelial barrier. Microb. Cell. Fact. 2020, 19, 23. https://doi.org/10.1186/s12934-020-1289-4.
117. Galligan, J.J. Beneficial actions of microbiota-derived tryptophan metabolites. Neurogastroenterol. Motil. 2018, 30, e13283.
https://doi.org/10.1111/nmo.13283.
118. Wu, Y.; Li, J.; Ding, W.; Ruan, Z.; Zhang, L. Enhanced Intestinal Barriers by Puerarin in Combination with Tryptophan. J. Agric.
Food Chem. 2021, 69, 15575–15584. https://doi.org/10.1021/acs.jafc.1c05830.
119. Sivaprakasam, S.; Bhutia, Y.D.; Ramachandran, S.; Ganapathy, V. Cell-Surface and Nuclear Receptors in the Colon as Targets
for Bacterial Metabolites and Its Relevance to Colon Health. Nutrients 2017, 9, 856. https://doi.org/10.3390/nu9080856.
120. Dutta, M.; Lim, J.J.; Cui, J.Y. PXR and the gut-liver axis: A recent update. Drug Metab. Dispos. 2021, 50, 000415.
https://doi.org/10.1124/dmd.121.000415.
121. Venkatesh, M.; Mukherjee, S.; Wang, H.; Li, H.; Sun, K.; Benechet, A.P.; Qiu, Z.; Maher, L.; Redinbo, M.R.; Phillips, R.S.; et al.
Symbiotic bacterial metabolites regulate gastrointestinal barrier function via the xenobiotic sensor PXR and Toll-like receptor 4.
Immunity 2014, 41, 296–310. https://doi.org/10.1016/j.immuni.2014.06.014.
122. Lee, D.M.; Ecton, K.E.; Trikha, S.R.J.; Wrigley, S.D.; Thomas, K.N.; Battson, M.L.; Wei, Y.; Johnson, S.A.; Weir, T.L.; Gentile, C.L.
Microbial metabolite indole-3-propionic acid supplementation does not protect mice from the cardiometabolic consequences of
a Western diet. Am. J. Physiol. Gastrointest. Liver. Physiol. 2020, 319, G51–G62. https://doi.org/10.1152/ajpgi.00375.2019.
123. Jennis, M.; Cavanaugh, C.R.; Leo, G.C.; Mabus, J.R.; Lenhard, J.; Hornby, P.J. Microbiota-derived tryptophan indoles increase
after gastric bypass surgery and reduce intestinal permeability in vitro and in vivo. Neurogastroenterol. Motil. 2018, 30, e13178.
https://doi.org/10.1111/nmo.13178.
124. Niwa, T. Indoxyl sulfate is a nephro-vascular toxin. J. Ren. Nutr. 2010, 20, S2–S6. https://doi.org/10.1053/j.jrn.2010.05.002.
125. Nakano, T.; Katsuki, S.; Chen, M.; Decano, J.L.; Halu, A.; Lee, L.H.; Pestana, D.V.S.; Kum, A.S.T.; Kuromoto, R.K.; Golden, W.S.;
et al. Uremic toxin indoxyl sulfate promotes proinflammatory macrophage activation via the interplay of OATP2B1 and Dll4-
notch signaling. Circulation 2019, 139, 78–96. https://doi.org/10.1161/CIRCULATIONAHA.118.034588.
Int. J. Mol. Sci. 2022, 23, 1222 18 of 21
126. Tanaka, S.; Watanabe, H.; Nakano, T.; Imafuku, T.; Kato, H.; Tokumaru, K.; Arimura, N.; Enoki, Y.; Maeda, H.; Tanaka, M.; et al.
Indoxyl sulfate contributes to adipose tissue inflammation through the activation of NADPH oxidase. Toxins 2020, 12, 502.
https://doi.org/10.3390/toxins12080502.
127. Yisireyili, M.; Takeshita, K.; Saito, S.; Murohara, T.; Niwa, T. Indole-3-propionic acid suppresses indoxyl sulfate-induced
expression of fibrotic and inflammatory genes in proximal tubular cells. Nagoya J. Med. Sci. 2017, 79, 477–486.
https://doi.org/10.18999/nagjms.79.4.477.
128. Poeggeler, B.; Pappolla, M.A.; Hardeland, R.; Rassoulpour, A.; Hodgkins, P.S.; Guidetti, P.; Schwarcz, R. Indole-3-propionate: A
potent hydroxyl radical scavenger in rat brain. Brain Res. 1999, 815, 382–388. https://doi.org/10.1016/s0006-8993(98)01027-0.
129. Hardeland, R.; Zsizsik, B.K.; Poeggeler, B.; Fuhrberg, B.; Holst, S.; Coto-Montes, A. Indole-3-pyruvic and -propionic acids,
kynurenic acid, and related metabolites as luminophores and free-radical scavengers. Adv. Exp. Med. Biol. 1999, 467, 389–395.
https://doi.org/10.1007/978-1-4615-4709-9_49.
130. Karbownik, M.; Stasiak, M.; Zasada, K.; Zygmunt, A.; Lewinski, A. Comparison of potential protective effects of melatonin,
indole-3-propionic acid, and propylthiouracil against lipid peroxidation caused by potassium bromate in the thyroid gland. J.
Cell Biochem. 2005, 95, 131–138. https://doi.org/10.1002/jcb.20404.
131. Karbownik, M.; Stasiak, M.; Zygmunt, A.; Zasada, K.; Lewinski, A. Protective effects of melatonin and indole-3-propionic acid
against lipid peroxidation, caused by potassium bromate in the rat kidney. Cell Biochem. Funct. 2006, 24, 483–489.
https://doi.org/10.1002/cbf.1321.
132. Stasiak, M.; Zasada, K.; Lewinski, A.; Karbownik-Lewinska, M. Melatonin restores the basal level of lipid peroxidation in rat
tissues exposed to potassium bromate in vitro. Neuro. Endocrinol. Lett. 2010, 31, 363–369.
133. Rynkowska, A.; Stepniak, J.; Karbownik-Lewinska, M. Melatonin and Indole-3-Propionic acid reduce oxidative damage to
membrane lipids induced by high iron concentrations in porcine skin. Membranes 2021, 11, 571.
https://doi.org/10.3390/membranes11080571.
134. Karbownik, M.; Gitto, E.; Lewinski, A.; Reiter, R.J. Relative efficacies of indole antioxidants in reducing autoxidation and iron-
induced lipid peroxidation in hamster testes. J. Cell. Biochem. 2001, 81, 693–699. https://doi.org/10.1002/jcb.1100.
135. Karbownik, M.; Reiter, R.J.; Garcia, J.J.; Cabrera, J.; Burkhardt, S.; Osuna, C.; Lewinski, A. Indole-3-propionic acid, a melatonin-
related molecule, protects hepatic microsomal membranes from iron-induced oxidative damage: Relevance to cancer reduction.
