Next Article in Journal
Discovery of Novel Biomarkers with Extended Non-Coding RNA Interactor Networks from Genetic and Protein Biomarkers
Previous Article in Journal
bZIP Transcription Factor PavbZIP6 Regulates Anthocyanin Accumulation by Increasing Abscisic Acid in Sweet Cherry
Previous Article in Special Issue
Unveiling the Role of Endothelial Dysfunction: A Possible Key to Enhancing Catheter Ablation Success in Atrial Fibrillation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Gut Microbial Metabolites in Cardiovascular Diseases—Current Insights and the Road Ahead

1
Department of Pharmacological and Pharmaceutical Sciences, College of Pharmacy, University of Houston, Houston, TX 77204, USA
2
Department of Pharmaceutical Sciences, Irma Lerma College of Pharmacy, Texas A&M University, Kingsville, TX 78363, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(18), 10208; https://doi.org/10.3390/ijms251810208
Submission received: 19 August 2024 / Revised: 18 September 2024 / Accepted: 19 September 2024 / Published: 23 September 2024
(This article belongs to the Special Issue Endothelial Dysfunction and Cardiovascular Diseases)

Abstract

:
Cardiovascular diseases (CVDs) are the leading cause of premature morbidity and mortality globally. The identification of novel risk factors contributing to CVD onset and progression has enabled an improved understanding of CVD pathophysiology. In addition to the conventional risk factors like high blood pressure, diabetes, obesity and smoking, the role of gut microbiome and intestinal microbe-derived metabolites in maintaining cardiovascular health has gained recent attention in the field of CVD pathophysiology. The human gastrointestinal tract caters to a highly diverse spectrum of microbes recognized as the gut microbiota, which are central to several physiologically significant cascades such as metabolism, nutrient absorption, and energy balance. The manipulation of the gut microbial subtleties potentially contributes to CVD, inflammation, neurodegeneration, obesity, and diabetic onset. The existing paradigm of studies suggests that the disruption of the gut microbial dynamics contributes towards CVD incidence. However, the exact mechanistic understanding of such a correlation from a signaling perspective remains elusive. This review has focused upon an in-depth characterization of gut microbial metabolites and their role in varied pathophysiological conditions, and highlights the potential molecular and signaling mechanisms governing the gut microbial metabolites in CVDs. In addition, it summarizes the existing courses of therapy in modulating the gut microbiome and its metabolites, limitations and scientific gaps in our current understanding, as well as future directions of studies involving the modulation of the gut microbiome and its metabolites, which can be undertaken to develop CVD-associated treatment options. Clarity in the understanding of the molecular interaction(s) and associations governing the gut microbiome and CVD shall potentially enable the development of novel druggable targets to ameliorate CVD in the years to come.

1. Introduction

Cardiovascular diseases (CVDs) constitute the most frequent known non-communicable diseases, accounting for an approximately 17.9 million mortalities globally [1,2]. Several risk factors play a pivotal role in CVD onset and progression, including hypertension, tobacco smoking, unrestrained alcohol consumption, dyslipidemia, diabetic incidence and a sedentary lifestyle [3,4]. CVDs encompass a broad spectrum of cardiovascular dysfunctions that chiefly include coronary heart disease [5,6], cerebrovascular anomalies [7,8], peripheral arterial dysfunction [9,10], rheumatic cardiac malfunction [11,12], congenital heart defects [13,14], myocardial infarction (MI) [15,16,17] and pulmonary embolism [18,19]. The commonality among these dysfunctions essentially results in defects in cardiac morphogenesis, repair, and cardiac functioning rhythm [17,20]. The heart encounters a marked decrease in proliferative cardiomyocyte count upon CVD onset. Adult cardiomyocytes exhibit a compromised ability to proliferate, which limits the cardiac regenerative capacity, accounting for ventricular remodeling, a loss in muscle strength, contractility and subsequent congestive heart failure [21,22]. Although significant progress has been made in CVD therapy development [23,24], it is important to recognize the emerging risk factors and novel molecular signaling cascades governing CVD pathophysiology.
The identification of the role of risk factors in contributing towards CVD onset and progression has enabled a better understanding of the plethora of biochemical and histological cascades driving CVD onset and progression, characterized by vascular and endothelial dysfunction [25,26], inflammatory episodes [27,28], prothrombotic status [29,30], impact on basal metabolism [31,32] and lipid profile imbalance [33,34]. Observational cohort studies and clinical trials emphasize the persistence of residual inflammatory risk in CVD patients post statin therapy [35,36,37]. Although precision medicine-based studies hold great promise for improved, personalized, integrated and patient-specific approaches for CVD therapy, patient-centric needs demand the identification of new risk factors and druggable targets for better outcomes of CVD therapy in cardiac patients [38]. In this context, gut microbiome and their metabolites have recently been recognized as novel risk factors affecting a variety of physiological processes and resulting in several disease pathologies, including CVDs.
The human gastrointestinal tract accommodates a diverse and dynamic spectrum of the microbial population, collectively referred to as gut microbiota [39,40]. It extends across the intestinal epithelium and is chiefly composed of physiologically beneficial populations of Firmicutes, Bacteroidetes, Actinobacteria, Proteobacteria, and Verrucomicrobia [41,42,43]. The advent of state-of-the-art techniques such as metagenomic analysis, marker gene analysis, and meta-transcriptome have enabled an improved characterization of the gut microbiota [44,45,46]. From a functional perspective, the gut microbiota primarily develop and maintain the intestinal barrier, the activation of the immune system and the metabolic breakdown of nutrients and drugs [47,48,49]. At the endocrine level, the gut microbiota modulates the metabolism of leptin, ghrelin, and cortisol [50]. Dysbiosis of the gut microbial dynamics, caused due to anomalies in diet, enteric pathogen population, toxins, and other environmental factors, results in diverse ailments like inflammatory bowel disease, diabetes, allergic onset, and neurological complications [51,52,53,54]. The gut microbial breakdown of dietary food components generates a broad spectrum of metabolites, such as short-chain fatty acids (SCFA), branched-chain fatty acids, amines, phenolic compounds, hydrogen sulfide, and methane [55,56,57,58,59,60]. Studies have proven that these gut microbial metabolites play a pivotal role in affecting cardiovascular functioning [61,62,63]. The gut microbiome and the heart interact with each other through these metabolites that are reabsorbed and subsequently transferred to systemic circulation [64,65]. Certain gut-microbe-derived-metabolites are beneficial to cardiac functioning, while some have been reported to be detrimental [65,66]. Although a cause–effect relationship between gut microbial metabolites and cardiovascular anomalies has been reported, the exact signaling mechanism governing such interactions remains unclear.
This review focuses on the signaling mechanisms identified so far, that drive the interaction between gut microbial metabolites and cardiovascular function. The metabolites that play a pivotal role in the CVD onset and progression and the status of their druggability is comprehensively discussed. An in-depth understanding of the signaling cross-talks involved in the gut–heart axis shall potentially benefit towards the development of novel and effective druggable targets to ameliorate CVD over the long term.

2. Pathophysiology of Gut Microbial Metabolites in CVDs

The gut microbial/host metabolic interactions play a pivotal role in the generation of a broad spectrum of metabolites and associated small molecules [67,68,69]. The extent of the generation of these metabolites is usually determined by interindividual genomic differences and the availability of dietary substrates, among multiple other environmental factors [70,71,72]. These metabolites undergo absorption across the host gut, enter systemic body circulation and influence vital physiological processes [68,73].

2.1. Short-Chain Fatty Acids (SCFAs)

Short-chain fatty acids (SCFAs) constitute a majority of the gut microbial metabolites produced because of the gut microbial fermentation of dietary macromolecules. The gut microbiota participate in the metabolic breakdown of sugars via the saccharolytic pathway and protein fermentation to generate SCFAs [74]. Protein fermentation additionally contributes towards the generation of co-metabolites like amines, ammonia, thiols, indoles, and other polyphenolic compounds [75]. The SCFAs exercise tissue-specific physiological influence, primarily driven by their concentration differences in the gut lumen and circulating blood, as well as the hepatic and peripheral sites [76,77]. Functionally, the SCFAs contribute towards the prevention of the onset and progression of metabolic dysfunctions, inflammatory complications, and cancer [78,79]. SCFAs inhibit the formation of foam cells and the influx of monocytes, as well as reducing the production of pro-inflammatory cytokines by the endothelium, and thus contribute to the recovery of endothelial dysfunction. SCFAs protect against obesity, improve glucose tolerance, and attenuate ulcerative colitis, Crohn’s disease, and antibiotic-associated diarrhea [80,81]. SCFAs augment postprandial plasma peptide YY (PYY) and glucagon-like peptide-1 (GLP-1) release from colon cells and downregulate intra-abdominal adipose tissue distribution and intrahepatocellular lipid content [82].
From a molecular perspective, SCFAs mediate their protective functions predominantly via the activation of the SCFA receptor-G-protein-coupled receptor 43 (GPR43), also identified as free fatty acid receptor 2 (FFAR2) [83]. GPR43 is chiefly expressed in the intestinal endocrine L-cells and its activation by SCFAs augments GLP-1 secretion [84]. SCFA-induced GPR43 activation also accounts for the decrease in adipose-specific insulin signaling [85]. This, in turn, hinders fat accumulation and attenuates obesity [86]. Studies have also shown that SCFA-induced GPR43 activation strengthens the colonic T cell population and exhibits a protective function against colitis, and type 2 diabetes [87]. SCFAs also activate GPR41, also known as FFAR3, in the peripheral nerves, and improve insulin resistance [88]. GPR41 activation by SCFAs increases macrophage and dendritic cell precursors and protects against respiratory problems [89]. The SCFA-associated activation of the colonic epithelium localized GPR109A receptor promotes anti-inflammatory factors in colonic macrophages, and augments the differentiation of regulatory and interleukin (IL)-10 producing T cells. This, in turn, prevents colonic inflammation and subsequent cancer incidence [90,91,92,93]. Thus, SCFAs activate multiple GPRs and mediate a variety of physiological and pathological processes [94].

2.2. Bile Acids

Bile acids (BAs) are amphipathic molecules generated in the liver from cholesterol and stored in the gall bladder [95]. The overall composition of the gut microbiota and associated influence on BA biogenesis are key drivers of interspecies and interindividual BA pool heterogeneity [96,97]. The mammalian BA pool comprises primary BA and secondary BA. Primary BAs are generated by hepatocytes and stored in the gall bladder, while the secondary BAs are primarily derived from gut microbial metabolism [98]. Primary BAs are either produced by cholesterol-7α-hydroxylase (CYP7A1) via the classical pathway in hepatocytes or by 27-hydroxylase (CYP27A1) via the acidic pathway in extrahepatic sites [98]. The primary BAs undergo conjugation with glycine or taurine and form bile salts and secondary BAs to be stored in gall bladder [99]. BAs determine the overall metabolic capacity of the gut microbiota. Primary BAs upregulate the expression of BA-metabolizing genes in the small intestine [100,101]. Excessive amphipathic BAs induce DNA damage, protein misfolding and oxidative stress, which can lower gut bacterial viability [102,103]. However, BAs also exhibit a protective role towards gut microbial diversity [100]. This is proven by studies demonstrating that cholestasis downregulates gut microbial diversity and functioning [104]. BAs potentially serve as energy sources to accommodate the needs to sustain gut microbial diversity [104].
From a pathophysiological perspective, BAs directly impede the differentiation and activation of the Clostridium difficile infection [105]. The upregulation of primary BA levels and downregulation of secondary BAs have been reported in inflammatory bowel disease (IBD), primarily via the suppression of nuclear factor Kappa-light-chain-enhancer of activated B cells (NF-κB) via the Farnesoid X receptor (FXR) and Pregnane X receptor (PXR) activation [106,107,108]. BAs also exhibit a beneficial impact against hypertriglyceridemia by the activation of the Farnesoid X receptor and Pregnane X receptor, which attenuates pro-inflammatory NF-κB signaling and downregulates triglyceride levels [94] (Figure 1). BA-associated FXR/SHP activation intervenes in the fatty acid biosynthesis driven by the liver X receptor (LXR) and sterol regulatory element-binding protein 1c (SREBP-1c) [109,110]. Secondary BAs have also been reported to ameliorate metabolic syndrome, although the mechanism governing the process remains unclear [111]. The gut-microbiota-derived BAs have been reported to augment hepatocellular carcinoma incidence by increasing the levels of deoxycholic acid (DCA) by 7α-hydroxylases [112,113]. Secondary BAs have been reported to exhibit greater carcinogenic potential as compared to primary Bas, mainly because of increased hydrophobicity, the disruption of the cellular membrane and the subsequent induction of cell damage responses [113,114]. Increased secondary BAs also augment reactive oxygen species (ROS) levels via phospholipase A2 activation [115,116].

2.3. Trimethylamine N-Oxide (TMAO)

The gut microbiota generate trimethylamine N-oxide (TMAO) from dietary products, employing microbial trimethylamine (TMA) lyases [117,118]. The gut-microbiota-driven metabolic conversion of choline, phosphatidylcholine and L-carnitine to TMA is catalyzed by TMA lyase, which breaks the carbon–nitrogen bonds in nutrient macromolecules and generates TMA [119,120]. TMA enters systemic circulation through the vessel wall, reaches the liver via hepatic portal circulation and is subsequently oxidized to TMAO by hepatic flavin monooxygenase-3 (FMO3) [121,122]. The generated TMAO re-enters systemic circulation and eventually accumulates in body tissues [123]. Plasma TMAO levels in healthy humans usually remain less than or equal to 5 µM and can shoot up to about 40 µM in patients with renal damage [124]. Other factors influencing TMAO levels in the body include age, cholic acid intake in food, and sex hormone levels [125,126]. Although kidneys are the primary organs involved in the TMAO elimination from the body, TMAO can also be eliminated via sweat, respiration, or fecal matter [127,128].
Recent studies have shown that TMAO exhibits a large pathophysiological influence on the onset and progression of heart failure, chronic kidney diseases, atherosclerosis, obesity, and diabetes [129,130,131,132]. Increased TMAO facilitates atherosclerotic plaque formation [122]. TMAO also augments thrombotic risk, promoting arterial thrombosis formation and shortening the duration of vascular occlusion [133]. It has been reported that TMAO promotes thrombin, collagen, platelet hyperactivity, endogenous calcium release from platelets, and amplified platelet adhesion to collagen [133]. TMAO also promotes monocyte adhesion to endothelial cells during foam cell formation and increases the count of scavenger receptors on macrophages [134,135]. Increased TMAO levels alter the static electricity at the arterial wall and significantly affect the inflow/efflux of fatty deposits on the arterial wall [136]. Consequently, foam cell formation rises and culminates to atherosclerotic plaque formation [136]. TMAO also has significant impact on cholesterol metabolism by reducing reverse cholesterol transport (RCT) [137]. This increases blood cholesterol levels and aggravates atherosclerotic progression [138]. Thus, TMAO is involved in multiple pathophysiological processes, including atherosclerosis, heart failure and platelet hyperreactivity [94] (Figure 2).

2.4. Indole-3-Propionic Acid (IPA)

Indole-3-propionic acid (IPA) is a metabolite of the gut-microbiota-driven breakdown of dietary tryptophan [139]. Serum IPA levels usually vary between 1 and 10 µM in humans under homeostatic conditions [140,141]. The gut microbial population predominantly involved in IPA formation include Lactobacillus reuteri, Clostridium sporogenes, Clostridium caloritolerans and members of Peptostreptococci genus [142,143]. The gut-microbiota-derived IPA production is chiefly governed by tryptophan aminotransferase [144,145,146]. Metabolic studies reveal that IPA augments gut–blood barrier functions by promoting the expression of claudins and tight junction proteins [147,148]. IPA is also involved in the activation of aryl hydrocarbon receptor in colonic epithelium and culminates in anti-inflammatory and anti-cancer effects [149,150]. IPA also exhibits a protective function against oxidative stress and lipid peroxidation damage [151,152,153]. IPA attenuates hematopoietic and gastrointestinal side effects in breast cancer cells [154,155]. From a neurological perspective, IPA downregulates the accumulation of amyloid β-proteins and restores mitochondrial functionality in Alzheimer’s disease [156,157]. IPA also prevents oxidative-stress-induced brain damage and ROS-induced cell death in Parkinson’s disease [158,159].

2.5. Hydrogen Sulfide

Gut-microbiota-driven hydrogen sulfide (H2S) is primarily generated from sulfate-reducing bacteria in the gut, which mainly include Desulfovibrio piger, Desulfovibrio desulficans and bacteria belonging to the Desulfobulbus and Desulfotomaculu genera [160,161]. These bacteria usually need a sulfate substrate and an electron donor for the generation of H2S [162]. Apart from these sulfate-reducing bacteria, anaerobic bacterial groups like Escherichia coli, Salmonella enterica and Enterobacter aerogenes mediate the catalytic breakdown of sulfur-containing cysteine into H2S, pyruvate and ammonia via cysteine desulfhydrase [163,164]. Sulfite reductase, a component of Escherichia coli, Salmonella and Klebsiella, also mediates H2S production via sulfite reduction reactions [165,166,167,168]. H2S acts as a signaling molecule that induces free radical responses and strengthens gut bacterial resistance to antibiotics [169]. H2S improves colon barrier integrity and downregulates local and systemic inflammatory cascades via the modulation of macrophages and cluster of differentiation (CD)8+ T cells [170,171]. H2S also prevents Nucleotide-binding oligomerization domain Leucine-rich Repeat and Pyrin domain-containing 3 (NLRP3) inflammasome-driven neuroinflammation via purinoreceptor-7 (P2X7) 14 inhibition in immune cells [172,173,174].

2.6. Phenylacetylglutamine (PAGln)

Phenylalanine, an essential amino acid, constitutes a chief precursor towards the synthesis of phenylacetylglutamine (PAGln)—primarily mediated via the meta organismal signaling pathway [175,176]. Unabsorbed phenylalanine culminates in the generation of phenylacetic acid (PAA) [177]. PAA, upon intestinal absorption, interacts with available glutamine residues and forms PAGln—chiefly catalyzed by hepatic enzymes [176].
Non-targeted metabolomic studies have established a strong correlation between PAGln and CVD onset, especially in type-2 diabetes patients [178]. PAGln has been reported to influence platelet functioning [179]. Studies have also reported that PAGln augments platelet adhesion to the collagen matrix and intracytoplasmic Ca2+ concentration [180,181]. The structural similarity with catecholamines has also led to hypotheses that PAGln can potentially modulate adrenergic receptors, accounting for GPCR (α2A, α2B, β2)-induced platelet reactivity and thrombosis [175]. This has been further substantiated via animal model studies, whereby PAGln was reported to trigger platelet activation and thrombosis—primarily driven via adrenergic signaling [178]. Patient-centric studies have reported that PAGln levels in plasma were significantly raised in coronary artery disease (CAD) patients with stent stenosis and hyperplasia [182]. This consolidates the idea that a rise in bacterial PAGln synthase-induced circulating PAGln accounts for in-stent stenosis in CAD patients [176].