J. Cell Biochem. 2001, 81, 507–513.
136. Qi, W.; Reiter, R.J.; Tan, D.X.; Manchester, L.C.; Siu, A.W.; Garcia, J.J. Increased levels of oxidatively damaged DNA induced by
chromium(III) and H2O2: Protection by melatonin and related molecules. J. Pineal Res. 2000, 29, 54–61.
https://doi.org/10.1034/j.1600-079x.2000.290108.x.
137. Ortial, S.; Durand, G.; Poeggeler, B.; Polidori, A.; Pappolla, M.A.; Boker, J.; Hardeland, R.; Pucci, B. Fluorinated amphiphilic
amino acid derivatives as antioxidant carriers: A new class of protective agents. J. Med. Chem. 2006, 49, 2812–2820.
https://doi.org/10.1021/jm060027e.
138. Mandelbaum-Shavit, F.; Barak, V.; Saheb-Tamimi, K.; Grossowicz, N. Susceptibility of Legionella pneumophila grown
extracellularly and in human monocytes to indole-3-propionic acid. Antimicrob. Agents Chemother. 1991, 35, 2526–2530.
https://doi.org/10.1128/AAC.35.12.2526.
139. Grossowicz, N. Phytohormones as specific inhibitors of Legionella pneumophila growth. Isr. J. Med. Sci. 1990, 26, 187–190.
140. Chelala, C.A.; Margolin, P. Bactericidal photoproducts in medium containing riboflavin plus aromatic compounds and MnCl2.
Can. J. Microbiol. 1983, 29, 670–675. https://doi.org/10.1139/m83-109.
141. Tuomainen, M.; Lindstrom, J.; Lehtonen, M.; Auriola, S.; Pihlajamaki, J.; Peltonen, M.; Tuomilehto, J.; Uusitupa, M.; de Mello,
V.D.; Hanhineva, K. Associations of serum indolepropionic acid, a gut microbiota metabolite, with type 2 diabetes and low-
grade inflammation in high-risk individuals. Nutr. Diabetes 2018, 8, 35. https://doi.org/10.1038/s41387-018-0046-9.
142. Nystrom, S.; Govender, M.; Yap, S.H.; Kamarulzaman, A.; Rajasuriar, R.; Larsson, M. HIV-infected individuals on ART With
impaired immune recovery have altered plasma metabolite profiles. Open Forum Infect. Dis. 2021, 8, ofab288.
https://doi.org/10.1093/ofid/ofab288.
143. Jeffrey, A.M.; Williams, G.M. Risk assessment of DNA-reactive carcinogens in food. Toxicol. Appl. Pharmacol. 2005, 207, 628–635.
https://doi.org/10.1016/j.taap.2005.03.024.
144. Hecht, S.S. Tobacco carcinogens, their biomarkers and tobacco-induced cancer. Nat. Rev. Cancer 2003, 3, 733–744.
https://doi.org/10.1038/nrc1190.
145. Athreya, K.; Xavier, M.F. Antioxidants in the treatment of cancer. Nutr. Cancer 2017, 69, 1099–1104.
https://doi.org/10.1080/01635581.2017.1362445.
146. Harris, I.S.; DeNicola, G.M. The Complex Interplay between Antioxidants and ROS in cancer. Trends Cell Biol. 2020, 30, 440–451.
https://doi.org/10.1016/j.tcb.2020.03.002.
147. Karbownik, M.; Reiter, R.J.; Cabrera, J.; Garcia, J.J. Comparison of the protective effect of melatonin with other antioxidants in
the hamster kidney model of estradiol-induced DNA damage. Mutat. Res. 2001, 474, 87–92. https://doi.org/10.1016/s0027-
5107(00)00164-0.
148. Owumi, S.E.; Adedara, I.A.; Oyelere, A.K. Indole-3-propionic acid mitigates chlorpyrifos-mediated neurotoxicity by modulating
cholinergic and redox-regulatory systems, inflammatory stress, apoptotic responses and DNA damage in rats. Environ. Toxicol.
Pharmacol. 2022, 89, 103786. https://doi.org/10.1016/j.etap.2021.103786.
Int. J. Mol. Sci. 2022, 23, 1222 19 of 21
149. Karbownik, M.; Garcia, J.J.; Lewinski, A.; Reiter, R.J. Carcinogen-induced, free radical-mediated reduction in microsomal
membrane fluidity: Reversal by indole-3-propionic acid. J. Bioenerg. Biomembr. 2001, 33, 73–78.
https://doi.org/10.1023/a:1005628808688.
150. Tolan, D.; Gandin, V.; Morrison, L.; El-Nahas, A.; Marzano, C.; Montagner, D.; Erxleben, A. Oxidative stress induced by Pt(IV)
Pro-drugs based on the Cisplatin Scaffold and Indole Carboxylic Acids in Axial Position. Sci. Rep. 2016, 6, 29367.
https://doi.org/10.1038/srep29367.
151. Zhang, Z.; Bi, C.; Buac, D.; Fan, Y.; Zhang, X.; Zuo, J.; Zhang, P.; Zhang, N.; Dong, L.; Dou, Q.P. Organic cadmium complexes as
proteasome inhibitors and apoptosis inducers in human breast cancer cells. J. Inorg. Biochem. 2013, 123, 1–10.
https://doi.org/10.1016/j.jinorgbio.2013.02.004.
152. Tabassum, S.; Zaki, M.; Ahmad, M.; Afzal, M.; Srivastav, S.; Srikrishna, S.; Arjmand, F. Synthesis and crystal structure
determination of copper(II)-complex: In vitro DNA and HSA binding, pBR322 plasmid cleavage, cell imaging and cytotoxic
studies. Eur. J. Med. Chem. 2014, 83, 141–154. https://doi.org/10.1016/j.ejmech.2014.06.018.
153. Xiao, H.W.; Cui, M.; Li, Y.; Dong, J.L.; Zhang, S.Q.; Zhu, C.C.; Jiang, M.; Zhu, T.; Wang, B.; Wang, H.C.; et al. Gut microbiota-
derived indole 3-propionic acid protects against radiation toxicity via retaining acyl-CoA-binding protein. Microbiome 2020, 8,
69. https://doi.org/10.1186/s40168-020-00845-6.
154. Dugger, B.N.; Dickson, D.W. Pathology of Neurodegenerative Diseases. Cold Spring Harb. Perspect. Biol. 2017, 9, a028035.
https://doi.org/10.1101/cshperspect.a028035.
155. van der Plas, E.; Schultz, J.L.; Nopoulos, P.C. The Neurodevelopmental Hypothesis of Huntington’s Disease. J. Huntingtons Dis.
2020, 9, 217–229. https://doi.org/10.3233/JHD-200394.
156. Behl, C.; Davis, J.B.; Lesley, R.; Schubert, D. Hydrogen peroxide mediates amyloid beta protein toxicity. Cell 1994, 77, 817–827.
https://doi.org/10.1016/0092-8674(94)90131-7.
157. Pappolla, M.A.; Matsubara, E.; Vidal, R.; Pacheco-Quinto, J.; Poeggeler, B.; Zagorski, M.; Sambamurti, K. Melatonin treatment
enhances abeta lymphatic clearance in a transgenic mouse model of amyloidosis. Curr. Alzheimer Res. 2018, 15, 637–642.
https://doi.org/10.2174/1567205015666180411092551.
158. Bendheim, P.E.; Poeggeler, B.; Neria, E.; Ziv, V.; Pappolla, M.A.; Chain, D.G. Development of indole-3-propionic acid (OXIGON)
for Alzheimer's disease. J. Mol. Neurosci. 2002, 19, 213–217. https://doi.org/10.1007/s12031-002-0036-0.
159. Chyan, Y.J.; Poeggeler, B.; Omar, R.A.; Chain, D.G.; Frangione, B.; Ghiso, J.; Pappolla, M.A. Potent neuroprotective properties
against the Alzheimer beta-amyloid by an endogenous melatonin-related indole structure, indole-3-propionic acid. J. Biol. Chem.
1999, 274, 21937–21942. https://doi.org/10.1074/jbc.274.31.21937.
160. Dragicevic, N.; Copes, N.; O'Neal-Moffitt, G.; Jin, J.; Buzzeo, R.; Mamcarz, M.; Tan, J.; Cao, C.; Olcese, J.M.; Arendash, G.W.; et
al. Melatonin treatment restores mitochondrial function in Alzheimer’s mice: A mitochondrial protective role of melatonin
membrane receptor signaling. J. Pineal Res. 2011, 51, 75–86. https://doi.org/10.1111/j.1600-079X.2011.00864.x.
161. Huang, Y.L.; Lin, C.H.; Tsai, T.H.; Huang, C.H.; Li, J.L.; Chen, L.K.; Li, C.H.; Tsai, T.F.; Wang, P.N. Discovery of a metabolic
signature predisposing high risk patients with mild cognitive impairment to converting to Alzheimer’s disease. Int. J. Mol. Sci.
2021, 22, 10903. https://doi.org/10.3390/ijms222010903.
162. Morshedi, D.; Rezaei-Ghaleh, N.; Ebrahim-Habibi, A.; Ahmadian, S.; Nemat-Gorgani, M. Inhibition of amyloid fibrillation of
lysozyme by indole derivatives--possible mechanism of action. FEBS J. 2007, 274, 6415–6425. https://doi.org/10.1111/j.1742-
4658.2007.06158.x.
163. Mimori, S.; Kawada, K.; Saito, R.; Takahashi, M.; Mizoi, K.; Okuma, Y.; Hosokawa, M.; Kanzaki, T. Indole-3-propionic acid has
chemical chaperone activity and suppresses endoplasmic reticulum stress-induced neuronal cell death. Biochem. Biophys. Res.
Commun. 2019, 517, 623–628. https://doi.org/10.1016/j.bbrc.2019.07.074.
164. Mangalam, A.; Poisson, L.; Nemutlu, E.; Datta, I.; Denic, A.; Dzeja, P.; Rodriguez, M.; Rattan, R.; Giri, S. Profile of circulatory
metabolites in a relapsing-remitting animal model of multiple sclerosis using global metabolomics. J. Clin. Cell. Immunol. 2013,
4, 10.4172/2155-9899.1000150. https://doi.org/10.4172/2155–9899.1000150.
165. Gaetani, L.; Boscaro, F.; Pieraccini, G.; Calabresi, P.; Romani, L.; Di Filippo, M.; Zelante, T. Host and Microbial Tryptophan
metabolic profiling in multiple sclerosis. Front. Immunol. 2020, 11, 157. https://doi.org/10.3389/fimmu.2020.00157.
166. Cavero, I.; Lefevre-Borg, F.; Roach, A.G. Effects of mianserin, desipramine and maprotiline on blood pressure responses evoked
by acetylcholine, histamine and 5-hydroxytryptamine in rats. Br. J. Pharmacol. 1981, 74, 143–148. https://doi.org/10.1111/j.1476-
5381.1981.tb09966.x.
167. Gesper, M.; Nonnast, A.B.H.; Kumowski, N.; Stoehr, R.; Schuett, K.; Marx, N.; Kappel, B.A. Gut-derived metabolite Indole-3-
Propionic acid modulates mitochondrial function in cardiomyocytes and alters cardiac function. Front. Med. 2021, 8, 648259.
https://doi.org/10.3389/fmed.2021.648259.
168. Zhu, C.; Sawrey-Kubicek, L.; Beals, E.; Rhodes, C.H.; Houts, H.E.; Sacchi, R.; Zivkovic, A.M. Human gut microbiome composition
and tryptophan metabolites were changed differently by fast food and Mediterranean diet in 4 days: A pilot study. Nutr. Res.
2020, 77, 62–72. https://doi.org/10.1016/j.nutres.2020.03.005.
169. Du, H.; Rao, Y.; Liu, R.; Deng, K.; Guan, Y.; Luo, D.; Mao, Q.; Yu, J.; Bo, T.; Fan, Z.; et al. Proteomics and metabolomics analyses
reveal the full spectrum of inflammatory and lipid metabolic abnormalities in dyslipidemia. Biomed. Chromatogr. 2021, 35, e5183.
https://doi.org/10.1002/bmc.5183.
Int. J. Mol. Sci. 2022, 23, 1222 20 of 21
170. Cason, C.A.; Dolan, K.T.; Sharma, G.; Tao, M.; Kulkarni, R.; Helenowski, I.B.; Doane, B.M.; Avram, M.J.; McDermott, M.M.;
Chang, E.B.; et al. Plasma microbiome-modulated indole- and phenyl-derived metabolites associate with advanced
atherosclerosis and postoperative outcomes. J. Vasc. Surg. 2018, 68, 1552–1562. https://doi.org/10.1016/j.jvs.2017.09.029.
171. Zhong, V.W.; Van Horn, L.; Greenland, P.; Carnethon, M.R.; Ning, H.; Wilkins, J.T.; Lloyd-Jones, D.M.; Allen, N.B. Associations
of processed meat, unprocessed red meat, poultry, or fish intake with incident cardiovascular disease and all-cause mortality.
JAMA Intern. Med. 2020, 180, 503–512. https://doi.org/10.1001/jamainternmed.2019.6969.