3. Molecular Mechanisms of Gut Microbial Metabolites in Cardiovascular Diseases

Dysbiosis of the gut microbiota largely contributes towards the onset and progression of CVDs [183]. Heart failure (HF) patients exhibit a rise in Bifidobacterium and Enterobacteriaceae populations and a decreased count of Faecalibacterium prausnitzii [184,185,186]. Compromised gut microbial dynamics promote intestinal permeability and culminates in a leaky gut, prone to inflammation [187,188]. Gut microbial dysbiosis is attributed to diastolic dysfunction and augments left ventricular hypertrophy [189,190]. The gut microbiota interferes with the cardiac functioning through metabolites that undergo intestinal absorption and enter systemic circulation [191]. These metabolites have intrinsic signaling cross-talks in driving the gut microbiota/heart interaction axis, and contribute to CVDs.

3.1. Short-Chain Fatty Acids (SCFAs)

Several studies established that gut microbial derived SCFAs are central to therapeutic potential against hypertension [192]. SCFAs downregulate cardiac hypertrophy, fibrosis and associated vascular dysfunction through regulatory T cell (Treg) modulation [193]. SCFAs also play a central role in decreasing luminal pH and mucus production and the subsequent maintenance of the gut barrier integrity [183]. SCFAs exercise anti-inflammatory functions and modulate lipid metabolism and gluconeogenesis [194]. SCFAs block cholesterol generation and hepatic localization. These eventually culminate in a long-term impact on cardiometabolic functioning, thereby inhibiting the onset of coronary diseases [194]. The alteration of the gut microbiome dynamics interferes with SCFA production and is attributed to CVD pathophysiology. A decrease in SCFA production compromises its anti-inflammatory functions and culminates in atherogenesis [195]. Accelerated inflammatory cascades lead to cardiovascular injury via the downregulation of Treg-induced IL-10 production [196,197]. SCFA supplementation promotes Treg-associated anti-inflammatory response and reduces the local infiltration of immune cells [193]. This downregulates susceptibility to cardiac arrhythmia and atherosclerotic lesion formation [198]. SCFAs also reverse cardiomyocyte contraction anomalies by promoting the expression of protein phosphatases, endothelial nitric oxide synthase (eNOS), phosphorylated phosphatase and TENsin homologue (PTEN) and glycogen synthase kinase-3β (GSK-3β) [93]. In regulating blood pressure, SCFAs modulate GPR41 and olfactory receptor-78 (OLFR78) receptors [93]. GPR41 is typically a Gαi/o-coupled receptor. SCFA-induced GPR41 activation inhibits the formation of cyclic adenosine monophosphate (cAMP), protein kinase A [94] (Figure 3) and the activation of p38 and extracellular signal-regulated kinase (ERK) [199,200]. The activation of the olfactory receptors leads to an increase in cAMP and Ca2+ production [201,202].

3.2. Trimethylamine-N-Oxide (TMAO)

The gut microbiome also promotes atherosclerosis onset and progression through the enhanced production of TMAO [203]. Gut bacteria, primarily, Deferribacteraceae, Anaeroplasmataceae, Prevotellaceae, and Enterobacteriaceae, contribute towards TMAO formation [204]. TMAO causes vascular injury through the modulation of cholesterol metabolism, promoting ROS production, augmenting vascular endothelial damage, cell junction rupture and an increase in cellular permeability [205,206,207]. Increased TMAO levels promote the adhesion of foam cells and arterial plaque formation, eventually leading to massive cardiometabolic dysfunction and mortality [208]. The molecular signaling mechanisms triggered by TMAO include enhanced oxidative stress and ROS production via nucleotide-binding oligomerization domain Leucine-rich Repeat and Pyrin domain-containing 3 (NLRP3) inflammasome activation [209]. Biochemical studies have shown that TMAO-associated endothelial dysfunction occurs via high mobility group box-1 (HMGB1), an inflammatory mediator [210]. TMAO-induced CVD onset also occurs via the activation of sirtuin 3 and superoxide dismutase 2 (SIRT3-SOD2) ROS signaling. TMAO downregulates SIRT1 expression and augments oxidative stress via the p53/p21/retinoblastoma (Rb) signaling cascade [211]. TMAO also upregulates vascular oxidative stress via a rise in nicotinamide adenine dinucleotide phosphate (NADPH) oxidase activity [94] (Figure 4) [212,213]. TMAO upregulates inflammatory signaling by increasing the level of pro-inflammatory cytokines viz tumor necrosis factor (TNF)-α and IL-1β [214,215]. Augmented TNF-α levels via NF-κB signaling upregulation promotes leukocyte adhesion to endothelial walls, leading to an endothelial dysfunction that triggers thrombosis and atherosclerosis [216,217]. TMAO-associated protein kinase C activation promotes monocyte adhesion and a decrease in the self-repair capacity of the endothelium [218,219]. TMAO induces vascular cell adhesion protein 1 (VCAM1) via the methylation of NF-κB p65 subunit [134,220]. This increases monocyte adhesion through an increase in intercellular adhesion molecule 1 (ICAM1) and E-selectin, enhancing endothelial dysfunction [216,217]. TMAO also upregulates IL-6 levels, which is one of the chief promoters of foam cell formation [221,222]. TMAO plays a pivotal role in the aggregation of low-density lipoproteins (LDLs) in macrophages [122,223]. This occurs via augmenting CD36, lectin-like oxidized low-density lipoprotein receptor 1 (LOX-1) and the scavenger receptors of class A1 (SR-A1) [223]. Together, these facilitate cholesterol uptake and trigger atherosclerotic onset [224,225].

3.3. Bile Acids (BAs)

The gut microbiota also play a pivotal role in regulating cholesterol metabolism by altering the BA levels [226]. In the gut, the primary BAs viz cholic acid (CA) and cheno-DCA (CDCA) undergo deconjugation by the gut microbiome and bile salt hydrolase (BSH) [227,228], producing secondary BAs like DCA, lithocholic acid (LCA) and ursodeoxycholic acid (UDCA) [229]. As mentioned earlier, BAs can activate FXR, and G-protein-coupled bile acid receptor Gpbar-1-, also known as Takeda G protein-coupled receptor 5 (TGR5) [230]. Primary BAs can activate FXR to a greater extent relative to secondary BAs [231]. FXR downregulates BA uptake into the hepatocytes and augments the expression of ATP-binding cassette subfamily B member 11 (ABCB11) [232]. The plasma level ratio of secondary to primary BAs is an important marker for potential hypercholesterolemia and CAD incidence [233]. BAs regulate cardiovascular functioning by decreasing heart rate through the modulation of channel conductance and Ca2+ inflow in ventricular cardiomyocytes [234]. This influences vascular tone and overall cardiovascular function [235]. A reduction in secondary BA levels and a rise in primary BAs hyperactivates FXR, reduces BA production and augments CAD and hypercholesterolemia incidence [236]. Additionally, BA-induced PXR activation augments total cholesterol, very low-density lipoprotein (VLDL), and LDL, and downregulates high density lipoprotein (HDL) promoting atherosclerotic onset [237,238]. PXR activation influences the genes involved in lipoprotein transport and cholesterol metabolism, downregulating apolipoprotein A-IV (ApoA-IV) and cytochrome P-450 family 39 subfamily A member 1 (Cyp39a1) and augmenting CD36 [239]. CD36 upregulation promotes the uptake of oxidized LDL into macrophages and leads to foam cell formation, which precedes atherosclerosis [237].

3.4. Indole-3-Propionic Acid (IPA)

The regulation of blood pressure usually depends on mechanical capabilities of the myocardium and the peripheral vasculature [240]. A rise in blood pressure bears a pathophysiological correlation with chronic cardiac tissue injury and atherosclerosis [241]. IPA, a gut microbial metabolite of dietary tryptophan metabolism, induces myocardial contractility and the vasoconstriction of the endothelium-denuded mesenteric resistance arteries [242,243]. IPA is significantly downregulated in the serum of CAD patients [244,245]. Studies in murine models have demonstrated that dietary IPA intake attenuates atherosclerotic progression under Apolipoprotein E deficient condition [246]. IPA augments ATP-binding cassette class A1 (ABCA1) and enhances macrophage mediated reverse cholesterol transport [247], which attenuates foam cell formation and atherosclerotic onset [248,249,250,251].

3.5. Hydrogen Sulfide (H2S)

Gut-microbiota-derived H2S exhibits varied vascular protective functions by modulating vascular tone and blood pressure, and reducing leukocyte adhesion, platelet aggregate formation, and ROS-mediated oxidative stress [252,253,254]. H2S also augments endothelial cell proliferation, and decreases vascular proliferation, migration, and LDL oxidation [172]. H2S inhibits vascular smooth muscle cell proliferation by the downregulation of CX3CR1 expression in macrophages [255,256,257]. This occurs by promoting peroxisome proliferator-activated receptor (PPAR)-γ, coupled with the downregulation of pro-inflammatory NF-κB/ICAM-1 cascade [258]. The vasodilatory impact of H2S is exercised through a non-endothelial dependent mechanism characterized by the activation of vascular smooth muscle potassium (K+) channels [259,260], which in turn decreases intracellular pH and acidification and prevents atherosclerosis.

3.6. Phenylacetylglutamine (PAGln)

Beyond augmenting CVD risk in type-2 diabetes patients, PAGln has also been reported to bear a strong correlation with an increased risk of heart attack, myocardial infarction and consequent mortality [175,261,262]. In vivo studies suggest that PAGln fosters thrombus formation and a shorter time for blood flow cessation in a carotid artery injury mice model [175,263]. Clinical cohort studies suggest that PAGln augments heart-failure-associated biological activities with extensive PAGln exposure associated with sarcomere functional defects [264]. Pharmacological gain-of-function and loss-of-function studies confirm that PAGln chiefly interacts with α (α2A, α2B) and β (β2)-adrenergic receptors and influences cardiac and platelet functions that eventually culminate to atherosclerotic onset [262,265,266].
Thus, various gut microbial metabolites like SCFAs, TMAO, BAs, IPA and H2S play a key role in the cellular and molecular signaling responsible for modulating the CVDs’ onset and progression (Table 1).

4. Therapeutic Perspectives

The role of gut microbial metabolites in CVD onset and/or attenuation has gained increasing appreciation over this decade. Various strategies to efficiently manipulate these metabolite levels to ameliorate the onset and progression of heart failure, atherosclerosis, and other CVD types have been reported, as mentioned below.

4.1. Dietary and Lifestyle Modifications

Dietary modifications to maintain and improve gut microbial dynamics have been clinically advocated [267,268,269]. Prebiotics constitute the non-digestible carbohydrates that significantly influence gut microbiome composition and function [270]. Investigative studies revealed that regular exercise, the intake of prebiotics and/or probiotics, and associated lifestyle improvements increases the beneficial gut microbial population and decreases the propensity for cardiac fibrosis and hypertrophy [271]. Preclinical studies revealed that prebiotics empower the gut microbiota to retain its dynamics and produce metabolites responsible for cardioprotective functions [272]. This has also promoted the practice of individualized gut-microbiome-targeted approaches to provide new therapeutic strategies against cardiometabolic complications [273]. A high-salt diet reduces Lactobacillus abundance in the gut and augments T helper 17 cell population, resulting in the onset of salt-sensitive hypertension [274,275]. Geographical comparisons reveal that the Western diet bears greater proximity with CVD incidence compared to that of Mediterranean regions [276]. The Mediterranean diet comprises a lower level of saturated fatty acids, salts, and phosphate, and higher amounts of antioxidants, nitrates, and fiber [277,278], which support microbial balance and cause a reduction in oxidative stress, enhanced nitric oxide bioavailability, and antioxidant functions, leading to improved vascular and cardiac functioning [276].

4.2. Fecal Microbial Transplantation (FMT)

Fecal Microbial Transplantation (FMT) introduces fecal contents from healthy subjects into the gastrointestinal (GI) tract of gut-microbiota-compromised patients [279,280]. This technique has been found to be useful, especially in treating infectious episodes of Clostridium difficile invasion into the GI tract [281,282]. Recent courses of studies have identified FMT as a useful tool against the gut-microbiome-associated onset of CVDs [283]. FMT improves plasma triglyceride level and insulin sensitivity in its recipients [284]. This improves overall cardiometabolic health [285]. In experimental study models of autoimmune myocarditis, FMT has been shown to improve Bacteroidetes density and downregulate myocarditis incidence potential [285,286,287,288].

4.3. Pharmacological Interventions

Pharmacotherapy, targeting the composition and dynamics of the gut microbiota, largely contributes towards the attenuation of dysbiosis-associated CVD onset. Biguanides like metformin lower the risk of MI onset in type 2 diabetes patients [289]. Metformin functions by augmenting the peripheral glucose uptake and regulation of the incretin pathway through increasing the plasma and pancreatic levels of GLP-1, the chief molecule driving SCFA-associated cardioprotective function [290,291]. Metformin also downregulates BA absorption through hindering the apical sodium-dependent bile acid transporter (ASBT) [292], thus interfering with primary BA-mediated FXR activation [293,294]. Additionally, metformin augments the SCFA-producing Butyrivibrio, Bifidobacterium bifidium, and Megasphera populations [295,296].
α-glucosidase inhibitors (α-GIs), like acarbose, constitute another group of gut-microbiota-influencing drugs that reduce CVD risk in diabetic patients [297]. α-GIs upregulate bile salt hydrolase (BSH), activating the Lactobacillus and Bifidobacterium populations and downregulating the Bacteroides, Alistipes and Clostridium populations, thereby producing a positive impact on cardiometabolic homeostasis [298]. The pharmacological inhibition of Dipeptidyl Peptidase-4 (DPP-4) downregulates cardiovascular risk factors [299]. DPP-4 inhibition attenuates high-fat-diet-induced dysbiosis—primarily promoting the SCFA-releasing gut bacterial populations and decreasing the Firmicutes to Bacteroidetes ratio [300,301].

4.4. Targeting NLRP3 Inflammasomes

The NLRP3 inflammasome acts as a sensor to deleterious exogenous and endogenous entities to activate pro-inflammatory signaling and culminate in CVDs viz atherosclerosis [302,303]. Studies over the years have suggested that NLRP3 inflammasome is a key predisposing factor for atherosclerosis onset, mediated via cholesterol crystals [304]. Pharmacological studies along these lines suggest that TMAO is a significant triggering factor for NLRP3 inflammasome activation and subsequent endothelial hyperpermeability, endothelial barrier dysfunction and cardiac fibrosis [209,305].
The targeted downregulation of the NLRP3 inflammasome has been advocated across studies to attenuate atherosclerosis onset [306]. Statins, structural analogs of HMG-CoA reductase inhibitors, have been instrumental towards attenuating cardiac hypertrophy, cardiomyocyte apoptosis, endothelial dysfunction and oxidative stress [307,308]. Mechanistic studies in recent years suggest that statins impede TLR4/MyD88/NF-κB signaling which is also associated with NLRP3 inflammasome activation [309]. The suppression of NLRP3 inflammasome via Rosuvastatin has been implicated in the treatment of diabetic cardiomyopathy [310]. Simvastatin has also been implicated in hyperglycemia-associated endothelial dysfunction, downregulating vascular endothelial hyperpermeability and enhancing the expressional levels of tight and adherence junction proteins [311]. This, in turn, prevents NLRP3 inflammasome-mediated HMGB1 release in aortic endothelial cells [311]. Beyond statins, the use of Arglabin, a natural NLRP3 inflammasome inhibitor, has been found to exert anti-atherogenic effects in high-fat-diet mice model studies [312]. Together, these approaches strongly point towards NLRP3 inflammasome as a promising target for attenuating CVD onset and progression.

5. Current Limitations and Future Perspectives

Over the years, gut microbiota associated dysbiosis has been the focus of designing therapeutic strategies against CVD risk. Pharmacological interference, FMT, and dietary modifications have achieved considerable success in decreasing dysbiosis-induced CVD onset. However, existing limitations across these studies maintain a significant scientific gap and impede a clear understanding of the dysbiosis-associated CVD pathophysiological status.
From a pharmacological perspective, the cardioprotective impact of metformin in non-diabetic patients remain largely obscure [313]. A double-blinded placebo-controlled study with a small sample size postulates that metformin potentially improves vascular function and reduces ischemic risk in angina patients [314]. However, pilot clinical studies have mentioned no significant impact of metformin on other surrogate markers for CVD, like carotid intimal media thickness [315,316]. Similarly, the cardioprotective function of α-GIs in non-diabetic subjects remains debatable. Along these lines, studies reported that CAD patients with impaired glucose tolerance experienced no significant impact on cardiometabolic upliftment upon acarbose administration [317]. The inhibition of sodium glucose co-transporter 2 (SGT2) reduces cardiovascular inflammation, cellular apoptosis, and mitochondrial dysfunction [318]. However, its impact on gut microbial homeostasis remains largely unclear [319,320,321].
FMT constitutes an intriguing therapeutic strategy against gut-microbial-dysbiosis-driven extra-intestinal complications [284]. However, its utility remains largely restricted, owing to the associated risk of endotoxin and/or infectious substance transfer from the donor to the recipient [322,323]. The dietary intake of prebiotics strengthens gut microbial dynamics against CVD onset. However, such non-digestible carbohydrates increase the free sulfate concentrations in the colon [324], disrupting the mitochondrial complex IV function which, in turn, decreases butyrate oxidation and oxygen consumption [325]. Moreover, the emulsification property of these polysaccharides disturbs the gut microbial composition and augments bacterial translocation across the epithelium [326,327]. Considering the limitations of the existing strategies, it is plausible to say that it is important to carry out further studies to fill in the knowledge gaps in this area.
Gut microbial metabolites play a vital role in the onset as well as the attenuation of CVD, largely depending on the biochemical characteristics of the metabolite. In view of the beneficial metabolites and their cardioprotective impact, clustered regularly interspaced short palindromic repeats (CRISPR)/Cas9 technology can be advocated to alter the metabolic flow and ensure the increased and targeted expression of the cardioprotective metabolites [328,329]. Bacterial metagenomic and proteomic analysis can also be potentially combined with human metabolome to better understand the gut-microbiota-driven metabolic changes in CVD onset [330,331]. To confront the limitations associated with FMT, a transplantation of a specific group of bacterial population can be undertaken, instead of the fecal matter transplantations carried out so far [332]. From a pharmacological perspective, the development of greater numbers of cytostatic agents needs to be focused upon, instead of cytotoxic drugs, to ensure the gut microbial steady state and target specificity [333].