172. Wang, Z.; Bergeron, N.; Levison, B.S.; Li, X.S.; Chiu, S.; Jia, X.; Koeth, R.A.; Li, L.; Wu, Y.; Tang, W.H.W.; et al. Impact of chronic
dietary red meat, white meat, or non-meat protein on trimethylamine N-oxide metabolism and renal excretion in healthy men
and women. Eur. Heart J. 2019, 40, 583–594. https://doi.org/10.1093/eurheartj/ehy799.
173. Wu, W.; Zhang, L.; Xia, B.; Tang, S.; Liu, L.; Xie, J.; Zhang, H. Bioregional alterations in gut microbiome contribute to the plasma
metabolomic changes in pigs fed with inulin. Microorganisms 2020, 8, 111. https://doi.org/10.3390/microorganisms8010111.
174. Behera, J.; Ison, J.; Voor, M.J.; Tyagi, N. Probiotics stimulate bone formation in obese mice via histone methylations. Theranostics
2021, 11, 8605–8623. https://doi.org/10.7150/thno.63749.
175. Mach, F.; Baigent, C.; Catapano, A.L.; Koskinas, K.C.; Casula, M.; Badimon, L.; Chapman, M.J.; De Backer, G.G.; Delgado, V.;
Ference, B.A.; et al. 2019 ESC/EAS Guidelines for the management of dyslipidaemias: Lipid modification to reduce
cardiovascular risk. Eur. Heart J. 2020, 41, 111–188. https://doi.org/10.1093/eurheartj/ehz455.
176. Russo, G.; Piscitelli, P.; Giandalia, A.; Viazzi, F.; Pontremoli, R.; Fioretto, P.; De Cosmo, S. Atherogenic dyslipidemia and diabetic
nephropathy. J. Nephrol. 2020, 33, 1001–1008. https://doi.org/10.1007/s40620-020-00739-8.
177. Rinella, M.E. Nonalcoholic fatty liver disease: A systematic review. JAMA 2015, 313, 2263–2273.
https://doi.org/10.1001/jama.2015.5370.
178. Sehgal, R.; Ilha, M.; Vaittinen, M.; Kaminska, D.; Mannisto, V.; Karja, 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. https://doi.org/10.3390/nu13103509.
179. Liu, F.; Sun, C.; Chen, Y.; Du, F.; Yang, Y.; Wu, G. Indole-3-propionic Acid-aggravated CCl4-induced Liver Fibrosis via the TGF-
beta1/Smads Signaling Pathway. J. Clin. Transl. Hepatol. 2021, 9, 917–930. https://doi.org/10.14218/JCTH.2021.00032.
180. Meldrum, D.R.; Morris, M.A.; Gambone, J.C. Obesity pandemic: Causes, consequences, and solutions-but do we have the will?
Fertil. Steril. 2017, 107, 833–839. https://doi.org/10.1016/j.fertnstert.2017.02.104.
181. Cooke, A.A.; Connaughton, R.M.; Lyons, C.L.; McMorrow, A.M.; Roche, H.M. Fatty acids and chronic low grade inflammation
associated with obesity and the metabolic syndrome. Eur. J. Pharmacol. 2016, 785, 207–214.
https://doi.org/10.1016/j.ejphar.2016.04.021.
182. Zhao, L.; Wu, H.; Qiu, M.; Sun, W.; Wei, R.; Zheng, X.; Yang, Y.; Xin, X.; Zou, H.; Chen, T.; et al. Metabolic Signatures of Kidney
Yang Deficiency Syndrome and Protective Effects of Two Herbal Extracts in Rats Using GC/TOF MS. Evid. Based. Complement.
Alternat. Med. 2013, 2013, 540957. https://doi.org/10.1155/2013/540957.
183. Bornfeldt, K.E.; Tabas, I. Insulin resistance, hyperglycemia, and atherosclerosis. Cell Metab. 2011, 14, 575–585.
https://doi.org/10.1016/j.cmet.2011.07.015.
184. La Sala, L.; Prattichizzo, F.; Ceriello, A. The link between diabetes and atherosclerosis. Eur. J. Prev. Cardiol. 2019, 26, 15–24.
https://doi.org/10.1177/2047487319878373.
185. Patterson, E.; Ryan, P.M.; Cryan, J.F.; Dinan, T.G.; Ross, R.P.; Fitzgerald, G.F.; Stanton, C. Gut microbiota, obesity and diabetes.
Postgrad. Med. J. 2016, 92, 286–300. https://doi.org/10.1136/postgradmedj-2015-133285.
186. Abildgaard, A.; Elfving, B.; Hokland, M.; Wegener, G.; Lund, S. The microbial metabolite indole-3-propionic acid improves
glucose metabolism in rats, but does not affect behaviour. Arch. Physiol. Biochem. 2018, 124, 306–312.
https://doi.org/10.1080/13813455.2017.1398262.
187. Magder, S. The meaning of blood pressure. Crit. Care 2018, 22, 257. https://doi.org/10.1186/s13054-018-2171-1.
188. Ning, B.; Chen, Y.; Waqar, A.B.; Yan, H.; Shiomi, M.; Zhang, J.; Chen, Y.E.; Wang, Y.; Itabe, H.; Liang, J.; et al. Hypertension
enhances advanced atherosclerosis and induces cardiac death in watanabe heritable hyperlipidemic rabbits. Am. J. Pathol. 2018,
188, 2936–2947. https://doi.org/10.1016/j.ajpath.2018.08.007.
189. Pulakazhi Venu, V.K.; Saifeddine, M.; Mihara, K.; Tsai, Y.C.; Nieves, K.; Alston, L.; Mani, S.; McCoy, K.D.; Hollenberg, M.D.;
Hirota, S.A. The pregnane X receptor and its microbiota-derived ligand indole 3-propionic acid regulate endothelium-dependent
vasodilation. Am. J. Physiol. Endocrinol. Metab. 2019, 317, E350–E361. https://doi.org/10.1152/ajpendo.00572.2018.
190. Patterson, R.A.; Stankewicz, H.A. Penicillin Allergy. In Stat Pearls; Treasure Island (FL, US): 2021.
191. Zhu, D.; Sorg, J.A.; Sun, X. Clostridioides difficile biology: Sporulation, germination, and corresponding therapies for C. difficile
infection. Front. Cell Infect. Microbiol. 2018, 8, 29. https://doi.org/10.3389/fcimb.2018.00029.
192. Brown, S.A.; Riviere, J.E. Comparative pharmacokinetics of aminoglycoside antibiotics. J. Vet. Pharmacol. Ther. 1991, 14, 1–35.
https://doi.org/10.1111/j.1365-2885.1991.tb00801.x.