6. Conclusions

Several studies over the last few decades have proven the pivotal role of gut microbiome and its metabolites in cardiovascular health and disease, which has enabled a better understanding of the physiological and mechanistic alterations they drive in CVD onset and progression. The gut microbial breakdown of dietary food components produces a large variety of bioactive metabolites viz SCFAs, BA, TMAO, IPA, phenolic compounds, and hydrogen sulfide, among many. Precision medicine-based studies have promoted personalized, integrated and patient-specific approaches for CVD therapy. In this milieu, examining the microbiome of individual patients and analyzing their gut metabolites in plasma could provide us with necessary information to identify the detrimental effects of gut metabolites on cardiovascular health and equip us with personalized treatments. An in-depth comprehension of the signaling cross-talks involved in the gut–heart axis and alternative strategies to improve the existing therapeutic courses in relation to gut microbial metabolites shall potentially benefit the development of effective druggable targets and mitigate CVD over the long term.

Author Contributions

Conceptualization, S.D., K.M.B. and S.K.; writing—original draft preparation, S.D., writing—review and editing, S.D., S.I., S.P., K.M.B. and S.K.; funding acquisition, S.K. and K.M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institute of Health grant number R01HL148711.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Global Cardiovascular Risk Consortium. Global effect of modifiable risk factors on cardiovascular disease and mortality. N. Engl. J. Med. 2023, 389, 1273–1285. [Google Scholar] [CrossRef] [PubMed]
  2. World Health Organization. Cardiovascular Diseases. Available online: https://www.who.int/health-topics/cardiovascular-diseases#tab=tab_1 (accessed on 13 February 2024).
  3. Dahlöf, B. Cardiovascular disease risk factors: Epidemiology and risk assessment. Am. J. Cardiol. 2010, 105, 3A–9A. [Google Scholar] [CrossRef] [PubMed]
  4. Teo, K.K.; Rafiq, T. Cardiovascular risk factors and prevention: A perspective from developing countries. Can. J. Cardiol. 2021, 37, 733–743. [Google Scholar] [CrossRef] [PubMed]
  5. Katta, N.; Loethen, T.; Lavie, C.J.; Alpert, M.A. Obesity and coronary heart disease: Epidemiology, pathology, and coronary artery imaging. Curr. Probl. Cardiol. 2021, 46, 100655. [Google Scholar] [CrossRef]
  6. Liang, F.; Wang, Y. Coronary heart disease and atrial fibrillation: A vicious cycle. Am. J. Physiol.-Heart Circ. Physiol. 2021, 320, H1–H12. [Google Scholar] [CrossRef]
  7. Dzator, J.S.; Howe, P.R.; Wong, R.H. Profiling cerebrovascular function in migraine: A systematic review and meta-analysis. J. Cereb. Blood Flow Metab. 2021, 41, 919–944. [Google Scholar] [CrossRef]
  8. Borja, A.J.; Hancin, E.C.; Zhang, V.; Revheim, M.-E.; Alavi, A. Potential of PET/CT in Assessing Dementias with Emphasis on Cerebrovascular Disorders; Springer: Berlin/Heidelberg, Germany, 2020; Volume 47, pp. 2493–2498. [Google Scholar]
  9. Kim, K.; Anderson, E.M.; Scali, S.T.; Ryan, T.E. Skeletal muscle mitochondrial dysfunction and oxidative stress in peripheral arterial disease: A unifying mechanism and therapeutic target. Antioxidants 2020, 9, 1304. [Google Scholar] [CrossRef]
  10. Aday, A.W.; Matsushita, K. Epidemiology of peripheral artery disease and polyvascular disease. Circ. Res. 2021, 128, 1818–1832. [Google Scholar] [CrossRef] [PubMed]
  11. Passos, L.S.; Nunes, M.C.P.; Aikawa, E. Rheumatic heart valve disease pathophysiology and underlying mechanisms. Front. Cardiovasc. Med. 2021, 7, 612716. [Google Scholar] [CrossRef]
  12. Marijon, E.; Mocumbi, A.; Narayanan, K.; Jouven, X.; Celermajer, D.S. Persisting burden and challenges of rheumatic heart disease. Eur. Heart J. 2021, 42, 3338–3348. [Google Scholar] [CrossRef]
  13. Morton, S.U.; Quiat, D.; Seidman, J.G.; Seidman, C.E. Genomic frontiers in congenital heart disease. Nat. Rev. Cardiol. 2022, 19, 26–42. [Google Scholar] [CrossRef] [PubMed]
  14. Saef, J.M.; Ghobrial, J. Valvular heart disease in congenital heart disease: A narrative review. Cardiovasc. Diagn. Ther. 2021, 11, 818. [Google Scholar] [CrossRef] [PubMed]
  15. Mortensen, M.B.; Nordestgaard, B.G. Elevated LDL cholesterol and increased risk of myocardial infarction and atherosclerotic cardiovascular disease in individuals aged 70–100 years: A contemporary primary prevention cohort. Lancet 2020, 396, 1644–1652. [Google Scholar] [CrossRef] [PubMed]
  16. Mohammadi, H.; Ohm, J.; Discacciati, A.; Sundstrom, J.; Hambraeus, K.; Jernberg, T.; Svensson, P. Abdominal obesity and the risk of recurrent atherosclerotic cardiovascular disease after myocardial infarction. Eur. J. Prev. Cardiol. 2020, 27, 1944–1952. [Google Scholar] [CrossRef]
  17. Chandarana, M.; Curtis, A.; Hoskins, C. The use of nanotechnology in cardiovascular disease. Appl. Nanosci. 2018, 8, 1607–1619. [Google Scholar] [CrossRef]
  18. Kaptein, F.; Kroft, L.; Hammerschlag, G.; Ninaber, M.; Bauer, M.; Huisman, M.; Klok, F. Pulmonary infarction in acute pulmonary embolism. Thromb. Res. 2021, 202, 162–169. [Google Scholar] [CrossRef]
  19. Ortega-Paz, L.; Talasaz, A.H.; Sadeghipour, P.; Potpara, T.S.; Aronow, H.D.; Jara-Palomares, L.; Sholzberg, M.; Angiolillo, D.J.; Lip, G.Y.; Bikdeli, B. COVID-19-Associated Pulmonary Embolism: Review of the Pathophysiology, Epidemiology, Prevention, Diagnosis, and Treatment. In Seminars in Thrombosis and Hemostasis; Thieme Medical Publishers, Inc.: New York, NY, USA, 2022. [Google Scholar]
  20. Zhao, Y.; Ransom, J.F.; Li, A.; Vedantham, V.; von Drehle, M.; Muth, A.N.; Tsuchihashi, T.; McManus, M.T.; Schwartz, R.J.; Srivastava, D. Dysregulation of cardiogenesis, cardiac conduction, and cell cycle in mice lacking miRNA-1-2. Cell 2007, 129, 303–317. [Google Scholar] [CrossRef]
  21. Laflamme, M.A.; Murry, C.E. Heart regeneration. Nature 2011, 473, 326–335. [Google Scholar] [CrossRef]
  22. Prajnamitra, R.P.; Chen, H.-C.; Lin, C.-J.; Chen, L.-L.; Hsieh, P.C.-H. Nanotechnology approaches in tackling cardiovascular diseases. Molecules 2019, 24, 2017. [Google Scholar] [CrossRef]
  23. Smith, B.R.; Edelman, E.R. Nanomedicines for cardiovascular disease. Nat. Cardiovasc. Res. 2023, 2, 351–367. [Google Scholar] [CrossRef]
  24. Soumya, R.S.; Raghu, K.G. Recent advances on nanoparticle-based therapies for cardiovascular diseases. J. Cardiol. 2023, 81, 10–18. [Google Scholar] [CrossRef] [PubMed]
  25. Stefanadis, C.; Tsiamis, E.; Vlachopoulos, C.; Stratos, C.; Toutouzas, K.; Pitsavos, C.; Marakas, S.; Boudoulas, H.; Toutouzas, P. Unfavorable effect of smoking on the elastic properties of the human aorta. Circulation 1997, 95, 31–38. [Google Scholar] [CrossRef]
  26. Barua, R.S.; Ambrose, J.A.; Eales-Reynolds, L.-J.; DeVoe, M.C.; Zervas, J.G.; Saha, D.C. Dysfunctional endothelial nitric oxide biosynthesis in healthy smokers with impaired endothelium-dependent vasodilatation. Circulation 2001, 104, 1905–1910. [Google Scholar] [CrossRef] [PubMed]
  27. Lorenzatti, A.J.; Servato, M.L. New evidence on the role of inflammation in CVD risk. Curr. Opin. Cardiol. 2019, 34, 418–423. [Google Scholar] [CrossRef] [PubMed]
  28. Moore, K.J. Targeting inflammation in CVD: Advances and challenges. Nat. Rev. Cardiol. 2019, 16, 74–75. [Google Scholar] [CrossRef]
  29. Isordia-Salas, I.; Galván-Plata, M.E.; Leaños-Miranda, A.; Aguilar-Sosa, E.; Anaya-Gómez, F.; Majluf-Cruz, A.; Santiago-Germán, D. Proinflammatory and prothrombotic state in subjects with different glucose tolerance status before cardiovascular disease. J. Diabetes Res. 2014, 2014, 631902. [Google Scholar] [CrossRef]
  30. Nordestgaard, B.G.; Chapman, M.J.; Ray, K.; Borén, J.; Andreotti, F.; Watts, G.F.; Ginsberg, H.; Amarenco, P.; Catapano, A.; Descamps, O.S. Lipoprotein (a) as a cardiovascular risk factor: Current status. Eur. Heart J. 2010, 31, 2844–2853. [Google Scholar] [CrossRef]
  31. Li, Y.; Zhai, H.; Kang, L.; Chu, Q.; Zhao, X.; Li, R. Causal association between basal metabolic rate and risk of cardiovascular diseases: A univariable and multivariable Mendelian randomization study. Sci. Rep. 2023, 13, 12487. [Google Scholar] [CrossRef]
  32. Chen, K.; Zhang, Y.; Zhou, S.; Jin, C.; Xiang, M.; Ma, H. The association between the basal metabolic rate and cardiovascular disease: A two-sample Mendelian randomization study. Eur. J. Clin. Investig. 2024, 54, e14153. [Google Scholar] [CrossRef]
  33. Deprince, A.; Haas, J.T.; Staels, B. Dysregulated lipid metabolism links NAFLD to cardiovascular disease. Mol. Metab. 2020, 42, 101092. [Google Scholar] [CrossRef]
  34. Han, W.; Yang, S.; Xiao, H.; Wang, M.; Ye, J.; Cao, L.; Sun, G. Role of adiponectin in cardiovascular diseases related to glucose and lipid metabolism disorders. Int. J. Mol. Sci. 2022, 23, 15627. [Google Scholar] [CrossRef] [PubMed]
  35. Lu, Y.; Zhou, S.; Dreyer, R.P.; Spatz, E.S.; Geda, M.; Lorenze, N.P.; D’Onofrio, G.; Lichtman, J.H.; Spertus, J.A.; Ridker, P.M. Sex differences in inflammatory markers and health status among young adults with acute myocardial infarction: Results from the VIRGO (Variation in Recovery: Role of Gender on Outcomes of Young Acute Myocardial Infarction Patients) study. Circ. Cardiovasc. Qual. Outcomes 2017, 10, e003470. [Google Scholar] [CrossRef]
  36. Bohula, E.A.; Giugliano, R.P.; Cannon, C.P.; Zhou, J.; Murphy, S.A.; White, J.A.; Tershakovec, A.M.; Blazing, M.A.; Braunwald, E. Achievement of dual low-density lipoprotein cholesterol and high-sensitivity C-reactive protein targets more frequent with the addition of ezetimibe to simvastatin and associated with better outcomes in IMPROVE-IT. Circulation 2015, 132, 1224–1233. [Google Scholar] [CrossRef] [PubMed]
  37. Ridker, P.M.; Morrow, D.A.; Rose, L.M.; Rifai, N.; Cannon, C.P.; Braunwald, E. Relative efficacy of atorvastatin 80 mg and pravastatin 40 mg in achieving the dual goals of low-density lipoprotein cholesterol< 70 mg/dl and C-reactive protein< 2 mg/L: An analysis of the PROVE-IT TIMI-22 trial. J. Am. Coll. Cardiol. 2005, 45, 1644–1648. [Google Scholar]
  38. Sethi, Y.; Patel, N.; Kaka, N.; Kaiwan, O.; Kar, J.; Moinuddin, A.; Goel, A.; Chopra, H.; Cavalu, S. Precision Medicine and the future of Cardiovascular Diseases: A Clinically Oriented Comprehensive Review. J. Clin. Med. 2023, 12, 1799. [Google Scholar] [CrossRef] [PubMed]
  39. Thursby, E.; Juge, N. Introduction to the human gut microbiota. Biochem. J. 2017, 474, 1823–1836. [Google Scholar] [CrossRef] [PubMed]
  40. Mackie, R.I.; Sghir, A.; Gaskins, H.R. Developmental microbial ecology of the neonatal gastrointestinal tract. Am. J. Clin. Nutr. 1999, 69, 1035s–1045s. [Google Scholar] [CrossRef]
  41. Dubourg, G.; Lagier, J.-C.; Armougom, F.; Robert, C.; Audoly, G.; Papazian, L.; Raoult, D. High-level colonisation of the human gut by Verrucomicrobia following broad-spectrum antibiotic treatment. Int. J. Antimicrob. Agents 2013, 41, 149–155. [Google Scholar] [CrossRef]
  42. Grigor’eva, I.N. Gallstone disease, obesity and the Firmicutes/Bacteroidetes ratio as a possible biomarker of gut dysbiosis. J. Pers. Med. 2020, 11, 13. [Google Scholar] [CrossRef]
  43. Bai, J.; Hu, Y.; Bruner, D. Composition of gut microbiota and its association with body mass index and lifestyle factors in a cohort of 7–18 years old children from the American Gut Project. Pediatr. Obes. 2019, 14, e12480. [Google Scholar] [CrossRef]
  44. Ojima, M.; Motooka, D.; Shimizu, K.; Gotoh, K.; Shintani, A.; Yoshiya, K.; Nakamura, S.; Ogura, H.; Iida, T.; Shimazu, T. Metagenomic analysis reveals dynamic changes of whole gut microbiota in the acute phase of intensive care unit patients. Dig. Dis. Sci. 2016, 61, 1628–1634. [Google Scholar] [CrossRef] [PubMed]
  45. De la Cuesta-Zuluaga, J.; Escobar, J.S. Considerations for optimizing microbiome analysis using a marker gene. Front. Nutr. 2016, 3, 26. [Google Scholar] [CrossRef]
  46. Jovel, J.; Nimaga, A.; Jordan, T.; O’keefe, S.; Patterson, J.; Thiesen, A.; Hotte, N.; Bording-Jorgensen, M.; Subedi, S.; Hamilton, J. Metagenomics versus metatranscriptomics of the murine gut microbiome for assessing microbial metabolism during inflammation. Front. Microbiol. 2022, 13, 119. [Google Scholar] [CrossRef]
  47. Li, X.-Y.; He, C.; Zhu, Y.; Lu, N.-H. Role of gut microbiota on intestinal barrier function in acute pancreatitis. World J. Gastroenterol. 2020, 26, 2187. [Google Scholar] [CrossRef] [PubMed]
  48. Ottman, N.; Reunanen, J.; Meijerink, M.; Pietilä, T.E.; Kainulainen, V.; Klievink, J.; Huuskonen, L.; Aalvink, S.; Skurnik, M.; Boeren, S. Pili-like proteins of Akkermansia muciniphila modulate host immune responses and gut barrier function. PLoS ONE 2017, 12, e0173004. [Google Scholar] [CrossRef]
  49. Li, X.; Liu, L.; Cao, Z.; Li, W.; Li, H.; Lu, C.; Yang, X.; Liu, Y. Gut microbiota as an “invisible organ” that modulates the function of drugs. Biomed. Pharmacother. 2020, 121, 109653. [Google Scholar] [CrossRef]
  50. García-Ríos, A.; Garcia, A.C.; Perez-Jimenez, F.; Perez-Martinez, P. Gut microbiota: A new protagonist in the risk of cardiovascular disease? Clínica E Investig. En Arterioscler. (Engl. Ed.) 2019, 31, 178–185. [Google Scholar] [CrossRef]
  51. Kamada, N.; Seo, S.-U.; Chen, G.Y.; Núñez, G. Role of the gut microbiota in immunity and inflammatory disease. Nat. Rev. Immunol. 2013, 13, 321–335. [Google Scholar] [CrossRef]
  52. Qiao, R.; Deng, Y.; Zhang, S.; Wolosker, M.B.; Zhu, Q.; Ren, H.; Zhang, Y. Accumulation of different shapes of microplastics initiates intestinal injury and gut microbiota dysbiosis in the gut of zebrafish. Chemosphere 2019, 236, 124334. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, H.; Chen, Y.; Wang, Z.; Xie, G.; Liu, M.; Yuan, B.; Chai, H.; Wang, W.; Cheng, P. Implications of gut microbiota in neurodegenerative diseases. Front. Immunol. 2022, 13, 785644. [Google Scholar] [CrossRef]
  54. Karlsson, F.; Tremaroli, V.; Nielsen, J.; Bäckhed, F. Assessing the human gut microbiota in metabolic diseases. Diabetes 2013, 62, 3341–3349. [Google Scholar] [CrossRef] [PubMed]
  55. Nicholson, J.K.; Holmes, E.; Kinross, J.; Burcelin, R.; Gibson, G.; Jia, W.; Pettersson, S. Host-gut microbiota metabolic interactions. Science 2012, 336, 1262–1267. [Google Scholar] [CrossRef]
  56. Silva, Y.P.; Bernardi, A.; Frozza, R.L. The role of short-chain fatty acids from gut microbiota in gut-brain communication. Front. Endocrinol. 2020, 11, 25. [Google Scholar] [CrossRef]
  57. Oliphant, K.; Allen-Vercoe, E. Macronutrient metabolism by the human gut microbiome: Major fermentation by-products and their impact on host health. Microbiome 2019, 7, 91. [Google Scholar] [CrossRef]
  58. Augusti, P.R.; Conterato, G.M.; Denardin, C.C.; Prazeres, I.D.; Serra, A.T.; Bronze, M.R.; Emanuelli, T. Bioactivity, bioavailability, and gut microbiota transformations of dietary phenolic compounds: Implications for COVID-19. J. Nutr. Biochem. 2021, 97, 108787. [Google Scholar] [CrossRef]