193. Behr, C.; Kamp, H.; Fabian, E.; Krennrich, G.; Mellert, W.; Peter, E.; Strauss, V.; Walk, T.; Rietjens, I.; van Ravenzwaay, B. Gut
microbiome-related metabolic changes in plasma of antibiotic-treated rats. Arch. Toxicol. 2017, 91, 3439–3454.
https://doi.org/10.1007/s00204-017-1949-2.
194. Wilkins, T.; Sequoia, J. Probiotics for gastrointestinal conditions: A summary of the evidence. Am. Fam. Physician 2017, 96, 170–
178.
Int. J. Mol. Sci. 2022, 23, 1222 21 of 21
195. Sebastian Domingo, J.J. Review of the role of probiotics in gastrointestinal diseases in adults. Gastroenterol. Hepatol. 2017, 40, 417–
429. https://doi.org/10.1016/j.gastrohep.2016.12.003.
196. Kim, H.K.; Rutten, N.B.; Besseling-van der Vaart, I.; Niers, L.E.; Choi, Y.H.; Rijkers, G.T.; van Hemert, S. Probiotic
supplementation influences faecal short chain fatty acids in infants at high risk for eczema. Benef. Microbes. 2015, 6, 783–790.
https://doi.org/10.3920/BM2015.0056.
197. Primec, M.; Klemenak, M.; Di Gioia, D.; Aloisio, I.; Bozzi Cionci, N.; Quagliariello, A.; Gorenjak, M.; Micetic-Turk, D.; Langerholc,
T. Clinical intervention using Bifidobacterium strains in celiac disease children reveals novel microbial modulators of TNF-alpha
and short-chain fatty acids. Clin. Nutr. 2019, 38, 1373–1381. https://doi.org/10.1016/j.clnu.2018.06.931.
198. Abildgaard, A.; Elfving, B.; Hokland, M.; Wegener, G.; Lund, S. Probiotic treatment reduces depressive-like behaviour in rats
independently of diet. Psychoneuroendocrinology 2017, 79, 40–48. https://doi.org/10.1016/j.psyneuen.2017.02.014.
199. Savino, F.; Cordisco, L.; Tarasco, V.; Palumeri, E.; Calabrese, R.; Oggero, R.; Roos, S.; Matteuzzi, D. Lactobacillus reuteri DSM
17938 in infantile colic: A randomized, double-blind, placebo-controlled trial. Pediatrics 2010, 126, e526–e533.
https://doi.org/10.1542/peds.2010-0433.
200. Szajewska, H.; Gyrczuk, E.; Horvath, A. Lactobacillus reuteri DSM 17938 for the management of infantile colic in breastfed
infants: A randomized, double-blind, placebo-controlled trial. J. Pediatr. 2013, 162, 257–262.
https://doi.org/10.1016/j.jpeds.2012.08.004.
... It has been suggested that the number of human cells in the human body may be fewer than its microbial cells and that the number of human genes may be 10-100 times less than what is found in the intestine of the host [1,[7][8][9]. In the last few decades, the amount of research conducted on the influence of the gut microbiome on the host organism, in a field called microbiomics, has increased significantly [10][11][12]. A new theory about the microbiome being a human "organ" gave impetus to investigating mechanisms of interactions between bacterial communities and other organism systems. ...
... Tryptophan and its metabolites (melatonin is one of them) are also discussed as gut microbiome modifiers influenced by oxidative processes. Several bacteria transform tryptophan into its indolic derivatives, and some bacteria protect cells from oxidative damage [10]. Butyrate is one of the SCFAs produced by microflora, and in the OS condition, this compound counteracted the negative effect of ROS on tryptophan uptake [90]. ...
Article
Full-text available
Understanding how gut flora interacts with oxidative stress has been the subject of significant research in recent years. There is much evidence demonstrating the existence of the microbiome–oxidative stress interaction. However, the biochemical basis of this interaction is still unclear. In this narrative review, possible pathways of the gut microbiota and oxidative stress interaction are presented, among which genetic underpinnings play an important role. Trimethylamine-N-oxide, mitochondria, short-chain fatty acids, and melatonin also appear to play roles. Moreover, the relationship between oxidative stress and the gut microbiome in obesity, metabolic syndrome, chronic ethanol consumption, dietary supplements, and medications is considered. An investigation of the correlation between bacterial community features and OS parameter changes under normal and pathological conditions might provide information for the determination of new research methods. Furthermore, such research could contribute to establishing a foundation for determining the linkers in the microbiome–OS association.
... It was further known from the literature that UDCA and GUDCA were bile acid-like compounds mainly produced by the metabolism of intestinal ora 15,16 . IPA and I3C were metabolism products of the tryptophan by intestinal ora 17,18 . Consequently, we preliminarily deduced that the four metabolites might play important roles in the improving effect of MP-A for NAFLD. ...
Preprint
Full-text available
We investigated the ameliorative effect of mussel polysaccharide α-D-glucan (MP-A) on non-alcoholic fatty liver disease (NAFLD) and further investigated its action mechanism based on serum metabolomics. The results showed that MP-A improved the hepatic steatosis in the rat NAFLD model induced by high-fat diet (HFD) significantly. MP-A could improve the metabolic disorders in NAFLD models. Four metabolites, ursodeoxycholic acid (UDCA), indole-3-propionic acid (IPA), indole-3-carbinol (I3C) and glycine ursodeoxycholic acid (GUDCA), mainly produced by intestinal flora and absorbed into the blood were screened out through serum metabolomics techniques and analysis. Compared with the normal group, the levels of these four metabolites were reduced in the HFD group but increased through the MP-A intervention. The four metabolites significantly reduced the intracellular lipid droplet accumulation and triglyceride (TG) content in sodium oleate-induced HepG2 cells. Meanwhile, they upregulated the intracellular farnesoid X receptor (FXR) protein expression and downregulated the sterol regulatory element-binding protein-1C (SREBP-1C) and fatty acid synthase (FAS) proteins expression. These results suggested that MP-A might improve NAFLD by affecting the metabolites of intestinal flora, UDCA, IPA, I3C and GUDCA, which then regulated the FXR-SREBP-1C-FAS signaling pathway and liver lipid metabolism.
... Conversely, Blautia, Klebsiella, and Shigella have a greater abundance in patients with kidney diseases, and a similar association of Blautia was also found with adults receiving regular dialysis [44]. Under the anti-oxidative effects of indoxy-3-propionic acid [45], the abundance of those bacteria in non-CKD groups may indicate lower oxidative stress to renal tissue. Roseburia inulinivorans and Roseburia intestinalis are butyrate producers, which promote gut health [46]. ...