  59. Gui, D.-D.; Luo, W.; Yan, B.-J.; Ren, Z.; Tang, Z.-H.; Liu, L.-S.; Zhang, J.-F.; Jiang, Z.-S. Effects of gut microbiota on atherosclerosis through hydrogen sulfide. Eur. J. Pharmacol. 2021, 896, 173916. [Google Scholar] [CrossRef] [PubMed]
  60. Kasprzak-Drozd, K.; Oniszczuk, T.; Stasiak, M.; Oniszczuk, A. Beneficial effects of phenolic compounds on gut microbiota and metabolic syndrome. Int. J. Mol. Sci. 2021, 22, 3715. [Google Scholar] [CrossRef]
  61. Meyer, K.A.; Bennett, B.J. Diet and gut microbial function in metabolic and cardiovascular disease risk. Curr. Diabetes Rep. 2016, 16, 93. [Google Scholar] [CrossRef]
  62. Zhu, Y.; Shui, X.; Liang, Z.; Huang, Z.; Qi, Y.; He, Y.; Chen, C.; Luo, H.; Lei, W. Gut microbiota metabolites as integral mediators in cardiovascular diseases. Int. J. Mol. Med. 2020, 46, 936–948. [Google Scholar] [CrossRef] [PubMed]
  63. Xu, H.; Wang, X.; Feng, W.; Liu, Q.; Zhou, S.; Liu, Q.; Cai, L. The gut microbiota and its interactions with cardiovascular disease. Microb. Biotechnol. 2020, 13, 637–656. [Google Scholar] [CrossRef]
  64. Tang, W.W.; Li, D.Y.; Hazen, S.L. Dietary metabolism, the gut microbiome, and heart failure. Nat. Rev. Cardiol. 2019, 16, 137–154. [Google Scholar] [CrossRef] [PubMed]
  65. Brial, F.; Le Lay, A.; Dumas, M.-E.; Gauguier, D. Implication of gut microbiota metabolites in cardiovascular and metabolic diseases. Cell. Mol. Life Sci. 2018, 75, 3977–3990. [Google Scholar] [CrossRef]
  66. Branchereau, M.; Burcelin, R.; Heymes, C. The gut microbiome and heart failure: A better gut for a better heart. Rev. Endocr. Metab. Disord. 2019, 20, 407–414. [Google Scholar] [CrossRef]
  67. 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. [Google Scholar] [CrossRef]
  68. 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. A gut bacterial pathway metabolizes aromatic amino acids into nine circulating metabolites. Nature 2017, 551, 648–652. [Google Scholar] [CrossRef]
  69. Claus, S.P.; Tsang, T.M.; Wang, Y.; Cloarec, O.; Skordi, E.; Martin, F.P.; Rezzi, S.; Ross, A.; Kochhar, S.; Holmes, E. Systemic multicompartmental effects of the gut microbiome on mouse metabolic phenotypes. Mol. Syst. Biol. 2008, 4, 219. [Google Scholar] [CrossRef]
  70. Deschasaux, M.; Bouter, K.E.; Prodan, A.; Levin, E.; Groen, A.K.; Herrema, H.; Tremaroli, V.; Bakker, G.J.; Attaye, I.; Pinto-Sietsma, S.-J. Depicting the composition of gut microbiota in a population with varied ethnic origins but shared geography. Nat. Med. 2018, 24, 1526–1531. [Google Scholar] [CrossRef] [PubMed]
  71. Jha, A.R.; Davenport, E.R.; Gautam, Y.; Bhandari, D.; Tandukar, S.; Ng, K.M.; Fragiadakis, G.K.; Holmes, S.; Gautam, G.P.; Leach, J. Gut microbiome transition across a lifestyle gradient in Himalaya. PLoS Biol. 2018, 16, e2005396. [Google Scholar] [CrossRef]
  72. Smits, S.A.; Leach, J.; Sonnenburg, E.D.; Gonzalez, C.G.; Lichtman, J.S.; Reid, G.; Knight, R.; Manjurano, A.; Changalucha, J.; Elias, J.E. Seasonal cycling in the gut microbiome of the Hadza hunter-gatherers of Tanzania. Science 2017, 357, 802–806. [Google Scholar] [CrossRef]
  73. Martin, F.-P.J.; Sprenger, N.; Yap, I.K.; Wang, Y.; Bibiloni, R.; Rochat, F.; Rezzi, S.; Cherbut, C.; Kochhar, S.; Lindon, J.C. Panorganismal gut microbiome− host metabolic crosstalk. J. Proteome Res. 2009, 8, 2090–2105. [Google Scholar] [CrossRef]
  74. Sekirov, I.; Russell, S.L.; Antunes, L.C.M.; Finlay, B.B. Gut microbiota in health and disease. Physiol. Rev. 2010, 90, 859–904. [Google Scholar] [CrossRef]
  75. Nallu, A.; Sharma, S.; Ramezani, A.; Muralidharan, J.; Raj, D. Gut microbiome in chronic kidney disease: Challenges and opportunities. Transl. Res. 2017, 179, 24–37. [Google Scholar] [CrossRef]
  76. Juanola, O.; Ferrusquía-Acosta, J.; García-Villalba, R.; Zapater, P.; Magaz, M.; Marín, A.; Olivas, P.; Baiges, A.; Bellot, P.; Turon, F. Circulating levels of butyrate are inversely related to portal hypertension, endotoxemia, and systemic inflammation in patients with cirrhosis. FASEB J. 2019, 33, 11595–11605. [Google Scholar] [CrossRef]
  77. Cummings, J.H.; Pomare, E.; Branch, W.; Naylor, C.; MacFarlane, G. Short chain fatty acids in human large intestine, portal, hepatic and venous blood. Gut 1987, 28, 1221. [Google Scholar] [CrossRef]
  78. Fukuda, S.; Toh, H.; Hase, K.; Oshima, K.; Nakanishi, Y.; Yoshimura, K.; Tobe, T.; Clarke, J.M.; Topping, D.L.; Suzuki, T. Bifidobacteria can protect from enteropathogenic infection through production of acetate. Nature 2011, 469, 543–547. [Google Scholar] [CrossRef]
  79. Donohoe, D.R.; Garge, N.; Zhang, X.; Sun, W.; O‘Connell, T.M.; Bunger, M.K.; Bultman, S.J. The microbiome and butyrate regulate energy metabolism and autophagy in the mammalian colon. Cell Metab. 2011, 13, 517–526. [Google Scholar] [CrossRef]
  80. Lin, H.V.; Frassetto, A.; Kowalik, E.J., Jr.; Nawrocki, A.R.; Lu, M.M.; Kosinski, J.R.; Hubert, J.A.; Szeto, D.; Yao, X.; Forrest, G. Butyrate and propionate protect against diet-induced obesity and regulate gut hormones via free fatty acid receptor 3-independent mechanisms. PLoS ONE 2012, 7, e35240. [Google Scholar] [CrossRef]
  81. Gao, Z.; Yin, J.; Zhang, J.; Ward, R.E.; Martin, R.J.; Lefevre, M.; Cefalu, W.T.; Ye, J. Butyrate improves insulin sensitivity and increases energy expenditure in mice. Diabetes 2009, 58, 1509–1517. [Google Scholar] [CrossRef]
  82. Chambers, E.S.; Viardot, A.; Psichas, A.; Morrison, D.J.; Murphy, K.G.; Zac-Varghese, S.E.; MacDougall, K.; Preston, T.; Tedford, C.; Finlayson, G.S. Effects of targeted delivery of propionate to the human colon on appetite regulation, body weight maintenance and adiposity in overweight adults. Gut 2014, 64, 1744–1754. [Google Scholar] [CrossRef]
  83. Hong, Y.-H.; Nishimura, Y.; Hishikawa, D.; Tsuzuki, H.; Miyahara, H.; Gotoh, C.; Choi, K.-C.; Feng, D.D.; Chen, C.; Lee, H.-G. Acetate and propionate short chain fatty acids stimulate adipogenesis via GPCR43. Endocrinology 2005, 146, 5092–5099. [Google Scholar] [CrossRef]
  84. Ge, H.; Li, X.; Weiszmann, J.; Wang, P.; Baribault, H.; Chen, J.-L.; Tian, H.; Li, Y. Activation of G protein-coupled receptor 43 in adipocytes leads to inhibition of lipolysis and suppression of plasma free fatty acids. Endocrinology 2008, 149, 4519–4526. [Google Scholar] [CrossRef]
  85. Kimura, I.; Ozawa, K.; Inoue, D.; Imamura, T.; Kimura, K.; Maeda, T.; Terasawa, K.; Kashihara, D.; Hirano, K.; Tani, T. The gut microbiota suppresses insulin-mediated fat accumulation via the short-chain fatty acid receptor GPR43. Nat. Commun. 2013, 4, 1829. [Google Scholar] [CrossRef]
  86. Maslowski, K.M.; Vieira, A.T.; Ng, A.; Kranich, J.; Sierro, F.; Yu, D.; Schilter, H.C.; Rolph, M.S.; Mackay, F.; Artis, D. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 2009, 461, 1282–1286. [Google Scholar] [CrossRef]
  87. Smith, P.M.; Howitt, M.R.; Panikov, N.; Michaud, M.; Gallini, C.A.; Bohlooly-y, M.; Glickman, J.N.; Garrett, W.S. The microbial metabolites, short-chain fatty acids, regulate colonic Treg cell homeostasis. Science 2013, 341, 569–573. [Google Scholar] [CrossRef]
  88. Brown, A.J.; Goldsworthy, S.M.; Barnes, A.A.; Eilert, M.M.; Tcheang, L.; Daniels, D.; Muir, A.I.; Wigglesworth, M.J.; Kinghorn, I.; Fraser, N.J. The Orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. J. Biol. Chem. 2003, 278, 11312–11319. [Google Scholar] [CrossRef]
  89. Tazoe, H.; Otomo, Y.; Karaki, S.-i.; Kato, I.; Fukami, Y.; Terasaki, M.; Kuwahara, A. Expression of short-chain fatty acid receptor GPR41 in the human colon. Biomed. Res. 2009, 30, 149–156. [Google Scholar] [CrossRef]
  90. Singh, N.; Gurav, A.; Sivaprakasam, S.; Brady, E.; Padia, R.; Shi, H.; Thangaraju, M.; Prasad, P.D.; Manicassamy, S.; Munn, D.H. Activation of Gpr109a, receptor for niacin and the commensal metabolite butyrate, suppresses colonic inflammation and carcinogenesis. Immunity 2014, 40, 128–139. [Google Scholar] [CrossRef]
  91. Feingold, K.R.; Moser, A.; Shigenaga, J.K.; Grunfeld, C. Inflammation stimulates niacin receptor (GPR109A/HCA2) expression in adipose tissue and macrophages. J. Lipid Res. 2014, 55, 2501–2508. [Google Scholar] [CrossRef]
  92. Lukasova, M.; Malaval, C.; Gille, A.; Kero, J.; Offermanns, S. Nicotinic acid inhibits progression of atherosclerosis in mice through its receptor GPR109A expressed by immune cells. J. Clin. Investig. 2011, 121, 1163–1173. [Google Scholar] [CrossRef]
  93. Pluznick, J.L.; Protzko, R.J.; Gevorgyan, H.; Peterlin, Z.; Sipos, A.; Han, J.; Brunet, I.; Wan, L.-X.; Rey, F.; Wang, T. Olfactory receptor responding to gut microbiota-derived signals plays a role in renin secretion and blood pressure regulation. Proc. Natl. Acad. Sci. USA 2013, 110, 4410–4415. [Google Scholar] [CrossRef]
  94. Biorender. Create Professional Science Figures in Minutes. Available online: https://www.biorender.com/ (accessed on 25 February 2024).
  95. Fuchs, C.D.; Trauner, M. Role of bile acids and their receptors in gastrointestinal and hepatic pathophysiology. Nat. Rev. Gastroenterol. Hepatol. 2022, 19, 432–450. [Google Scholar] [CrossRef]
  96. Thakare, R.; Alamoudi, J.A.; Gautam, N.; Rodrigues, A.D.; Alnouti, Y. Species differences in bile acids I. Plasma and urine bile acid composition. J. Appl. Toxicol. 2018, 38, 1323–1335. [Google Scholar] [CrossRef]
  97. Wahlström, A.; Kovatcheva-Datchary, P.; Ståhlman, M.; Khan, M.-T.; Bäckhed, F.; Marschall, H.-U. Induction of farnesoid X receptor signaling in germ-free mice colonized with a human microbiota. J. Lipid Res. 2017, 58, 412–419. [Google Scholar] [CrossRef]
  98. Hamilton, J.P.; Xie, G.; Raufman, J.-P.; Hogan, S.; Griffin, T.L.; Packard, C.A.; Chatfield, D.A.; Hagey, L.R.; Steinbach, J.H.; Hofmann, A.F. Human cecal bile acids: Concentration and spectrum. Am. J. Physiol.-Gastrointest. Liver Physiol. 2007, 293, G256–G263. [Google Scholar] [CrossRef]
  99. Jones, B.V.; Begley, M.; Hill, C.; Gahan, C.G.; Marchesi, J.R. Functional and comparative metagenomic analysis of bile salt hydrolase activity in the human gut microbiome. Proc. Natl. Acad. Sci. USA 2008, 105, 13580–13585. [Google Scholar] [CrossRef]
  100. Van Best, N.; Rolle-Kampczyk, U.; Schaap, F.; Basic, M.; Olde Damink, S.; Bleich, A.; Savelkoul, P.; Von Bergen, M.; Penders, J.; Hornef, M. Bile acids drive the newborn’s gut microbiota maturation. Nat. Commun. 2020, 11, 3692. [Google Scholar] [CrossRef]
  101. Ridlon, J.M.; Kang, D.J.; Hylemon, P.B.; Bajaj, J.S. Bile acids and the gut microbiome. Curr. Opin. Gastroenterol. 2014, 30, 332–338. [Google Scholar] [CrossRef]
  102. Begley, M.; Gahan, C.G.; Hill, C. The interaction between bacteria and bile. FEMS Microbiol. Rev. 2005, 29, 625–651. [Google Scholar] [CrossRef]
  103. Miller, S.I. Antibiotic resistance and regulation of the gram-negative bacterial outer membrane barrier by host innate immune molecules. mBio 2016, 7, e01541-16. [Google Scholar] [CrossRef]
  104. Li, Y.; Tang, R.; Leung, P.S.; Gershwin, M.E.; Ma, X. Bile acids and intestinal microbiota in autoimmune cholestatic liver diseases. Autoimmun. Rev. 2017, 16, 885–896. [Google Scholar] [CrossRef]
  105. Theriot, C.M.; Bowman, A.A.; Young, V.B. Antibiotic-induced alterations of the gut microbiota alter secondary bile acid production and allow for Clostridium difficile spore germination and outgrowth in the large intestine. mSphere 2016, 1, e00045-15. [Google Scholar] [CrossRef]
  106. Duboc, H.; Rajca, S.; Rainteau, D.; Benarous, D.; Maubert, M.-A.; Quervain, E.; Thomas, G.; Barbu, V.; Humbert, L.; Despras, G. Connecting dysbiosis, bile-acid dysmetabolism and gut inflammation in inflammatory bowel diseases. Gut 2013, 62, 531–539. [Google Scholar] [CrossRef]
  107. Vavassori, P.; Mencarelli, A.; Renga, B.; Distrutti, E.; Fiorucci, S. The bile acid receptor FXR is a modulator of intestinal innate immunity. J. Immunol. 2009, 183, 6251–6261. [Google Scholar] [CrossRef]
  108. Mencarelli, A.; Renga, B.; Palladino, G.; Ricci, P.; Distrutti, E.; Barbanti, M.; Baldelli, F.; Fiorucci, S. Inhibition of NF-κB by a PXR-dependent pathway mediates counter-regulatory activities of rifaximin on innate immunity in intestinal epithelial cells. Eur. J. Pharmacol. 2011, 668, 317–324. [Google Scholar] [CrossRef]
  109. Watanabe, M.; Houten, S.M.; Wang, L.; Moschetta, A.; Mangelsdorf, D.J.; Heyman, R.A.; Moore, D.D.; Auwerx, J. Bile acids lower triglyceride levels via a pathway involving FXR, SHP, and SREBP-1c. J. Clin. Investig. 2004, 113, 1408–1418. [Google Scholar] [CrossRef]
  110. Wiegman, C.H.; Bandsma, R.H.; Ouwens, M.; van der Sluijs, F.H.; Havinga, R.; Boer, T.; Reijngoud, D.-J.; Romijn, J.A.; Kuipers, F. Hepatic VLDL production in ob/ob mice is not stimulated by massive de novo lipogenesis but is less sensitive to the suppressive effects of insulin. Diabetes 2003, 52, 1081–1089. [Google Scholar] [CrossRef]
  111. Katsuma, S.; Hirasawa, A.; Tsujimoto, G. Bile acids promote glucagon-like peptide-1 secretion through TGR5 in a murine enteroendocrine cell line STC-1. Biochem. Biophys. Res. Commun. 2005, 329, 386–390. [Google Scholar] [CrossRef]
  112. Bernstein, H.; Bernstein, C.; Payne, C.; Dvorakova, K.; Garewal, H. Bile acids as carcinogens in human gastrointestinal cancers. Mutat. Res./Rev. Mutat. Res. 2005, 589, 47–65. [Google Scholar] [CrossRef]
  113. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, inflammation, and cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef]
  114. Qiao, D.; Gaitonde, S.V.; Qi, W.; Martinez, J.D. Deoxycholic acid suppresses p53 by stimulating proteasome-mediated p53 protein degradation. Carcinogenesis 2001, 22, 957–964. [Google Scholar] [CrossRef]
  115. Payne, C.M.; Bernstein, C.; Dvorak, K.; Bernstein, H. Hydrophobic bile acids, genomic instability, Darwinian selection, and colon carcinogenesis. Clin. Exp. Gastroenterol. 2008, 1, 19–47. [Google Scholar] [CrossRef]
  116. Washo-Stultz, D.; Crowley-Weber, C.L.; Dvorakova, K.; Bernstein, C.; Bernstein, H.; Kunke, K.; Waltmire, C.N.; Garewal, H.; Payne, C.M. Role of mitochondrial complexes I and II, reactive oxygen species and arachidonic acid metabolism in deoxycholate-induced apoptosis. Cancer Lett. 2002, 177, 129–144. [Google Scholar] [CrossRef]
  117. Tang, W.W.; Hazen, S.L. The contributory role of gut microbiota in cardiovascular disease. J. Clin. Investig. 2014, 124, 4204–4211. [Google Scholar] [CrossRef]
  118. Tang, W.W.; 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. [Google Scholar] [CrossRef]
  119. Wang, Z.; Tang, W.W.; Buffa, J.A.; Fu, X.; Britt, E.B.; Koeth, R.A.; Levison, B.S.; Fan, Y.; Wu, Y.; Hazen, S.L. Prognostic value of choline and betaine depends on intestinal microbiota-generated metabolite trimethylamine-N-oxide. Eur. Heart J. 2014, 35, 904–910. [Google Scholar] [CrossRef]