Article
Full-text available
Microbiota tryptophan metabolism and the biosynthesis of indole derivatives play an important role in homeostasis and pathogenesis in the human body and can be affected by the gut microbiota. However, studies on the interplay between gut microbiota and tryptophan metabolites in patients undergoing dialysis are lacking. This study aimed to identify the gut microbiota, the indole pathway in tryptophan metabolism, and significant functional differences in ESRD patients with regular hemodialysis. We performed the shotgun metagenome sequencing of stool samples from 85 hemodialysis patients. Using the linear discriminant analysis effect size (LEfSe), we examined the composition of the gut microbiota and metabolic features across varying concentrations of tryptophan and indole metabolites. Higher tryptophan levels promoted tyrosine degradation I and pectin degradation I metabolic modules; lower tryptophan levels were associated with glutamate degradation I, fructose degradation, and valine degradation modules. Higher 3-indoxyl sulfate concentrations were characterized by alanine degradation I, anaerobic fatty acid beta-oxidation, sulfate reduction, and acetyl-CoA to crotonyl-CoA. Contrarily, lower 3-indoxyl sulfate levels were related to propionate production III, arabinoxylan degradation, the Entner–Doudoroff pathway, and glutamate degradation II. The present study provides a better understanding of the interaction between tryptophan, indole metabolites, and the gut microbiota as well as their gut metabolic modules in ESRD patients with regular hemodialysis.
... This axis plays a crucial role in maintaining overall balance within the body and regulates the immune system. By doing so, it exhibits anti-inflammatory and antioxidant properties, working in harmony to modulate various physiological processes [46,47]. We hypothesized that protection of microbial IPA production is one of the mechanisms through which RS2 plays a prebiotic role. ...
Article
Full-text available
Owing to the interplay of genetic and environmental factors, obesity has emerged as a significant global public health concern. To gain enhanced control over obesity, we examined the effects of type 2 resistant starch (RS2) and its promoted microbial-derived metabolite, indole-3-propionic acid (IPA), on hepatic steatosis, antioxidant activity, and gut microbiota in obese mice. Neither RS2 nor low-dose IPA (20 mg kg−1) exhibited a reduction in body weight or improved glucose and lipid metabolism in post-obesity state mice continuously fed the high-fat diet (HFD). However, both interventions improved hepatic steatosis, with RS2 being more effective in all measured parameters, potentially due to changes in gut microbiota and metabolites not solely attributed to IPA. LC-MS/MS analysis revealed increased serum IPA levels in both RS2 and IPA groups, which positively correlated with Bifidobacterium and Clostridium. Moreover, RS2 exhibited a more significant restoration of gut dysbiosis by promoting the abundance of health-promoting bacteria including Faecalibaculum and Bifidobacterium. These findings suggest that the regulatory role of RS2 on tryptophan metabolism only partially explains its prebiotic activity. Future studies should consider increasing the dose of IPA and combining RS2 and IPA to explore their potential interventions in obesity.
... Low tyrosine levels may indicate renal injury, which is a well-known metabolic alteration post radiation exposure [17]. Along with Trp, the metabolic disturbances in indole-3-lactic acid (I3LA), indole-3-propionic acid (I3PA), and tryptophol (indole-3-ethanol (IEA)) levels indicate disturbances in the gut microbiota-promoted shikimate pathway [18][19][20]. The upregulation of lathosterol, which is a precursor for cholesterols, also overlapped in both genders post-IR. ...
Article
Full-text available
Exposure to ionizing radiation, accidental or intentional, may lead to delayed effects of acute radiation exposure (DEARE) that manifest as injury to organ systems, including the kidney, heart, and brain. This study examines the role of activated protein C (APC), a known mitigator of radiation-induced early toxicity, in long-term plasma metabolite and lipid panels that may be associated with DEARE in APCHi mice. The APCHi mouse model used in the study was developed in a C57BL/6N background, expressing the D168F/N173K mouse analog of the hyper-activatable human D167F/D172K protein C variant. This modification enables increased circulating APC levels throughout the mouse’s lifetime. Male and female cohorts of C57BL/6N wild-type and APCHi transgenic mice were exposed to 9.5 Gy γ-rays with their hind legs shielded to allow long-term survival that is necessary to monitor DEARE, and plasma was collected at 6 months for LC-MS-based metabolomics and lipidomics. We observed significant dyslipidemia, indicative of inflammatory phenotype, upon radiation exposure. Additionally, observance of several other metabolic dysregulations was suggestive of gut damage, perturbations in TriCarboxylic Acid (TCA) and urea cycles, and arginine metabolism. We also observed gender- and genotype-modulated metabolic perturbations post radiation exposure. The APCHi mice showed near-normal abundance for several lipids. Moreover, restoration of plasma levels of some metabolites, including amino acids, citric acid, and hypoxanthine, in APCHi mice is indicative of APC-mediated protection from radiation injuries. With the help of these findings, the role of APC in plasma molecular events after acute γ-radiation exposure in a gender-specific manner can be established for the first time.
Article
Full-text available
Autism spectrum disorder (ASD) is a heterogeneous neurodevelopmental condition characterized by social communication deficiencies and stereotypic behaviors influenced by hereditary and/or environmental risk factors. There are currently no approved medications for treating the core symptoms of ASD. Human fecal microbiota transplantation (FMT) has emerged as a potential intervention to improve autistic symptoms, but the underlying mechanisms are not fully understood. In this study, we evaluated the effects of human-derived FMT on behavioral and multi-omics profiles of the BTBR mice, an established model for ASD. FMT effectively alleviated the social deficits in the BTBR mice and normalized their distinct plasma metabolic profile, notably reducing the elevated long-chain acylcarnitines. Integrative analysis linked these phenotypic changes to specific Bacteroides species and vitamin B6 metabolism. Indeed, vitamin B6 supplementation improved the social behaviors in BTBR mice. Collectively, these findings shed new light on the interplay between FMT and vitamin B6 metabolism and revealed a potential mechanism underlying the therapeutic role of FMT in ASD. IMPORTANCE Accumulating evidence supports the beneficial effects of human fecal microbiota transplantation (FMT) on symptoms associated with autism spectrum disorder (ASD). However, the precise mechanism by which FMT induces a shift in the microbiota and leads to symptom improvement remains incompletely understood. This study integrated data from colon-content metagenomics, colon-content metabolomics, and plasma metabolomics to investigate the effects of FMT treatment on the BTBR mouse model for ASD. The analysis linked the amelioration of social deficits following FMT treatment to the restoration of mitochondrial function and the modulation of vitamin B6 metabolism. Bacterial species and compounds with beneficial roles in vitamin B6 metabolism and mitochondrial function may further contribute to improving FMT products and designing novel therapies for ASD treatment.