  120. Craciun, S.; Balskus, E.P. Microbial conversion of choline to trimethylamine requires a glycyl radical enzyme. Proc. Natl. Acad. Sci. USA 2012, 109, 21307–21312. [Google Scholar] [CrossRef]
  121. Koeth, R.A.; Levison, B.S.; Culley, M.K.; Buffa, J.A.; Wang, Z.; Gregory, J.C.; Org, E.; Wu, Y.; Li, L.; Smith, J.D. γ-Butyrobetaine is a proatherogenic intermediate in gut microbial metabolism of L-carnitine to TMAO. Cell Metab. 2014, 20, 799–812. [Google Scholar] [CrossRef]
  122. Wang, Z.; Klipfell, E.; Bennett, B.J.; Koeth, R.; Levison, B.S.; DuGar, B.; Feldstein, A.E.; Britt, E.B.; Fu, X.; Chung, Y.-M. Gut flora metabolism of phosphatidylcholine promotes cardiovascular disease. Nature 2011, 472, 57–63. [Google Scholar] [CrossRef]
  123. Katz, D.L.; Gnanaraj, J.; Treu, J.A.; Ma, Y.; Kavak, Y.; Njike, V.Y. Effects of egg ingestion on endothelial function in adults with coronary artery disease: A randomized, controlled, crossover trial. Am. Heart J. 2015, 169, 162–169. [Google Scholar] [CrossRef]
  124. Ufnal, M.; Zadlo, A.; Ostaszewski, R. TMAO: A small molecule of great expectations. Nutrition 2015, 31, 1317–1323. [Google Scholar] [CrossRef]
  125. Wang, Z.; Levison, B.S.; Hazen, J.E.; Donahue, L.; Li, X.-M.; Hazen, S.L. Measurement of trimethylamine-N-oxide by stable isotope dilution liquid chromatography tandem mass spectrometry. Anal. Biochem. 2014, 455, 35–40. [Google Scholar] [CrossRef]
  126. Zeisel, S.H.; Warrier, M. Trimethylamine N-oxide, the microbiome, and heart and kidney disease. Annu. Rev. Nutr. 2017, 37, 157–181. [Google Scholar] [CrossRef]
  127. Fennema, D.; Phillips, I.R.; Shephard, E.A. Trimethylamine and trimethylamine N-oxide, a flavin-containing monooxygenase 3 (FMO3)-mediated host-microbiome metabolic axis implicated in health and disease. Drug Metab. Dispos. 2016, 44, 1839–1850. [Google Scholar] [CrossRef]
  128. Gao, C.; Catucci, G.; Castrignanò, S.; Gilardi, G.; Sadeghi, S.J. Inactivation mechanism of N61S mutant of human FMO3 towards trimethylamine. Sci. Rep. 2017, 7, 14668. [Google Scholar] [CrossRef]
  129. Organ, C.L.; Li, Z.; Sharp, T.E., III; Polhemus, D.J.; Gupta, N.; Goodchild, T.T.; Tang, W.W.; Hazen, S.L.; Lefer, D.J. Nonlethal Inhibition of Gut Microbial Trimethylamine N-oxide Production Improves Cardiac Function and Remodeling in a Murine Model of Heart Failure. J. Am. Heart Assoc. 2020, 9, e016223. [Google Scholar] [CrossRef]
  130. Zhang, W.; Miikeda, A.; Zuckerman, J.; Jia, X.; Charugundla, S.; Zhou, Z.; Kaczor-Urbanowicz, K.E.; Magyar, C.; Guo, F.; Wang, Z. Inhibition of microbiota-dependent TMAO production attenuates chronic kidney disease in mice. Sci. Rep. 2021, 11, 518. [Google Scholar] [CrossRef]
  131. Gupta, N.; Buffa, J.A.; Roberts, A.B.; Sangwan, N.; Skye, S.M.; Li, L.; Ho, K.J.; Varga, J.; DiDonato, J.A.; Tang, W.W. Targeted inhibition of gut microbial trimethylamine N-oxide production reduces renal tubulointerstitial fibrosis and functional impairment in a murine model of chronic kidney disease. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 1239–1255. [Google Scholar] [CrossRef]
  132. Sánchez-Alcoholado, L.; Ordóñez, R.; Otero, A.; Plaza-Andrade, I.; Laborda-Illanes, A.; Medina, J.A.; Ramos-Molina, B.; Gómez-Millán, J.; Queipo-Ortuño, M.I. Gut microbiota-mediated inflammation and gut permeability in patients with obesity and colorectal cancer. Int. J. Mol. Sci. 2020, 21, 6782. [Google Scholar] [CrossRef]
  133. Zhu, W.; Gregory, J.C.; Org, E.; Buffa, J.A.; Gupta, N.; Wang, Z.; Li, L.; Fu, X.; Wu, Y.; Mehrabian, M. Gut microbial metabolite TMAO enhances platelet hyperreactivity and thrombosis risk. Cell 2016, 165, 111–124. [Google Scholar] [CrossRef]
  134. Ma, G.; Pan, B.; Chen, Y.; Guo, C.; Zhao, M.; Zheng, L.; Chen, B. Trimethylamine N-oxide in atherogenesis: Impairing endothelial self-repair capacity and enhancing monocyte adhesion. Biosci. Rep. 2017, 37, BSR20160244. [Google Scholar] [CrossRef]
  135. Mohammadi, A.; Najar, A.G.; Yaghoobi, M.M.; Jahani, Y.; Vahabzadeh, Z. Trimethylamine-N-oxide treatment induces changes in the ATP-binding cassette transporter A1 and scavenger receptor A1 in murine macrophage J774A. 1 cells. Inflammation 2016, 39, 393–404. [Google Scholar] [CrossRef]
  136. Mondal, J.A. Effect of trimethylamine N-oxide on interfacial electrostatics at phospholipid monolayer–water interfaces and its relevance to cardiovascular disease. J. Phys. Chem. Lett. 2016, 7, 1704–1708. [Google Scholar] [CrossRef]
  137. 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. Intestinal microbiota metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat. Med. 2013, 19, 576–585. [Google Scholar] [CrossRef]
  138. Ding, L.; Chang, M.; Guo, Y.; Zhang, L.; Xue, C.; Yanagita, T.; Zhang, T.; Wang, Y. Trimethylamine-N-oxide (TMAO)-induced atherosclerosis is associated with bile acid metabolism. Lipids Health Dis. 2018, 17, 286. [Google Scholar] [CrossRef]
  139. Danaceau, J.P.; Anderson, G.M.; McMahon, W.M.; Crouch, D.J. A liquid chromatographic-tandem mass spectrometric method for the analysis of serotonin and related indoles in human whole blood. J. Anal. Toxicol. 2003, 27, 440–444. [Google Scholar] [CrossRef]
  140. Negatu, D.A.; Gengenbacher, M.; Dartois, V.; Dick, T. Indole propionic acid, an unusual antibiotic produced by the gut microbiota, with anti-inflammatory and antioxidant properties. Front. Microbiol. 2020, 11, 575586. [Google Scholar] [CrossRef]
  141. Tuomainen, M.; Lindström, J.; Lehtonen, M.; Auriola, S.; Pihlajamäki, 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. [Google Scholar] [CrossRef]
  142. Pappolla, M.A.; Perry, G.; Fang, X.; Zagorski, M.; Sambamurti, K.; Poeggeler, B. Indoles as essential mediators in the gut-brain axis. Their Role Alzheimer’s disease. Neurobiol. Dis. 2021, 156, 105403. [Google Scholar]
  143. Parthasarathy, A.; Cross, P.J.; Dobson, R.C.; 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. [Google Scholar] [CrossRef]
  144. Melhem, N.J.; Taleb, S. Tryptophan: From diet to cardiovascular diseases. Int. J. Mol. Sci. 2021, 22, 9904. [Google Scholar] [CrossRef]
  145. Agus, A.; Planchais, J.; Sokol, H. Gut microbiota regulation of tryptophan metabolism in health and disease. Cell Host Microbe 2018, 23, 716–724. [Google Scholar] [CrossRef]
  146. Roager, H.M.; Licht, T.R. Microbial tryptophan catabolites in health and disease. Nat. Commun. 2018, 9, 3294. [Google Scholar] [CrossRef] [PubMed]
  147. 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. 2020, 69, 1487–1495. [Google Scholar] [CrossRef]
  148. 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. [Google Scholar] [CrossRef]
  149. 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. Microbiota-derived indole metabolites promote human and murine intestinal homeostasis through regulation of interleukin-10 receptor. Am. J. Pathol. 2018, 188, 1183–1194. [Google Scholar] [CrossRef]
  150. 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. [Google Scholar] [CrossRef]
  151. 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. Neuroendocr. Lett 2010, 31, 363–369. [Google Scholar]
  152. Karbownik, M.; Stasiak, M.; Zygmunt, A.; Zasada, K.; Lewiński, A. Protective effects of melatonin and indole-3-propionic acid against lipid peroxidation, caused by potassium bromate in the rat kidney. Cell Biochem. Funct. Cell. Biochem. Its Modul. By Act. Agents Or Dis. 2006, 24, 483–489. [Google Scholar] [CrossRef]
  153. 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. [Google Scholar] [CrossRef]
  154. 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. [Google Scholar] [CrossRef]
  155. 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. Gut microbiota-derived indole 3-propionic acid protects against radiation toxicity via retaining acyl-CoA-binding protein. Microbiome 2020, 8, 69. [Google Scholar] [CrossRef] [PubMed]
  156. Pappolla, M.; Matsubara, E.; Vidal, R.; Pacheco-Quinto, J.; Poeggeler, B.; Zagorski, M.; Sambamurti, K. Melatonin treatment enhances Aβ lymphatic clearance in a transgenic mouse model of amyloidosis. Curr. Alzheimer Res. 2018, 15, 637–642. [Google Scholar] [CrossRef] [PubMed]
  157. 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. [Google Scholar] [CrossRef] [PubMed]
  158. 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, 1–25. [Google Scholar]
  159. 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. [Google Scholar] [CrossRef]
  160. Rabus, R.; Venceslau, S.S.; Woehlbrand, L.; Voordouw, G.; Wall, J.D.; Pereira, I.A. A post-genomic view of the ecophysiology, catabolism and biotechnological relevance of sulphate-reducing prokaryotes. Adv. Microb. Physiol. 2015, 66, 55–321. [Google Scholar]
  161. Rey, F.E.; Gonzalez, M.D.; Cheng, J.; Wu, M.; Ahern, P.P.; Gordon, J.I. Metabolic niche of a prominent sulfate-reducing human gut bacterium. Proc. Natl. Acad. Sci. USA 2013, 110, 13582–13587. [Google Scholar] [CrossRef]
  162. Gibson, G.; Macfarlane, G.; Cummings, J. Occurrence of sulphate-reducing bacteria in human faeces and the relationship of dissimilatory sulphate reduction to methanogenesis in the large gut. J. Appl. Bacteriol. 1988, 65, 103–111. [Google Scholar] [CrossRef]
  163. Awano, N.; Wada, M.; Mori, H.; Nakamori, S.; Takagi, H. Identification and functional analysis of Escherichia coli cysteine desulfhydrases. Appl. Environ. Microbiol. 2005, 71, 4149–4152. [Google Scholar] [CrossRef]
  164. Blachier, F.; Davila, A.-M.; Mimoun, S.; Benetti, P.-H.; Atanasiu, C.; Andriamihaja, M.; Benamouzig, R.; Bouillaud, F.; Tomé, D. Luminal sulfide and large intestine mucosa: Friend or foe? Amino Acids 2010, 39, 335–347. [Google Scholar] [CrossRef]
  165. Distrutti, E.; Sediari, L.; Mencarelli, A.; Renga, B.; Orlandi, S.; Russo, G.; Caliendo, G.; Santagada, V.; Cirino, G.; Wallace, J.L. 5-Amino-2-hydroxybenzoic acid 4-(5-thioxo-5H-[1, 2] dithiol-3yl)-phenyl ester (ATB-429), a hydrogen sulfide-releasing derivative of mesalamine, exerts antinociceptive effects in a model of postinflammatory hypersensitivity. J. Pharmacol. Exp. Ther. 2006, 319, 447–458. [Google Scholar] [CrossRef] [PubMed]
  166. Martin, G.R.; McKnight, G.W.; Dicay, M.S.; Coffin, C.S.; Ferraz, J.G.; Wallace, J.L. Hydrogen sulphide synthesis in the rat and mouse gastrointestinal tract. Dig. Liver Dis. 2010, 42, 103–109. [Google Scholar] [CrossRef] [PubMed]
  167. Wallace, J.L.; Vong, L.; McKnight, W.; Dicay, M.; Martin, G.R. Endogenous and exogenous hydrogen sulfide promotes resolution of colitis in rats. Gastroenterology 2009, 137, 569–578.e561. [Google Scholar] [CrossRef]
  168. Fiorucci, S.; Antonelli, E.; Distrutti, E.; Rizzo, G.; Mencarelli, A.; Orlandi, S.; Zanardo, R.; Renga, B.; Di Sante, M.; Morelli, A. Inhibition of hydrogen sulfide generation contributes to gastric injury caused by anti-inflammatory nonsteroidal drugs. Gastroenterology 2005, 129, 1210–1224. [Google Scholar] [CrossRef]
  169. Shatalin, K.; Shatalina, E.; Mironov, A.; Nudler, E. H2S: A universal defense against antibiotics in bacteria. Science 2011, 334, 986–990. [Google Scholar] [CrossRef] [PubMed]
  170. Zhao, H.; Yan, R.; Zhou, X.; Ji, F.; Zhang, B. Hydrogen sulfide improves colonic barrier integrity in DSS-induced inflammation in Caco-2 cells and mice. Int. Immunopharmacol. 2016, 39, 121–127. [Google Scholar] [CrossRef]
  171. Figliuolo, V.R.; Dos Santos, L.M.; Abalo, A.; Nanini, H.; Santos, A.; Brittes, N.M.; Bernardazzi, C.; de Souza, H.S.P.; Vieira, L.Q.; Coutinho-Silva, R. Sulfate-reducing bacteria stimulate gut immune responses and contribute to inflammation in experimental colitis. Life Sci. 2017, 189, 29–38. [Google Scholar] [CrossRef]
  172. Lohninger, L.; Tomasova, L.; Praschberger, M.; Hintersteininger, M.; Erker, T.; Gmeiner, B.M.; Laggner, H. Hydrogen sulphide induces HIF-1α and Nrf2 in THP-1 macrophages. Biochimie 2015, 112, 187–195. [Google Scholar] [CrossRef]
  173. Yang, R.; Qu, C.; Zhou, Y.; Konkel, J.E.; Shi, S.; Liu, Y.; Chen, C.; Liu, S.; Liu, D.; Chen, Y. Hydrogen sulfide promotes Tet1-and Tet2-mediated Foxp3 demethylation to drive regulatory T cell differentiation and maintain immune homeostasis. Immunity 2015, 43, 251–263. [Google Scholar] [CrossRef]
  174. Zhao, H.; Pan, P.; Yang, Y.; Ge, H.; Chen, W.; Qu, J.; Shi, J.; Cui, G.; Liu, X.; Feng, H. Endogenous hydrogen sulphide attenuates NLRP3 inflammasome-mediated neuroinflammation by suppressing the P2X7 receptor after intracerebral haemorrhage in rats. J. Neuroinflamm. 2017, 14, 163. [Google Scholar] [CrossRef]
  175. Nemet, I.; Saha, P.P.; Gupta, N.; Zhu, W.; Romano, K.A.; Skye, S.M.; Cajka, T.; Mohan, M.L.; Li, L.; Wu, Y. A cardiovascular disease-linked gut microbial metabolite acts via adrenergic receptors. Cell 2020, 180, 862–877.e822. [Google Scholar] [CrossRef] [PubMed]
  176. Wen, Y.; Sun, Z.; Xie, S.; Hu, Z.; Lan, Q.; Sun, Y.; Yuan, L.; Zhai, C. Intestinal flora derived metabolites affect the occurrence and development of cardiovascular disease. J. Multidiscip. Healthc. 2022, 15, 2591–2603. [Google Scholar] [CrossRef] [PubMed]
  177. Chen, W.; Zhang, S.; Wu, J.; Ye, T.; Wang, S.; Wang, P.; Xing, D. Butyrate-producing bacteria and the gut-heart axis in atherosclerosis. Clin. Chim. Acta 2020, 507, 236–241. [Google Scholar] [CrossRef]
  178. Huynh, K. Novel gut microbiota-derived metabolite promotes platelet thrombosis via adrenergic receptor signalling. Nat. Rev. Cardiol. 2020, 17, 265. [Google Scholar] [CrossRef] [PubMed]
  179. Yu, F.; Li, X.; Feng, X.; Wei, M.; Luo, Y.; Zhao, T.; Xiao, B.; Xia, J. Phenylacetylglutamine, a novel biomarker in acute ischemic stroke. Front. Cardiovasc. Med. 2021, 8, 798765. [Google Scholar] [CrossRef]
  180. Chiorescu, R.M.; Mocan, M.; Inceu, A.I.; Buda, A.P.; Blendea, D.; Vlaicu, S.I. Vulnerable atherosclerotic plaque: Is there a molecular signature? Int. J. Mol. Sci. 2022, 23, 13638. [Google Scholar] [CrossRef]
  181. Lupu, V.V.; Adam Raileanu, A.; Mihai, C.M.; Morariu, I.D.; Lupu, A.; Starcea, I.M.; Frasinariu, O.E.; Mocanu, A.; Dragan, F.; Fotea, S. The implication of the gut microbiome in heart failure. Cells 2023, 12, 1158. [Google Scholar] [CrossRef]
  182. Fang, C.; Zuo, K.; Fu, Y.; Li, J.; Wang, H.; Xu, L.; Yang, X. Dysbiosis of gut microbiota and metabolite phenylacetylglutamine in coronary artery disease patients with stent stenosis. Front. Cardiovasc. Med. 2022, 9, 832092. [Google Scholar] [CrossRef] [PubMed]
  183. Polsinelli, V.B.; Marteau, L.; Shah, S.J. The role of splanchnic congestion and the intestinal microenvironment in the pathogenesis of advanced heart failure. Curr. Opin. Support. Palliat. Care 2019, 13, 24. [Google Scholar] [CrossRef]
  184. Wang, Z.; Cai, Z.; Ferrari, M.W.; Liu, Y.; Li, C.; Zhang, T.; Lyu, G. The correlation between gut microbiota and serum metabolomic in elderly patients with chronic heart failure. Mediat. Inflamm. 2021, 2021, 5587428. [Google Scholar] [CrossRef]