Article
In modern scientific literature, close attention is paid to the biological role of tryptophan catabolites both in normal conditions and in various pathologies. There are more and more reports that tryptophan metabolism catabolites play a signaling role in the human body and in the intestinal microbial community. Receptors and signaling pathways in the human body, the so-called tryptophan signaling molecules (TrySM), their cellular targets, physiological and metabolic effects are being actively studied. It has now been established that almost all catabolites of tryptophan metabolism are signaling molecules. Many of them realize their signaling role through aryl hydrocarbon receptors (АhR). The dominant pathway of tryptophan metabolism for the human body is the kynurenine pathway, which is the source of universal signaling molecules – kynurenine, quinolinic and kynurenic acids. The indole pathway of tryptophan catabolism, the main one for the microbiota, with the exception of indole formation reactions in immunocompetent cells, is a source of interregional and interspecies signaling molecules – indole and its derivatives: indole-3-pyruvate, indole-3-lactate, indole-3-acetate, indole-3-propionate, indole-3-acrylate, indole-3-butyrate and indole-3-acetaldehyde. Serotonin and melatonin are also universal signaling molecules and have been widely studied in various diseases of the nervous system.
Article
Full-text available
It is well-known that the pregnane X receptor (PXR/Nr1i2) is a critical xenobiotic-sensing nuclear receptor enriched in liver and intestine and is responsible for drug-drug interactions (DDI), due to their versatile ligand binding domain (LBD) and target genes involved in xenobiotic biotransformation. PXR can be modulated by various xenobiotics including pharmaceuticals, nutraceuticals, dietary factors, and environmental chemicals. Microbial metabolites such as certain secondary bile acids (BAs) and the tryptophan metabolite indole-3-propionic acid (IPA) are endogenous PXR activators. Gut microbiome is increasingly recognized as an important regulator for host xenobiotic biotransformation and intermediary metabolism. PXR regulates and is regulated by the gut-liver axis. This review summarizes recent research advancements leveraging pharmaco- and toxico-metagenomic approaches that have redefined the previous understanding of PXR. Key topics covered in this review include 1) genome-wide investigations on novel PXR-target genes, novel PXR-DNA interaction patterns, and novel PXR-targeted intestinal bacteria; 2) key PXR-modulating activators and suppressors of exogenous and endogenous sources; 3) novel bi-directional interactions between PXR and gut microbiome under physiological, pathophysiological, pharmacological, and toxicological conditions; and 4) modifying factors of PXR-signaling including species- and sex difference, and time (age, critical windows of exposure, and circadian rhythm). The review also discusses critical knowledge gaps and important future research topics centering around PXR. Significance Statement This review summarizes recent research advancements leveraging O'mics approaches that have redefined the previous understanding of the xenobiotic-sensing nuclear receptor PXR. Key topics include: 1) genome-wide investigations on novel PXR-targeted host genes and intestinal bacteria as well as novel PXR-DNA interaction patterns; 2) key PXR modulators including microbial metabolites under physiological, pathophysiological, pharmacological, and toxicological conditions; and 3) modifying factors including species, sex, and time.
Article
Full-text available
Clostridium sporogenes (C. sporogenes), as a potential probiotic, metabolizes tryptophan and produces an anti-inflammatory metabolite, indole-3-propionic acid (IPA). Herein, we studied the effects of C. sporogenes and its bioactive metabolite, IPA, on skeletal muscle development and chronic inflammation in mice. In the in vivo study, the muscle tissues and serum samples of mice with C. sporogenes supplementation were used to analyze the effects of C. sporogenes on muscle metabolism; the IPA content was determined by metabonomics and ELISA. In an in vitro study, C2C12 cells were exposed to lipopolysaccharide (LPS) alone or LPS + IPA to verify the effect of IPA on muscle cell inflammation by transcriptome, and the involved mechanism was revealed by different functional assays. We observed that C. sporogenes colonization significantly increased the body weight and muscle weight gain, as well as the myogenic regulatory factors’ (MRFs) expression. In addition, C. sporogenes significantly improved host IPA content and decreased pro-inflammatory cytokine levels in the muscle tissue of mice. Subsequently, we confirmed that IPA promoted C2C12 cells’ proliferation by activating MRF signaling. IPA also effectively protected against LPS-induced C2C12 cells inflammation by activating Pregnane X Receptor and restoring the inhibited miR-26a-2-3p expression. miR-26a-2-3p serves as a novel muscle inflammation regulatory factor that could directly bind to the 3′-UTR of IL-1β, a key initiator factor in inflammation. The results suggested that C. sporogenes with its functional metabolite IPA not only helps muscle growth development, but also protects against inflammation, partly by the IPA/ miR-26a-2-3p /IL-1β cascade.
Article
Full-text available
Tryptophan is an essential amino acid whose metabolites play key roles in diverse physiological processes. Due to low reserves in the body, especially under various catabolic conditions, tryptophan deficiency manifests itself rapidly, and both the serotonin and kynurenine pathways of metabolism are clinically significant in critically ill patients. In this review, we highlight these pathways as sources of serotonin and melatonin, which then regulate neurotransmission, influence circadian rhythm, cognitive functions, and the development of delirium. Kynurenines serve important signaling functions in inter-organ communication and modulate endogenous inflammation. Increased plasma kynurenine levels and kynurenine-tryptophan ratios are early indicators for the development of sepsis. They also influence the regulation of skeletal muscle mass and thereby the development of polyneuromyopathy in critically ill patients. The modulation of tryptophan metabolism could help prevent and treat age-related disease with low grade chronic inflammation as well as post intensive care syndrome in all its varied manifestations: cognitive decline (including delirium or dementia), physical impairment (catabolism, protein breakdown, loss of muscle mass and tone), and mental impairment (depression, anxiety or post-traumatic stress disorder).
Article
Full-text available
Background and aims: Gut microbiota-derived metabolites play a vital role in maintenance of human health and progression of disorders, including obesity and type 2 diabetes (T2D). Indole-3-propionic acid (IPA), a gut-derived tryptophan metabolite, has been recently shown to be lower in individuals with obesity and T2D. IPA's beneficial effect on liver health has been also explored in rodent and cell models. In this study, we investigated the association of IPA with human liver histology and transcriptomics, and the potential of IPA to reduce hepatic stellate cell activation in vitro. Methods: A total of 233 subjects (72% women; age 48.3 ± 9.3 years; BMI 43.1 ± 5.4 kg/m2) undergoing bariatric surgery with detailed liver histology were included. Circulating IPA levels were measured using LC-MS and liver transcriptomics with total RNA-sequencing. LX-2 cells were used to study hepatoprotective effect of IPA in cells activated by TGF-β1. Results: Circulating IPA levels were found to be lower in individuals with liver fibrosis compared to those without fibrosis (p = 0.039 for all participants; p = 0.013 for 153 individuals without T2D). Accordingly, levels of circulating IPA associated with expression of 278 liver transcripts (p < 0.01) that were enriched for the genes regulating hepatic stellate cells (HSCs) activation and hepatic fibrosis signaling. Our results suggest that IPA may have hepatoprotective potential because it is able to reduce cell adhesion, cell migration and mRNA gene expression of classical markers of HSCs activation in LX-2 cells (all p < 0.05). Conclusion: The association of circulating IPA with liver fibrosis and the ability of IPA to reduce activation of LX-2 cells suggests that IPA may have a therapeutic potential. Further molecular studies are needed to investigate the mechanisms how IPA can ameliorate hepatic fibrosis.
Article
Full-text available
Assessing dementia conversion in patients with mild cognitive impairment (MCI) remains challenging owing to pathological heterogeneity. While many MCI patients ultimately proceed to Alzheimer’s disease (AD), a subset of patients remain stable for various times. Our aim was to characterize the plasma metabolites of nineteen MCI patients proceeding to AD (P-MCI) and twenty-nine stable MCI (S-MCI) patients by untargeted metabolomics profiling. Alterations in the plasma metabolites between the P-MCI and S-MCI groups, as well as between the P-MCI and AD groups, were compared over the observation period. With the help of machine learning-based stratification, a 20-metabolite signature panel was identified that was associated with the presence and progression of AD. Furthermore, when the metabolic signature panel was used for classification of the three patient groups, this gave an accuracy of 73.5% using the panel. Moreover, when specifically classifying the P-MCI and S-MCI subjects, a fivefold cross-validation accuracy of 80.3% was obtained using the random forest model. Importantly, indole-3-propionic acid, a bacteria-generated metabolite from tryptophan, was identified as a predictor of AD progression, suggesting a role for gut microbiota in AD pathophysiology. Our study establishes a metabolite panel to assist in the stratification of MCI patients and to predict conversion to AD.
Preprint
Full-text available
Cardiovascular disease (CVD) is one of the major causes of mortality worldwide. Inflammation is the underlying common mechanism involved in CVD. It has been recently related to amino acid metabolism, which acts as a critical regulator of innate and adaptive immune responses. Among different metabolites that have emerged as important regulators of immune and inflammatory responses, tryptophan (Trp) metabolites have been shown to play a pivotal role in CVD. Here, we provide an overview of the fundamental aspects of Trp metabolism and the interplay between the dysregulation of the main actors involved in Trp metabolism such as indoleamine 2, 3-dioxygenase 1 (IDO) and CVD, including atherosclerosis and myocardial infarction. IDO has a prominent and complex role. Its activity, impacting on several biological pathways, complicates our understanding of its function, particularly in CVD, where it is still under debate. The discrepancy of the observed IDO effects could be potentially explained by its specific cell and tissue contribution, encouraging further investigations regarding the role of this enzyme. Thus, improving our understanding of the function of Trp as well as its derived metabolites will help to move one step closer towards tailored therapies aiming to treat CVD.
Article
Full-text available
Iron excess in tissues results in increased oxidative damage. Among different tissues, the skin can particularly be severely damaged by oxidative stress, as it is exposed not only to endogenous but also directly to exogenous pro-oxidants. The skin is especially vulnerable to harmful oxidative stress. Melatonin and indole-3-propionic acid (IPA), two indole substances, are efficient antioxidants. This study aims to evaluate the potential protective effects of melatonin and IPA against oxidative damage to membrane lipids (lipid peroxidation (LPO)), induced in porcine skin homogenates by the Fenton reaction (Fe2+ + H2O2 → Fe3+ + •OH + OH−) when iron is used in extremely high concentrations. Skin homogenates were incubated in the presence of FeSO4 (2400, 1200, 600, 300, 150 and 75 µM) + H2O2 (5 mM) with/without melatonin or IPA. LPO level (MDA + 4-HDA/mg protein) was measured spectrophotometrically. Melatonin, in its highest used concentration (5.0 mM), prevented FeSO4 (1200 mM)-induced LPO, whereas it was effective in concentrations as low as 2.5 mM against all lower iron concentrations. IPA was protective in concentrations as low as 2.5 mM independently of FeSO4 concentration. In conclusion, melatonin and IPA effectively protect against oxidative damage to membrane lipids induced by high concentrations of iron in porcine skin; therefore, both can be considered pharmacological agents in the case of disorders associated with excessive iron accumulation in the skin.
Article
This study probed the neuroprotective influence of indole-3-propionic acid (IPA) in rats exposed to chlorpyrifos (CPF) alone at 5 mg/kg body weight or co-administered with IPA at 12.5 and 25 mg/kg for 14 days. Behavioral data indicated that IPA significantly (p < 0.05) abated CPF-mediated anxiogenic‐like behaviors with concomitant improvement in the locomotor and exploratory behaviors as substantiated by track plots and heat maps data. Also, IPA mitigated CPF-mediated diminution in cholinergic and antioxidant defense systems whereas it markedly improved thioredoxin level and thioredoxin reductase activity in cerebral and cerebellar tissues of the animals. Co-administration of IPA significantly enhanced anti-inflammatory cytokine, interleukin-10 but suppressed oxidative and inflammatory stress, caspase‐9 and caspase-3 activation with concomitant reduction in 8-hydroxy-2'-deoxyguanosine (8-OHdG) level and histological damage. Collectively, IPA‐mediated neuroprotection involves modulation of cholinergic and redox-regulatory systems, inflammatory stress, apoptotic responses and DNA damage in cerebrum and cerebellum of rats.
Article
Recent evidence suggests that gut bacteria-derived metabolites interact with the cardiovascular system and alter blood pressure (BP) in mammals. Here, we evaluated the effect of indole-3-propionic acid (IPA), a gut bacteria-derived metabolite of tryptophan, on the circulatory system. Arterial BP, electrocardiographic, and echocardiographic (ECHO) parameters were recorded in male, anesthetized, 12-wk-old Wistar-Kyoto rats at baseline and after intravenous administration of either IPA or vehicle. In additional experiments, rats were pretreated with prazosin or pentolinium to evaluate the involvement of the autonomic nervous system in cardiovascular responses to IPA. IPA’s concentrations were measured using ultra-high performance liquid chromatography tandem mass spectrometry. The reactivity of endothelium-intact and -denuded mesenteric resistance arteries was tested. Cells’ viability and lactate dehydrogenase (LDH) cytotoxicity assays were performed on cultured cardiomyocytes. IPA increased BP with a concomitant bradycardic response but no significant change in QTc interval. The pretreatment with prazosin and pentolinium reduced the hypertensive response. ECHO showed increased contractility of the heart after the administration of IPA. Ex vivo, IPA constricted predilated and endothelium-denuded mesenteric resistance arteries and increased metabolic activity of cardiomyocytes. IPA increases BP via cardiac and vascular mechanisms in rats. Furthermore, IPA increases cardiac contractility and metabolic activity of cardiomyocytes. Our study suggests that IPA may act as a mediator between gut microbiota and the circulatory system.