  185. Cui, X.; Ye, L.; Li, J.; Jin, L.; Wang, W.; Li, S.; Bao, M.; Wu, S.; Li, L.; Geng, B. Metagenomic and metabolomic analyses unveil dysbiosis of gut microbiota in chronic heart failure patients. Sci. Rep. 2018, 8, 635. [Google Scholar] [CrossRef] [PubMed]
  186. Jie, Z.; Xia, H.; Zhong, S.-L.; Feng, Q.; Li, S.; Liang, S.; Zhong, H.; Liu, Z.; Gao, Y.; Zhao, H. The gut microbiome in atherosclerotic cardiovascular disease. Nat. Commun. 2017, 8, 845. [Google Scholar] [CrossRef] [PubMed]
  187. Navab-Moghadam, F.; Sedighi, M.; Khamseh, M.E.; Alaei-Shahmiri, F.; Talebi, M.; Razavi, S.; Amirmozafari, N. The association of type II diabetes with gut microbiota composition. Microb. Pathog. 2017, 110, 630–636. [Google Scholar] [CrossRef] [PubMed]
  188. Karlsson, F.H.; Tremaroli, V.; Nookaew, I.; Bergström, G.; Behre, C.J.; Fagerberg, B.; Nielsen, J.; Bäckhed, F. Gut metagenome in European women with normal, impaired and diabetic glucose control. Nature 2013, 498, 99–103. [Google Scholar] [CrossRef] [PubMed]
  189. Mamic, P.; Chaikijurajai, T.; Tang, W.W. Gut microbiome-A potential mediator of pathogenesis in heart failure and its comorbidities: State-of-the-art review. J. Mol. Cell. Cardiol. 2021, 152, 105–117. [Google Scholar] [CrossRef]
  190. Cui, L.; Zhao, T.; Hu, H.; Zhang, W.; Hua, X. Association study of gut flora in coronary heart disease through high-throughput sequencing. BioMed Res. Int. 2017, 2017, 3796359. [Google Scholar] [CrossRef]
  191. Espín, J.C.; González-Sarrías, A.; Tomás-Barberán, F.A. The gut microbiota: A key factor in the therapeutic effects of (poly) phenols. Biochem. Pharmacol. 2017, 139, 82–93. [Google Scholar] [CrossRef]
  192. Gallo, A.; Macerola, N.; Favuzzi, A.M.; Nicolazzi, M.A.; Gasbarrini, A.; Montalto, M. The gut in heart failure: Current knowledge and novel frontiers. Med. Princ. Pract. 2022, 31, 203–214. [Google Scholar] [CrossRef]
  193. Bartolomaeus, H.; Balogh, A.; Yakoub, M.; Homann, S.; Markó, L.; Höges, S.; Tsvetkov, D.; Krannich, A.; Wundersitz, S.; Avery, E.G. Short-chain fatty acid propionate protects from hypertensive cardiovascular damage. Circulation 2019, 139, 1407–1421. [Google Scholar] [CrossRef]
  194. Gerdes, V.; Gueimonde, M.; Pajunen, L.; Nieuwdorp, M.; Laitinen, K. How strong is the evidence that gut microbiota composition can be influenced by lifestyle interventions in a cardio-protective way? Atherosclerosis 2020, 311, 124–142. [Google Scholar] [CrossRef]
  195. Brandsma, E.; Kloosterhuis, N.J.; Koster, M.; Dekker, D.C.; Gijbels, M.J.; Van Der Velden, S.; Ríos-Morales, M.; Van Faassen, M.J.; Loreti, M.G.; De Bruin, A. A proinflammatory gut microbiota increases systemic inflammation and accelerates atherosclerosis. Circ. Res. 2019, 124, 94–100. [Google Scholar] [CrossRef]
  196. Emoto, T.; Yamashita, T.; Kobayashi, T.; Sasaki, N.; Hirota, Y.; Hayashi, T.; So, A.; Kasahara, K.; Yodoi, K.; Matsumoto, T. Characterization of gut microbiota profiles in coronary artery disease patients using data mining analysis of terminal restriction fragment length polymorphism: Gut microbiota could be a diagnostic marker of coronary artery disease. Heart Vessel. 2017, 32, 39–46. [Google Scholar] [CrossRef] [PubMed]
  197. Emoto, T.; Yamashita, T.; Sasaki, N.; Hirota, Y.; Hayashi, T.; So, A.; Kasahara, K.; Yodoi, K.; Matsumoto, T.; Mizoguchi, T. Analysis of gut microbiota in coronary artery disease patients: A possible link between gut microbiota and coronary artery disease. J. Atheroscler. Thromb. 2016, 23, 908–921. [Google Scholar] [CrossRef] [PubMed]
  198. Li, L.; Hua, Y.; Ren, J. Short-chain fatty acid propionate alleviates Akt2 knockout-induced myocardial contractile dysfunction. J. Diabetes Res. 2012, 2012, 851717. [Google Scholar] [CrossRef]
  199. Gelis, L.; Jovancevic, N.; Veitinger, S.; Mandal, B.; Arndt, H.-D.; Neuhaus, E.M.; Hatt, H. Functional characterization of the odorant receptor 51E2 in human melanocytes. J. Biol. Chem. 2016, 291, 17772–17786. [Google Scholar] [CrossRef]
  200. Dorsam, R.T.; Gutkind, J.S. G-protein-coupled receptors and cancer. Nat. Rev. Cancer 2007, 7, 79–94. [Google Scholar] [CrossRef] [PubMed]
  201. Fleischer, J.; Bumbalo, R.; Bautze, V.; Strotmann, J.; Breer, H. Expression of odorant receptor Olfr78 in enteroendocrine cells of the colon. Cell Tissue Res. 2015, 361, 697–710. [Google Scholar] [CrossRef]
  202. Pluznick, J.L. Renal and cardiovascular sensory receptors and blood pressure regulation. Am. J. Physiol.-Ren. Physiol. 2013, 305, F439–F444. [Google Scholar] [CrossRef]
  203. Zhu, Y.; Jameson, E.; Crosatti, M.; Schäfer, H.; Rajakumar, K.; Bugg, T.D.; Chen, Y. Carnitine metabolism to trimethylamine by an unusual Rieske-type oxygenase from human microbiota. Proc. Natl. Acad. Sci. USA 2014, 111, 4268–4273. [Google Scholar] [CrossRef]
  204. Romano, K.A.; Vivas, E.I.; Amador-Noguez, D.; Rey, F.E. Intestinal microbiota composition modulates choline bioavailability from diet and accumulation of the proatherogenic metabolite trimethylamine-N-oxide. mBio 2015, 6, e02481-14. [Google Scholar] [CrossRef]
  205. Dong, Z.; Liang, Z.; Guo, M.; Hu, S.; Shen, Z.; Hai, X. The association between plasma levels of trimethylamine N-oxide and the risk of coronary heart disease in Chinese patients with or without type 2 diabetes mellitus. Dis. Markers 2018, 2018, 1578320. [Google Scholar] [CrossRef] [PubMed]
  206. Chan, M.M.; Yang, X.; Wang, H.; Saaoud, F.; Sun, Y.; Fong, D. The microbial metabolite trimethylamine N-oxide links vascular dysfunctions and the autoimmune disease rheumatoid arthritis. Nutrients 2019, 11, 1821. [Google Scholar] [CrossRef] [PubMed]
  207. Cho, C.E.; Taesuwan, S.; Malysheva, O.V.; Bender, E.; Tulchinsky, N.F.; Yan, J.; Sutter, J.L.; Caudill, M.A. Trimethylamine-N-oxide (TMAO) response to animal source foods varies among healthy young men and is influenced by their gut microbiota composition: A randomized controlled trial. Mol. Nutr. Food Res. 2017, 61, 1600324. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, D.D.; Nguyen, L.H.; Li, Y.; Yan, Y.; Ma, W.; Rinott, E.; Ivey, K.L.; Shai, I.; Willett, W.C.; Hu, F.B. The gut microbiome modulates the protective association between a Mediterranean diet and cardiometabolic disease risk. Nat. Med. 2021, 27, 333–343. [Google Scholar] [CrossRef] [PubMed]
  209. Boini, K.M.; Hussain, T.; Li, P.-L.; Koka, S.S. Trimethylamine-N-oxide instigates NLRP3 inflammasome activation and endothelial dysfunction. Cell. Physiol. Biochem. 2018, 44, 152–162. [Google Scholar] [CrossRef]
  210. Singh, G.B.; Zhang, Y.; Boini, K.M.; Koka, S. High mobility group box 1 mediates TMAO-induced endothelial dysfunction. Int. J. Mol. Sci. 2019, 20, 3570. [Google Scholar] [CrossRef]
  211. Ke, Y.; Li, D.; Zhao, M.; Liu, C.; Liu, J.; Zeng, A.; Shi, X.; Cheng, S.; Pan, B.; Zheng, L. Gut flora-dependent metabolite Trimethylamine-N-oxide accelerates endothelial cell senescence and vascular aging through oxidative stress. Free Radic. Biol. Med. 2018, 116, 88–100. [Google Scholar] [CrossRef]
  212. González-Loyola, A.; Petrova, T.V. Development and aging of the lymphatic vascular system. Adv. Drug Deliv. Rev. 2021, 169, 63–78. [Google Scholar] [CrossRef]
  213. Brunt, V.E.; LaRocca, T.J.; Bazzoni, A.E.; Sapinsley, Z.J.; Miyamoto-Ditmon, J.; Gioscia-Ryan, R.A.; Neilson, A.P.; Link, C.D.; Seals, D.R. The gut microbiome–derived metabolite trimethylamine N-oxide modulates neuroinflammation and cognitive function with aging. GeroScience 2021, 43, 377–394. [Google Scholar] [CrossRef]
  214. Chen, M.l.; Zhu, X.h.; Ran, L.; Lang, H.d.; Yi, L.; Mi, M.t. Trimethylamine-N-Oxide induces vascular inflammation by activating the NLRP3 inflammasome through the SIRT3-SOD2-mtROS signaling pathway. J. Am. Heart Assoc. 2017, 6, e006347. [Google Scholar] [CrossRef]
  215. Boutagy, N.E.; Neilson, A.P.; Osterberg, K.L.; Smithson, A.T.; Englund, T.R.; Davy, B.M.; Hulver, M.W.; Davy, K.P. Probiotic supplementation and trimethylamine-N-oxide production following a high-fat diet. Obesity 2015, 23, 2357–2363. [Google Scholar] [CrossRef] [PubMed]
  216. Seldin, M.M.; Meng, Y.; Qi, H.; Zhu, W.; Wang, Z.; Hazen, S.L.; Lusis, A.J.; Shih, D.M. Trimethylamine N-oxide promotes vascular inflammation through signaling of mitogen-activated protein kinase and nuclear factor-κB. J. Am. Heart Assoc. 2016, 5, e002767. [Google Scholar] [CrossRef] [PubMed]
  217. Yang, S.; Li, X.; Yang, F.; Zhao, R.; Pan, X.; Liang, J.; Tian, L.; Li, X.; Liu, L.; Xing, Y. Gut microbiota-dependent marker TMAO in promoting cardiovascular disease: Inflammation mechanism, clinical prognostic, and potential as a therapeutic target. Front. Pharmacol. 2019, 10, 1360. [Google Scholar] [CrossRef]
  218. Liu, Y.; Dai, M. Trimethylamine N-oxide generated by the gut microbiota is associated with vascular inflammation: New insights into atherosclerosis. Mediat. Inflamm. 2020, 2020, 4634172. [Google Scholar] [CrossRef]
  219. Durpès, M.-C.; Morin, C.; Paquin-Veillet, J.; Beland, R.; Paré, M.; Guimond, M.-O.; Rekhter, M.; King, G.L.; Geraldes, P. PKC-β activation inhibits IL-18-binding protein causing endothelial dysfunction and diabetic atherosclerosis. Cardiovasc. Res. 2015, 106, 303–313. [Google Scholar] [CrossRef]
  220. Liu, H.; Jia, K.; Ren, Z.; Sun, J.; Pan, L.-L. PRMT5 critically mediates TMAO-induced inflammatory response in vascular smooth muscle cells. Cell Death Dis. 2022, 13, 299. [Google Scholar] [CrossRef]
  221. Zheng, Y.; Li, Y.; Rimm, E.B.; Hu, F.B.; Albert, C.M.; Rexrode, K.M.; Manson, J.E.; Qi, L. Dietary phosphatidylcholine and risk of all-cause and cardiovascular-specific mortality among US women and men. Am. J. Clin. Nutr. 2016, 104, 173–180. [Google Scholar] [CrossRef]
  222. Zhang, J.-M.; An, J. Cytokines, inflammation and pain. Int. Anesthesiol. Clin. 2007, 45, 27. [Google Scholar] [CrossRef] [PubMed]
  223. Gui, Y.; Zheng, H.; Cao, R.Y. Foam cells in atherosclerosis: Novel insights into its origins, consequences, and molecular mechanisms. Front. Cardiovasc. Med. 2022, 9, 845942. [Google Scholar] [CrossRef]
  224. Bentzon, J.F.; Otsuka, F.; Virmani, R.; Falk, E. Mechanisms of plaque formation and rupture. Circ. Res. 2014, 114, 1852–1866. [Google Scholar] [CrossRef]
  225. Zhou, S.; Xue, J.; Shan, J.; Hong, Y.; Zhu, W.; Nie, Z.; Zhang, Y.; Ji, N.; Luo, X.; Zhang, T. Gut-flora-dependent metabolite trimethylamine-N-oxide promotes atherosclerosis-associated inflammation responses by indirect ROS stimulation and signaling involving AMPK and SIRT1. Nutrients 2022, 14, 3338. [Google Scholar] [CrossRef] [PubMed]
  226. Hu, H.; Shao, W.; Liu, Q.; Liu, N.; Wang, Q.; Xu, J.; Zhang, X.; Weng, Z.; Lu, Q.; Jiao, L. Gut microbiota promotes cholesterol gallstone formation by modulating bile acid composition and biliary cholesterol secretion. Nat. Commun. 2022, 13, 252. [Google Scholar] [CrossRef] [PubMed]
  227. Guzior, D.V.; Okros, M.; Shivel, M.; Armwald, B.; Bridges, C.; Fu, Y.; Martin, C.; Schilmiller, A.L.; Miller, W.M.; Ziegler, K.M. Bile salt hydrolase acyltransferase activity expands bile acid diversity. Nature 2024, 626, 852–858. [Google Scholar] [CrossRef] [PubMed]
  228. Lu, Q.; Chen, J.; Jiang, L.; Geng, T.; Tian, S.; Liao, Y.; Yang, K.; Zheng, Y.; He, M.; Tang, H. Gut microbiota-derived secondary bile acids, bile acids receptor polymorphisms, and risk of cardiovascular disease in individuals with newly diagnosed type 2 diabetes: A cohort study. Am. J. Clin. Nutr. 2023, 119, 324–332. [Google Scholar] [CrossRef] [PubMed]
  229. Rodríguez-Morató, J.; Matthan, N.R. Nutrition and gastrointestinal microbiota, microbial-derived secondary bile acids, and cardiovascular disease. Curr. Atheroscler. Rep. 2020, 22, 47. [Google Scholar] [CrossRef] [PubMed]
  230. Chiang, J.Y.; Ferrell, J.M. Bile acid receptors FXR and TGR5 signaling in fatty liver diseases and therapy. Am. J. Physiol.-Gastrointest. Liver Physiol. 2020, 318, G554–G573. [Google Scholar] [CrossRef] [PubMed]
  231. Jiao, N.; Baker, S.S.; Chapa-Rodriguez, A.; Liu, W.; Nugent, C.A.; Tsompana, M.; Mastrandrea, L.; Buck, M.J.; Baker, R.D.; Genco, R.J. Suppressed hepatic bile acid signalling despite elevated production of primary and secondary bile acids in NAFLD. Gut 2018, 67, 1881–1891. [Google Scholar] [CrossRef] [PubMed]
  232. Calkin, A.C.; Tontonoz, P. Transcriptional integration of metabolism by the nuclear sterol-activated receptors LXR and FXR. Nat. Rev. Mol. Cell Biol. 2012, 13, 213–224. [Google Scholar] [CrossRef]
  233. Li, W.; Shu, S.; Cheng, L.; Hao, X.; Wang, L.; Wu, Y.; Yuan, Z.; Zhou, J. Fasting serum total bile acid level is associated with coronary artery disease, myocardial infarction and severity of coronary lesions. Atherosclerosis 2020, 292, 193–200. [Google Scholar] [CrossRef]
  234. Fiorucci, S.; Zampella, A.; Cirino, G.; Bucci, M.; Distrutti, E. Decoding the vasoregulatory activities of bile acid-activated receptors in systemic and portal circulation: Role of gaseous mediators. Am. J. Physiol.-Heart Circ. Physiol. 2017, 312, H21–H32. [Google Scholar] [CrossRef]
  235. Sauerbruch, T.; Hennenberg, M.; Trebicka, J.; Beuers, U. Bile acids, liver cirrhosis, and extrahepatic vascular dysfunction. Front. Physiol. 2021, 12, 718783. [Google Scholar] [CrossRef] [PubMed]
  236. Jia, B.; Zou, Y.; Han, X.; Bae, J.-W.; Jeon, C.O. Gut microbiome-mediated mechanisms for reducing cholesterol levels: Implications for ameliorating cardiovascular disease. Trends Microbiol. 2023, 31, 76–91. [Google Scholar] [CrossRef] [PubMed]
  237. Zhou, C.; King, N.; Chen, K.Y.; Breslow, J.L. Activation of PXR induces hypercholesterolemia in wild-type and accelerates atherosclerosis in apoE deficient mice [S]. J. Lipid Res. 2009, 50, 2004–2013. [Google Scholar] [CrossRef]
  238. Sui, Y.; Xu, J.; Rios-Pilier, J.; Zhou, C. Deficiency of PXR decreases atherosclerosis in apoE-deficient mice. J. Lipid Res. 2011, 52, 1652–1659. [Google Scholar] [CrossRef]
  239. Park, Y.M. CD36, a scavenger receptor implicated in atherosclerosis. Exp. Mol. Med. 2014, 46, e99. [Google Scholar] [CrossRef] [PubMed]
  240. Magder, S. The meaning of blood pressure. Crit. Care 2018, 22, 257. [Google Scholar] [CrossRef]
  241. Ning, B.; Chen, Y.; Waqar, A.B.; Yan, H.; Shiomi, M.; Zhang, J.; Chen, Y.E.; Wang, Y.; Itabe, H.; Liang, J. Hypertension enhances advanced atherosclerosis and induces cardiac death in Watanabe heritable hyperlipidemic rabbits. Am. J. Pathol. 2018, 188, 2936–2947. [Google Scholar] [CrossRef]
  242. Huc, T.; Nowinski, A.; Drapala, A.; Konopelski, P.; Ufnal, M. Indole and indoxyl sulfate, gut 422 bacteria metabolites of tryptophan, change arterial blood pressure via peripheral and 423 central mechanisms in rats. Pharmacol Res 2018, 130, 424. [Google Scholar] [CrossRef]
  243. Gesper, M.; Nonnast, A.B.; 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. [Google Scholar] [CrossRef]
  244. 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. [Google Scholar] [CrossRef]
  245. Venu, V.K.P.; 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. [Google Scholar] [CrossRef]
  246. 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. Plasma microbiome-modulated indole-and phenyl-derived metabolites associate with advanced atherosclerosis and postoperative outcomes. J. Vasc. Surg. 2018, 68, 1552–1562.e1557. [Google Scholar] [CrossRef] [PubMed]
  247. Xue, H.; Chen, X.; Yu, C.; Deng, Y.; Zhang, Y.; Chen, S.; Chen, X.; Chen, K.; Yang, Y.; Ling, W. Gut microbially produced indole-3-propionic acid inhibits atherosclerosis by promoting reverse cholesterol transport and its deficiency is causally related to atherosclerotic cardiovascular disease. Circ. Res. 2022, 131, 404–420. [Google Scholar] [CrossRef] [PubMed]
  248. Cuchel, M.; Rader, D.J. Macrophage reverse cholesterol transport: Key to the regression of atherosclerosis? Circulation 2006, 113, 2548–2555. [Google Scholar] [CrossRef]
  249. Rosenson, R.S.; Brewer, H.B., Jr.; Davidson, W.S.; Fayad, Z.A.; Fuster, V.; Goldstein, J.; Hellerstein, M.; Jiang, X.-C.; Phillips, M.C.; Rader, D.J. Cholesterol efflux and atheroprotection: Advancing the concept of reverse cholesterol transport. Circulation 2012, 125, 1905–1919. [Google Scholar] [CrossRef] [PubMed]
  250. Westerterp, M.; Fotakis, P.; Ouimet, M.; Bochem, A.E.; Zhang, H.; Molusky, M.M.; Wang, W.; Abramowicz, S.; la Bastide-van Gemert, S.; Wang, N. Cholesterol efflux pathways suppress inflammasome activation, NETosis, and atherogenesis. Circulation 2018, 138, 898–912. [Google Scholar] [CrossRef]
  251. Wang, H.H.; Garruti, G.; Liu, M.; Portincasa, P.; Wang, D.Q. Cholesterol and lipoprotein metabolism and atherosclerosis: Recent advances in reverse cholesterol transport. Ann. Hepatol. 2018, 16, 27–42. [Google Scholar] [CrossRef]
  252. Mani, S.; Untereiner, A.; Wu, L.; Wang, R. Hydrogen sulfide and the pathogenesis of atherosclerosis. Antioxid. Redox Signal. 2014, 20, 805–817. [Google Scholar] [CrossRef]
  253. Zhang, Y.; Tang, Z.-H.; Ren, Z.; Qu, S.-L.; Liu, M.-H.; Liu, L.-S.; Jiang, Z.-S. Hydrogen sulfide, the next potent preventive and therapeutic agent in aging and age-associated diseases. Mol. Cell. Biol. 2013, 33, 1104–1113. [Google Scholar] [CrossRef]
  254. Zhao, Z.-Z.; Wang, Z.; Li, G.-H.; Wang, R.; Tan, J.-M.; Cao, X.; Suo, R.; Jiang, Z.-S. Hydrogen sulfide inhibits macrophage-derived foam cell formation. Exp. Biol. Med. 2011, 236, 169–176. [Google Scholar] [CrossRef]
  255. Yang, G.; Wu, L.; Wang, R. Pro-apoptotic effect of endogenous H2S on human aorta smooth muscle cells. FASEB J. 2006, 20, 553–555. [Google Scholar] [CrossRef] [PubMed]
  256. Zhang, H.; Guo, C.; Zhang, A.; Fan, Y.; Gu, T.; Wu, D.; Sparatore, A.; Wang, C. Effect of S-aspirin, a novel hydrogen-sulfide-releasing aspirin (ACS14), on atherosclerosis in apoE-deficient mice. Eur. J. Pharmacol. 2012, 697, 106–116. [Google Scholar] [CrossRef]
  257. Wang, Y.; Zhao, X.; Jin, H.; Wei, H.; Li, W.; Bu, D.; Tang, X.; Ren, Y.; Tang, C.; Du, J. Role of hydrogen sulfide in the development of atherosclerotic lesions in apolipoprotein E knockout mice. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 173–179. [Google Scholar] [CrossRef]
  258. Eme, L.; Spang, A.; Lombard, J.; Stairs, C.W.; Ettema, T.J. Archaea and the origin of eukaryotes. Nat. Rev. Microbiol. 2017, 15, 711–723. [Google Scholar] [CrossRef]
  259. Gheibi, S.; Jeddi, S.; Kashfi, K.; Ghasemi, A. Regulation of vascular tone homeostasis by NO and H2S: Implications in hypertension. Biochem. Pharmacol. 2018, 149, 42–59. [Google Scholar] [CrossRef]
  260. Kiss, L.; Deitch, E.A.; Szabó, C. Hydrogen sulfide decreases adenosine triphosphate levels in aortic rings and leads to vasorelaxation via metabolic inhibition. Life Sci. 2008, 83, 589–594. [Google Scholar] [CrossRef]
  261. Wang, J.; Gareri, C.; Rockman, H.A. G-protein–coupled receptors in heart disease. Circ. Res. 2018, 123, 716–735. [Google Scholar] [CrossRef]
  262. Iaccarino, G.; Ciccarelli, M.; Sorriento, D.; Galasso, G.; Campanile, A.; Santulli, G.; Cipolletta, E.; Cerullo, V.; Cimini, V.; Altobelli, G.G. Ischemic neoangiogenesis enhanced by β2-adrenergic receptor overexpression: A novel role for the endothelial adrenergic system. Circ. Res. 2005, 97, 1182–1189. [Google Scholar] [CrossRef]
  263. Jansen, V.L.; Gerdes, V.E.; Middeldorp, S.; van Mens, T.E. Gut microbiota and their metabolites in cardiovascular disease. Best Pract. Res. Clin. Endocrinol. Metab. 2021, 35, 101492. [Google Scholar] [CrossRef] [PubMed]
  264. Romano, K.A.; Nemet, I.; Prasad Saha, P.; Haghikia, A.; Li, X.S.; Mohan, M.L.; Lovano, B.; Castel, L.; Witkowski, M.; Buffa, J.A. Gut microbiota-generated phenylacetylglutamine and heart failure. Circ. Heart Fail. 2023, 16, e009972. [Google Scholar] [CrossRef] [PubMed]
  265. Huang, L. Phenylacetylglutamine causes a pathologic inflammation state and enhances atherosclerosis through the b2-adrenergic receptor cAMP PKA NF-kappaB pathway in diabetes. Eur. Heart J. 2023, 44, ehad655.3289. [Google Scholar] [CrossRef]
  266. Xu, H.; Yang, N.; Wang, B.Y.; Zhou, L.; Xu, L.L.; Chen, Y.; Wang, D.J.; Ge, W.H. Phenylacetyl glutamine (PAGln) enhances cardiomyocyte death after myocardial infarction through β1 adrenergic receptor. Environ. Toxicol. 2024, 39, 1682–1699. [Google Scholar] [CrossRef]
  267. Adithya, K.; Rajeev, R.; Selvin, J.; Seghal Kiran, G. Dietary influence on the dynamics of the human gut microbiome: Prospective implications in interventional therapies. ACS Food Sci. Technol. 2021, 1, 717–736. [Google Scholar] [CrossRef]
  268. Leeming, E.R.; Johnson, A.J.; Spector, T.D.; Le Roy, C.I. Effect of diet on the gut microbiota: Rethinking intervention duration. Nutrients 2019, 11, 2862. [Google Scholar] [CrossRef]
  269. Shetty, S.A.; Hugenholtz, F.; Lahti, L.; Smidt, H.; de Vos, W.M. Intestinal microbiome landscaping: Insight in community assemblage and implications for microbial modulation strategies. FEMS Microbiol. Rev. 2017, 41, 182–199. [Google Scholar] [CrossRef] [PubMed]
  270. Adamberg, K.; Kolk, K.; Jaagura, M.; Vilu, R.; Adamberg, S. The composition and metabolism of faecal microbiota is specifically modulated by different dietary polysaccharides and mucin: An isothermal microcalorimetry study. Benef. Microbes 2018, 9, 21–34. [Google Scholar] [CrossRef]
  271. Gentile, C.L.; Weir, T.L. The gut microbiota at the intersection of diet and human health. Science 2018, 362, 776–780. [Google Scholar] [CrossRef] [PubMed]
  272. Parnell, J.A.; Reimer, R.A. Weight loss during oligofructose supplementation is associated with decreased ghrelin and increased peptide YY in overweight and obese adults. Am. J. Clin. Nutr. 2009, 89, 1751–1759. [Google Scholar] [CrossRef]
  273. Roberfroid, M.; Gibson, G.R.; Hoyles, L.; McCartney, A.L.; Rastall, R.; Rowland, I.; Wolvers, D.; Watzl, B.; Szajewska, H.; Stahl, B. Prebiotic effects: Metabolic and health benefits. Br. J. Nutr. 2010, 104, S1–S63. [Google Scholar] [CrossRef]
  274. Bier, A.; Braun, T.; Khasbab, R.; Di Segni, A.; Grossman, E.; Haberman, Y.; Leibowitz, A. A high salt diet modulates the gut microbiota and short chain fatty acids production in a salt-sensitive hypertension rat model. Nutrients 2018, 10, 1154. [Google Scholar] [CrossRef]
  275. Miranda, P.M.; De Palma, G.; Serkis, V.; Lu, J.; Louis-Auguste, M.P.; McCarville, J.L.; Verdu, E.F.; Collins, S.M.; Bercik, P. High salt diet exacerbates colitis in mice by decreasing Lactobacillus levels and butyrate production. Microbiome 2018, 6, 57. [Google Scholar] [CrossRef] [PubMed]
  276. Barrea, L.; Annunziata, G.; Muscogiuri, G.; Laudisio, D.; Di Somma, C.; Maisto, M.; Tenore, G.C.; Colao, A.; Savastano, S. Trimethylamine N-oxide, Mediterranean diet, and nutrition in healthy, normal-weight adults: Also a matter of sex? Nutrition 2019, 62, 7–17. [Google Scholar] [CrossRef] [PubMed]
  277. Mafra, D.; Borges, N.A.; Lindholm, B.; Shiels, P.G.; Evenepoel, P.; Stenvinkel, P. Food as medicine: Targeting the uraemic phenotype in chronic kidney disease. Nat. Rev. Nephrol. 2021, 17, 153–171. [Google Scholar] [CrossRef]
  278. Goñi, I.; Hernández-Galiot, A. Intake of nutrient and non-nutrient dietary antioxidants. contribution of macromolecular antioxidant polyphenols in an elderly Mediterranean population. Nutrients 2019, 11, 2165. [Google Scholar] [CrossRef]
  279. Colman, R.J.; Rubin, D.T. Fecal microbiota transplantation as therapy for inflammatory bowel disease: A systematic review and meta-analysis. J. Crohn’s Colitis 2014, 8, 1569–1581. [Google Scholar] [CrossRef]
  280. Brandt, L.J.; Aroniadis, O.C.; Mellow, M.; Kanatzar, A.; Kelly, C.; Park, T.; Stollman, N.; Rohlke, F.; Surawicz, C. Long-Term Follow-Up of Colonoscopic Fecal Microbiota Transplant for RecurrentClostridium difficileInfection. Off. J. Am. Coll. Gastroenterol. ACG 2012, 107, 1079–1087. [Google Scholar] [CrossRef]
  281. Van Nood, E.; Vrieze, A.; Nieuwdorp, M.; Fuentes, S.; Zoetendal, E.G.; de Vos, W.M.; Visser, C.E.; Kuijper, E.J.; Bartelsman, J.F.; Tijssen, J.G. Duodenal infusion of donor feces for recurrent Clostridium difficile. N. Engl. J. Med. 2013, 368, 407–415. [Google Scholar] [CrossRef]
  282. Khan, M.Y.; Dirweesh, A.; Khurshid, T.; Siddiqui, W.J. Comparing fecal microbiota transplantation to standard-of-care treatment for recurrent Clostridium difficile infection: A systematic review and meta-analysis. Eur. J. Gastroenterol. Hepatol. 2018, 30, 1309–1317. [Google Scholar] [CrossRef] [PubMed]
  283. Murphy, K.; O’Donovan, A.N.; Caplice, N.M.; Ross, R.P.; Stanton, C. Exploring the gut microbiota and cardiovascular disease. Metabolites 2021, 11, 493. [Google Scholar] [CrossRef]
  284. Vrieze, A.; Van Nood, E.; Holleman, F.; Salojärvi, J.; Kootte, R.S.; Bartelsman, J.F.; Dallinga–Thie, G.M.; Ackermans, M.T.; Serlie, M.J.; Oozeer, R. Transfer of intestinal microbiota from lean donors increases insulin sensitivity in individuals with metabolic syndrome. Gastroenterology 2012, 143, 913–916.e917. [Google Scholar] [CrossRef]
  285. Hu, X.-F.; Zhang, W.-Y.; Wen, Q.; Chen, W.-J.; Wang, Z.-M.; Chen, J.; Zhu, F.; Liu, K.; Cheng, L.-X.; Yang, J. Fecal microbiota transplantation alleviates myocardial damage in myocarditis by restoring the microbiota composition. Pharmacol. Res. 2019, 139, 412–421. [Google Scholar] [CrossRef] [PubMed]
  286. Toral, M.; Robles-Vera, I.; De la Visitacion, N.; Romero, M.; Yang, T.; Sánchez, M.; Gómez-Guzmán, M.; Jiménez, R.; Raizada, M.K.; Duarte, J. Critical role of the interaction gut microbiota–sympathetic nervous system in the regulation of blood pressure. Front. Physiol. 2019, 10, 231. [Google Scholar] [CrossRef] [PubMed]
  287. Smits, L.P.; Kootte, R.S.; Levin, E.; Prodan, A.; Fuentes, S.; Zoetendal, E.G.; Wang, Z.; Levison, B.S.; Cleophas, M.C.; Kemper, E.M. Effect of vegan fecal microbiota transplantation on carnitine-and choline-derived trimethylamine-N-oxide production and vascular inflammation in patients with metabolic syndrome. J. Am. Heart Assoc. 2018, 7, e008342. [Google Scholar] [CrossRef] [PubMed]
  288. Kim, T.T.; Parajuli, N.; Sung, M.M.; Bairwa, S.C.; Levasseur, J.; Soltys, C.-L.M.; Wishart, D.S.; Madsen, K.; Schertzer, J.D.; Dyck, J.R. Fecal transplant from resveratrol-fed donors improves glycaemia and cardiovascular features of the metabolic syndrome in mice. Am. J. Physiol.-Endocrinol. Metab. 2018, 315, E511–E519. [Google Scholar] [CrossRef] [PubMed]
  289. Morgan, C.L.; Mukherjee, J.; Jenkins-Jones, S.; Holden, S.; Currie, C. Association between first-line monotherapy with sulphonylurea versus metformin and risk of all-cause mortality and cardiovascular events: A retrospective, observational study. Diabetes Obes. Metab. 2014, 16, 957–962. [Google Scholar] [CrossRef] [PubMed]
  290. Viollet, B.; Guigas, B.; Garcia, N.S.; Leclerc, J.; Foretz, M.; Andreelli, F. Cellular and molecular mechanisms of metformin: An overview. Clin. Sci. 2012, 122, 253–270. [Google Scholar] [CrossRef]
  291. Montandon, S.A.; Jornayvaz, F.R. Effects of antidiabetic drugs on gut microbiota composition. Genes 2017, 8, 250. [Google Scholar] [CrossRef]
  292. Napolitano, A.; Miller, S.; Nicholls, A.W.; Baker, D.; Van Horn, S.; Thomas, E.; Rajpal, D.; Spivak, A.; Brown, J.R.; Nunez, D.J. Novel gut-based pharmacology of metformin in patients with type 2 diabetes mellitus. PLoS ONE 2014, 9, e100778. [Google Scholar] [CrossRef] [PubMed]
  293. Van Stee, M.F.; de Graaf, A.A.; Groen, A.K. Actions of metformin and statins on lipid and glucose metabolism and possible benefit of combination therapy. Cardiovasc. Diabetol. 2018, 17, 94. [Google Scholar] [CrossRef]
  294. Brønden, A.; Albér, A.; Rohde, U.; Rehfeld, J.F.; Holst, J.J.; Vilsbøll, T.; Knop, F.K. Single-dose metformin enhances bile acid–induced glucagon-like peptide-1 secretion in patients with type 2 diabetes. J. Clin. Endocrinol. Metab. 2017, 102, 4153–4162. [Google Scholar] [CrossRef]
  295. De La Cuesta-Zuluaga, J.; Mueller, N.T.; Corrales-Agudelo, V.; Velásquez-Mejía, E.P.; Carmona, J.A.; Abad, J.M.; Escobar, J.S. Metformin is associated with higher relative abundance of mucin-degrading Akkermansia muciniphila and several short-chain fatty acid–producing microbiota in the gut. Diabetes Care 2017, 40, 54–62. [Google Scholar] [CrossRef]
  296. Forslund, K.; Hildebrand, F.; Nielsen, T.; Falony, G.; Le Chatelier, E.; Sunagawa, S.; Prifti, E.; Vieira-Silva, S.; Gudmundsdottir, V.; Krogh Pedersen, H. Disentangling type 2 diabetes and metformin treatment signatures in the human gut microbiota. Nature 2015, 528, 262–266. [Google Scholar] [CrossRef]
  297. Ning, F.; Tuomilehto, J.; Pyörälä, K.; Onat, A.; Söderberg, S.; Qiao, Q.; Group, D.S. Cardiovascular disease mortality in Europeans in relation to fasting and 2-h plasma glucose levels within a normoglycemic range. Diabetes Care 2010, 33, 2211–2216. [Google Scholar] [CrossRef]
  298. Gu, Y.; Wang, X.; Li, J.; Zhang, Y.; Zhong, H.; Liu, R.; Zhang, D.; Feng, Q.; Xie, X.; Hong, J. Analyses of gut microbiota and plasma bile acids enable stratification of patients for antidiabetic treatment. Nat. Commun. 2017, 8, 1785. [Google Scholar] [CrossRef] [PubMed]
  299. Scheen, A.J. Cardiovascular effects of new oral glucose-lowering agents: DPP-4 and SGLT-2 inhibitors. Circ. Res. 2018, 122, 1439–1459. [Google Scholar] [CrossRef] [PubMed]
  300. Zhang, Q.; Xiao, X.; Li, M.; Yu, M.; Ping, F.; Zheng, J.; Wang, T.; Wang, X. Vildagliptin increases butyrate-producing bacteria in the gut of diabetic rats. PLoS ONE 2017, 12, e0184735. [Google Scholar] [CrossRef] [PubMed]
  301. Green, J.B.; Bethel, M.A.; Armstrong, P.W.; Buse, J.B.; Engel, S.S.; Garg, J.; Josse, R.; Kaufman, K.D.; Koglin, J.; Korn, S. Effect of sitagliptin on cardiovascular outcomes in type 2 diabetes. N. Engl. J. Med. 2015, 373, 232–242. [Google Scholar] [CrossRef]
  302. Duewell, P.; Kono, H.; Rayner, K.J.; Sirois, C.M.; Vladimer, G.; Bauernfeind, F.G.; Abela, G.S.; Franchi, L.; Nunez, G.; Schnurr, M. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature 2010, 464, 1357–1361. [Google Scholar] [CrossRef]
  303. Rajamäki, K.; Lappalainen, J.; Öörni, K.; Välimäki, E.; Matikainen, S.; Kovanen, P.T.; Eklund, K.K. Cholesterol crystals activate the NLRP3 inflammasome in human macrophages: A novel link between cholesterol metabolism and inflammation. PLoS ONE 2010, 5, e11765. [Google Scholar] [CrossRef]
  304. Zhang, X.-N.; Yu, Z.-L.; Chen, J.-Y.; Li, X.-Y.; Wang, Z.-P.; Wu, M.; Liu, L.-T. The crosstalk between NLRP3 inflammasome and gut microbiome in atherosclerosis. Pharmacol. Res. 2022, 181, 106289. [Google Scholar] [CrossRef]
  305. Li, X.; Geng, J.; Zhao, J.; Ni, Q.; Zhao, C.; Zheng, Y.; Chen, X.; Wang, L. Trimethylamine N-oxide exacerbates cardiac fibrosis via activating the NLRP3 inflammasome. Front. Physiol. 2019, 10, 866. [Google Scholar] [CrossRef]
  306. Shao, B.-Z.; Xu, H.-Y.; Zhao, Y.-C.; Zheng, X.-R.; Wang, F.; Zhao, G.-R. NLRP3 inflammasome in atherosclerosis: Putting out the fire of inflammation. Inflammation 2023, 46, 35–46. [Google Scholar] [CrossRef]
  307. Han, L.; Li, M.; Liu, X. Effects of long-term atorvastatin treatment on cardiac aging. Exp. Ther. Med. 2013, 6, 721–726. [Google Scholar] [CrossRef]
  308. Gong, X.; Ma, Y.; Ruan, Y.; Fu, G.; Wu, S. Long-term atorvastatin improves age-related endothelial dysfunction by ameliorating oxidative stress and normalizing eNOS/iNOS imbalance in rat aorta. Exp. Gerontol. 2014, 52, 9–17. [Google Scholar] [CrossRef]
  309. Kong, F.; Ye, B.; Lin, L.; Cai, X.; Huang, W.; Huang, Z. Atorvastatin suppresses NLRP3 inflammasome activation via TLR4/MyD88/NF-κB signaling in PMA-stimulated THP-1 monocytes. Biomed. Pharmacother. 2016, 82, 167–172. [Google Scholar] [CrossRef]
  310. Luo, B.; Li, B.; Wang, W.; Liu, X.; Liu, X.; Xia, Y.; Zhang, C.; Zhang, Y.; Zhang, M.; An, F. Rosuvastatin alleviates diabetic cardiomyopathy by inhibiting NLRP3 inflammasome and MAPK pathways in a type 2 diabetes rat model. Cardiovasc. Drugs Ther. 2014, 28, 33–43. [Google Scholar] [CrossRef]
  311. Lv, Z.-H.; Phuong, T.A.; Jin, S.-J.; Li, X.-X.; Xu, M. Protection by simvastatin on hyperglycemia-induced endothelial dysfunction through inhibiting NLRP3 inflammasomes. Oncotarget 2017, 8, 91291. [Google Scholar] [CrossRef]
  312. Abderrazak, A.; Couchie, D.; Mahmood, D.F.D.; Elhage, R.; Vindis, C.; Laffargue, M.; Matéo, V.; Büchele, B.; Ayala, M.R.; El Gaafary, M. Anti-inflammatory and antiatherogenic effects of the NLRP3 inflammasome inhibitor arglabin in ApoE2. Ki mice fed a high-fat diet. Circulation 2015, 131, 1061–1070. [Google Scholar] [CrossRef]
  313. Luo, F.; Das, A.; Chen, J.; Wu, P.; Li, X.; Fang, Z. Metformin in patients with and without diabetes: A paradigm shift in cardiovascular disease management. Cardiovasc. Diabetol. 2019, 18, 54. [Google Scholar] [CrossRef] [PubMed]
  314. Jadhav, S.; Ferrell, W.; Greer, I.A.; Petrie, J.R.; Cobbe, S.M.; Sattar, N. Effects of metformin on microvascular function and exercise tolerance in women with angina and normal coronary arteries: A randomized, double-blind, placebo-controlled study. J. Am. Coll. Cardiol. 2006, 48, 956–963. [Google Scholar] [CrossRef] [PubMed]
  315. Mohan, M.; Al-Talabany, S.; McKinnie, A.; Mordi, I.R.; Singh, J.S.; Gandy, S.J.; Baig, F.; Hussain, M.S.; Bhalraam, U.; Khan, F. A randomized controlled trial of metformin on left ventricular hypertrophy in patients with coronary artery disease without diabetes: The MET-REMODEL trial. Eur. Heart J. 2019, 40, 3409–3417. [Google Scholar] [CrossRef] [PubMed]
  316. Preiss, D.; Lloyd, S.M.; Ford, I.; McMurray, J.J.; Holman, R.R.; Welsh, P.; Fisher, M.; Packard, C.J.; Sattar, N. Metformin for non-diabetic patients with coronary heart disease (the CAMERA study): A randomised controlled trial. Lancet Diabetes Endocrinol. 2014, 2, 116–124. [Google Scholar] [CrossRef]
  317. Holman, R.R.; Coleman, R.L.; Chan, J.C.; Chiasson, J.-L.; Feng, H.; Ge, J.; Gerstein, H.C.; Gray, R.; Huo, Y.; Lang, Z. Effects of acarbose on cardiovascular and diabetes outcomes in patients with coronary heart disease and impaired glucose tolerance (ACE): A randomised, double-blind, placebo-controlled trial. Lancet Diabetes Endocrinol. 2017, 5, 877–886. [Google Scholar] [CrossRef]
  318. Lahnwong, S.; Chattipakorn, S.C.; Chattipakorn, N. Potential mechanisms responsible for cardioprotective effects of sodium–glucose co-transporter 2 inhibitors. Cardiovasc. Diabetol. 2018, 17, 101. [Google Scholar] [CrossRef]
  319. Lee, H.; Kim, J.; An, J.; Lee, S.; Choi, D.; Kong, H.; Song, Y.; Park, I.H.; Lee, C.-K.; Kim, K. Downregulation of IL-18 expression in the gut by metformin-induced gut microbiota modulation. Immune Netw. 2019, 19, e28. [Google Scholar] [CrossRef]
  320. Li, J.; Lin, S.; Vanhoutte, P.M.; Woo, C.W.; Xu, A. Akkermansia muciniphila protects against atherosclerosis by preventing metabolic endotoxemia-induced inflammation in Apoe−/− mice. Circulation 2016, 133, 2434–2446. [Google Scholar] [CrossRef]
  321. Tang, W.W.; Bäckhed, F.; Landmesser, U.; Hazen, S.L. Intestinal microbiota in cardiovascular health and disease: JACC state-of-the-art review. J. Am. Coll. Cardiol. 2019, 73, 2089–2105. [Google Scholar] [CrossRef] [PubMed]
  322. Schwartz, M.; Gluck, M.; Koon, S. Norovirus gastroenteritis after fecal microbiota transplantation for treatment ofclostridium difficileinfection despite asymptomatic donors and lack of sick contacts. Off. J. Am. Coll. Gastroenterol. ACG 2013, 108, 1367. [Google Scholar] [CrossRef]
  323. De Leon, L.M.; Watson, J.B.; Kelly, C.R. Transient flare of ulcerative colitis after fecal microbiota transplantation for recurrent Clostridium difficile infection. Clin. Gastroenterol. Hepatol. 2013, 11, 1036–1038. [Google Scholar] [CrossRef]
  324. Ndeh, D.; Gilbert, H.J. Biochemistry of complex glycan depolymerisation by the human gut microbiota. FEMS Microbiol. Rev. 2018, 42, 146–164. [Google Scholar] [CrossRef]
  325. Hickey, C.A.; Kuhn, K.A.; Donermeyer, D.L.; Porter, N.T.; Jin, C.; Cameron, E.A.; Jung, H.; Kaiko, G.E.; Wegorzewska, M.; Malvin, N.P. Colitogenic Bacteroides thetaiotaomicron antigens access host immune cells in a sulfatase-dependent manner via outer membrane vesicles. Cell Host Microbe 2015, 17, 672–680. [Google Scholar] [CrossRef]
  326. Roberts, C.L.; Keita, Å.V.; Duncan, S.H.; O’Kennedy, N.; Söderholm, J.D.; Rhodes, J.M.; Campbell, B.J. Translocation of Crohn’s disease Escherichia coli across M-cells: Contrasting effects of soluble plant fibres and emulsifiers. Gut 2010, 59, 1331–1339. [Google Scholar] [CrossRef]
  327. Chassaing, B.; Koren, O.; Goodrich, J.K.; Poole, A.C.; Srinivasan, S.; Ley, R.E.; Gewirtz, A.T. Dietary emulsifiers impact the mouse gut microbiota promoting colitis and metabolic syndrome. Nature 2015, 519, 92–96. [Google Scholar] [CrossRef]
  328. Russell, B.J.; Brown, S.D.; Siguenza, N.; Mai, I.; Saran, A.R.; Lingaraju, A.; Maissy, E.S.; Machado, A.C.D.; Pinto, A.F.; Sanchez, C. Intestinal transgene delivery with native E. coli chassis allows persistent physiological changes. Cell 2022, 185, 3263–3277.e3215. [Google Scholar]
  329. Praveschotinunt, P.; Duraj-Thatte, A.M.; Gelfat, I.; Bahl, F.; Chou, D.B.; Joshi, N.S. Engineered E. coli Nissle 1917 for the delivery of matrix-tethered therapeutic domains to the gut. Nat. Commun. 2019, 10, 5580. [Google Scholar] [CrossRef]
  330. Ahmad, A.F.; Caparrós-Martin, J.A.; Gray, N.; Lodge, S.; Wist, J.; Lee, S.; O’Gara, F.; Dwivedi, G.; Ward, N.C. Gut microbiota and metabolomics profiles in patients with chronic stable angina and acute coronary syndrome. Physiol. Genom. 2024, 56, 48–64. [Google Scholar] [CrossRef]
  331. Ahmad, A.F.; Caparrós-Martin, J.A.; Gray, N.; Lodge, S.; Wist, J.; Lee, S.; O’Gara, F.; Shah, A.; Ward, N.C.; Dwivedi, G. Insights into the associations between the gut microbiome, its metabolites, and heart failure. Am. J. Physiol.-Heart Circ. Physiol. 2023, 325, H1325–H1336. [Google Scholar] [CrossRef]
  332. Wymore Brand, M.; Wannemuehler, M.J.; Phillips, G.J.; Proctor, A.; Overstreet, A.-M.; Jergens, A.E.; Orcutt, R.P.; Fox, J.G. The altered Schaedler flora: Continued applications of a defined murine microbial community. ILAR J. 2015, 56, 169–178. [Google Scholar] [CrossRef]
  333. Hsu, B.B.; Gibson, T.E.; Yeliseyev, V.; Liu, Q.; Lyon, L.; Bry, L.; Silver, P.A.; Gerber, G.K. Dynamic modulation of the gut microbiota and metabolome by bacteriophages in a mouse model. Cell Host Microbe 2019, 25, 803–814.e805. [Google Scholar] [CrossRef]
Figure 1. A schematic representation of the protective role of BAs in hypertriglyceridemia. The BA-associated activation of the Farnesoid X receptor and Pregnane X receptor attenuates pro-inflammatory NF-κB signaling and downregulates triglyceride levels.
Figure 1. A schematic representation of the protective role of BAs in hypertriglyceridemia. The BA-associated activation of the Farnesoid X receptor and Pregnane X receptor attenuates pro-inflammatory NF-κB signaling and downregulates triglyceride levels.
Ijms 25 10208 g001
Figure 2. Schematic representation of TMAO generation and associated pathophysiological correlations. High fat dietary intake leads to trimethylamine (TMA) generation upon gut microbial breakdown. TMA undergoes oxidization via hepatic flavin monooxygenase-3 (FMO3) and produces TMAO. This TMAO correlates with varied pathophysiological onset viz atherosclerosis, heart failure and platelet hyper-reactivity.
Figure 2. Schematic representation of TMAO generation and associated pathophysiological correlations. High fat dietary intake leads to trimethylamine (TMA) generation upon gut microbial breakdown. TMA undergoes oxidization via hepatic flavin monooxygenase-3 (FMO3) and produces TMAO. This TMAO correlates with varied pathophysiological onset viz atherosclerosis, heart failure and platelet hyper-reactivity.
Ijms 25 10208 g002
Figure 3. Schematic representation of SCFA-mediated cardioprotective signaling influence. SCFA-mediated GPR41 activation deactivates adenylyl cyclase and attenuates cAMP release, ameliorating blood pressure anomalies. SCFAs also augment regulatory T cell (Treg)-associated anti-inflammatory IL-10 signaling and attenuate cardiovascular injury.
Figure 3. Schematic representation of SCFA-mediated cardioprotective signaling influence. SCFA-mediated GPR41 activation deactivates adenylyl cyclase and attenuates cAMP release, ameliorating blood pressure anomalies. SCFAs also augment regulatory T cell (Treg)-associated anti-inflammatory IL-10 signaling and attenuate cardiovascular injury.
Ijms 25 10208 g003
Figure 4. Schematic representation of TMAO-driven CVD risk. TMAO increases NLRP3 inflammasome activity, mitochondrial ROS levels and NADPH oxidase activity and culminates in CVD onset viz atherosclerosis and deep vein thrombosis.
Figure 4. Schematic representation of TMAO-driven CVD risk. TMAO increases NLRP3 inflammasome activity, mitochondrial ROS levels and NADPH oxidase activity and culminates in CVD onset viz atherosclerosis and deep vein thrombosis.
Ijms 25 10208 g004
Table 1. A summary of the source, mechanism of action and CVD impact of key gut microbial metabolites associated with CVD onset and progression.
Table 1. A summary of the source, mechanism of action and CVD impact of key gut microbial metabolites associated with CVD onset and progression.
MetaboliteSourceMechanism of ActionImpact on CVDsReferences
Short-Chain Fatty acids (SCFAs)Gut microbial fermentation of dietary macromoleculesBlock cholesterol generation and their hepatic localization, promote Treg-induced anti-inflammatory signaling and reduce the local infiltration of immune cells, promote the expression of protein phosphatases, eNOS, PTEN and GSK-3β Downregulate cardiac hypertrophy, fibrosis and associated vascular dysfunction, inhibit onset of coronary diseases, reverses cardiomyocyte contraction defects[185,193,196,198]
Trimethylamine-N-oxide (TMAO)Gut-microbiota-driven breakdown of high-fat dietary food components via microbial TMA lyasesEnhances oxidative stress and ROS production via NLRP3 inflammasome activation, downregulates SIRT1 levels and promotes oxidative stress via the p53/p21/retinoblastoma (Rb) signaling, upregulates vascular oxidative stress via an upsurge in NADPH oxidase activity Promotes monocyte adhesion and deteriorates endothelium self-repair capacity, upregulates foam cell formation, augments macrophage LDL aggregation and is cumulatively predisposed towards atherosclerosis [214,216,217,218,219,220,225]
Bile acids (BAs)Amphipathic molecules generated in the liver from cholesterol metabolismActivate FXR, TGR5, modulation of Ca2+ inflow in ventricular cardiomyocytes, induce PXR activation, downregulating ApoA-IV, Cyp39a1 and augments CD36 Influence vascular tone and overall cardiovascular function, a reduction in secondary BA and a rise in primary BAs promotes CAD and hypercholesterolemia, increases total cholesterol, VLDL, and LDL, and decreases HDL that lead to atherosclerosis[229,230,232,234,235]
Indole-3-propionic acid (IPA)Gut-microbiota-driven tryptophan breakdownEnhances ABCA1 and macrophage-associated reverse cholesterol transportInduces myocardial contractility, the vasoconstriction of endothelium-denuded mesenteric arteries[242,243,244,245]
Hydrogen sulfide (H2S)Sulfate-reducing gut microbiotaInduces the downregulation of macrophage CX3CR1 expression, promotes PPAR-γ and downregulates NF-κB/ICAM-1 Exhibits vascular protective functions, regulates vascular tone, reduces leukocyte adhesion, mitochondrial ROS stress, decreases vascular proliferation and LDL oxidation.[255,256,257,259,260]
Phenylacetylglutamine (PAGln)Phenylalanine metabolism, primarily driven via bacterial PAGln synthaseInteracts and activates α2A, α2B and β2- adrenergic GPCRsAssociated with sarcomere functional defects, influences cardiac and platelet functions that eventually culminate to atherosclerosis [175,262,263,264]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Datta, S.; Pasham, S.; Inavolu, S.; Boini, K.M.; Koka, S. Role of Gut Microbial Metabolites in Cardiovascular Diseases—Current Insights and the Road Ahead. Int. J. Mol. Sci. 2024, 25, 10208. https://doi.org/10.3390/ijms251810208

AMA Style

Datta S, Pasham S, Inavolu S, Boini KM, Koka S. Role of Gut Microbial Metabolites in Cardiovascular Diseases—Current Insights and the Road Ahead. International Journal of Molecular Sciences. 2024; 25(18):10208. https://doi.org/10.3390/ijms251810208

Chicago/Turabian Style

Datta, Sayantap, Sindhura Pasham, Sriram Inavolu, Krishna M. Boini, and Saisudha Koka. 2024. "Role of Gut Microbial Metabolites in Cardiovascular Diseases—Current Insights and the Road Ahead" International Journal of Molecular Sciences 25, no. 18: 10208. https://doi.org/10.3390/ijms251810208

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop