Next Article in Journal
A Droplet Digital PCR-Based Approach for Quantitative Analysis of the Adulteration of Atlantic Salmon with Rainbow Trout
Previous Article in Journal
Antiproliferative Cancer Cell and Fungicidal Effects of Yellow and Red Araçá (Psidium cattleianum Sabine) Fruit Extract
Previous Article in Special Issue
Reducing Antigenicity and Improving Antioxidant Capacity of β-Lactoglobulin through Covalent Interaction with Six Flavonoids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of Gut Microbiota on Blood Cholesterol: A Review on Mechanisms

by
Chuanling Deng
1,
Jingjin Pan
1,
Hanyue Zhu
1,* and
Zhen-Yu Chen
2,*
1
School of Food Science and Engineering/National Technical Center (Foshan) for Quality Control of Famous and Special Agricultural Products (CAQS-GAP-KZZX043), Foshan University, Foshan 528011, China
2
School of Life Sciences, The Chinese University of Hong Kong, Shatin, NT, Hong Kong, China
*
Authors to whom correspondence should be addressed.
Foods 2023, 12(23), 4308; https://doi.org/10.3390/foods12234308
Submission received: 13 October 2023 / Revised: 24 November 2023 / Accepted: 27 November 2023 / Published: 29 November 2023
(This article belongs to the Special Issue Health Foods: Molecular Nutrition Mechanisms and Product Development)

Abstract

:
The gut microbiota serves as a pivotal mediator between diet and human health. Emerging evidence has shown that the gut microbiota may play an important role in cholesterol metabolism. In this review, we delve into five possible mechanisms by which the gut microbiota may influence cholesterol metabolism: (1) the gut microbiota changes the ratio of free bile acids to conjugated bile acids, with the former being eliminated into feces and the latter being reabsorbed back into the liver; (2) the gut microbiota can ferment dietary fiber to produce short-chain fatty acids (SCFAs) which are absorbed and reach the liver where SCFAs inhibit cholesterol synthesis; (3) the gut microbiota can regulate the expression of some genes related to cholesterol metabolism through their metabolites; (4) the gut microbiota can convert cholesterol to coprostanol, with the latter having a very low absorption rate; and (5) the gut microbiota could reduce blood cholesterol by inhibiting the production of lipopolysaccharides (LPS), which increases cholesterol synthesis and raises blood cholesterol. In addition, this review will explore the natural constituents in foods with potential roles in cholesterol regulation, mainly through their interactions with the gut microbiota. These include polysaccharides, polyphenolic entities, polyunsaturated fatty acids, phytosterols, and dicaffeoylquinic acid. These findings will provide a scientific foundation for targeting hypercholesterolemia and cardiovascular diseases through the modulation of the gut microbiota.

1. Introduction

The intestinal microbial structure embodies a complex bacterial consortium encompassing over 35,000 distinct bacterial species [1]. This consortium predominantly consists of four phyla, namely, Firmicutes, Bacteroides, Actinobacteria, and Proteobacteria, with Firmicutes and Bacteroides dominating, accounting for a substantial 90% of the total species count [2]. Intestinal microbes could profoundly impact host health, influencing not only host’s metabolic processes, including the absorption of nutrients, but also the metabolism of detrimental substances [3]. The composition and functionality of the gut microbiota are subjected to modulation by many factors, encapsulating both internal elements and external determinants such as genetics, age, diet, lifestyle, and medications [4]. Notably, diet is one of most pivotal factors underpinning alterations in the intestinal microbial structure [5,6]. Functional constituents in diets can regulate the growth and metabolic activities of the gut microbiota, thereby influencing the microbial composition. It is imperative to highlight that discernible disparities exist in the genetic content of intestinal microbes between adults and infants, reflecting the divergent intestinal functional needs at different life stages [7]. Through the production of an array of metabolic products such as short-chain fatty acids (SCFAs), branched-chain fatty acids (BCFAs), bile salt hydrolase (BSH), and lipopolysaccharides (LPS), the gut microbiota partakes in and regulates the host’s metabolism. Notably, SCFAs and BCFAs play a crucial role in this process and furnish vital nutritional sources for intestinal cells [8]. SCFAs can stimulate the proliferation and differentiation of intestinal epithelial cells, contributing to the maintenance of the mineral balance and the absorption of iron, calcium, and magnesium [9]. Prevalent SCFAs include acetate (Ac), propionate (Pr), and butyrate (Bu), whereas, in the colon, others such as valerate (Va), caproate (Ca), and isobutyrate are relatively low, constituting approximately 5–10% of the total SCFAs [10]. LPS is identified as an endotoxin derived from Gram-negative bacteria. Elevated LPS levels are associated with some metabolic diseases, inflammation, the infiltration of adipose macrophages, endothelial cell apoptosis, fatty-liver-related diseases, and insulin resistance [11,12,13]. Additionally, intestinal microbes can produce BSH, an enzyme capable of hydrolyzing N-acylamide bonds, facilitating the release of taurine or glycine (Gly) through the hydrolysis of bile salts [14].
Cholesterol is a sterol synthesized endogenously by animals, serving as an indispensable component of cell membranes. Beyond its structural role, it acts as a signaling molecule in various biological processes, including cellular transport and neurotransmission. It also functions as a precursor for synthesizing vitamin D, steroid hormones (such as progesterone and estrogen), and bile acids [15]. Due to its low water solubility, cholesterol predominantly circulates within the body through lipoproteins [16]. Depending on the density, size, and composition of these lipoproteins, they are classified into chylomicrons (CMs), very-low-density lipoproteins (VLDL), intermediate-density lipoproteins (IDL), low-density lipoproteins (LDL), and high-density lipoproteins (HDL) [17]. The homeostasis of cholesterol plays a vital role in physiological functions. It has been long known that an elevated serum cholesterol level, or hypercholesterolemia, is a principal cause of atherosclerosis and coronary heart disease [18]. Conversely, an overly reduced cholesterol level may also pose a health risk, including an increased susceptibility to hemorrhagic stroke and a correlation with higher mortality rates due to late-stage heart failure [19,20].
Many recent studies have unveiled a significant correlation between gut microbiota dysbiosis and cholesterol metabolism. In this regard, a profound exploration into the influence exerted by the intestinal microbiota on cholesterol metabolism is of paramount importance. In general, the gut microbiota can engage in cholesterol metabolism through the following possible pathways: modulating the ratio of free to conjugated bile acids [21]; enhancing the abundance of the SCFA-producing microbiota to increase the concentration of SCFAs in the intestinal lumen [22]; and regulating the expression of cholesterol-metabolism-related genes, epitomized by the activity of Lactobacillus strains [23]. In addition, the gut microbiota facilitates the conversion of cholesterol into fecal neutral sterols for excretion, while it concurrently reduces the production of LPS [24,25]. Therefore, optimizing the abundance of beneficial bacteria in the colon represents one of the primary strategies for cholesterol reduction. Currently, probiotics and prebiotics as health supplements are widely advocated [26]. Compared with the direct intake of supplements, bioactive compounds in natural foods boast higher bioavailability, potentially conferring additional health benefits [27]. For instance, polysaccharides, ubiquitously present in various plants, can be hydrolyzed and fermented to produce SCFAs by the intestinal microbiota [28]. Polyphenolic compounds, phytosterols (PS), and polyunsaturated fatty acids (PUFAs) are noted for their diversity and abundance, with the capacity to stimulate the growth of a beneficial microbiota [29,30]. Recent studies have highlighted the potential of dicaffeoylquinic acid (DCQA), a functional component found in Ilex kudingcha, in promoting the growth of SCFA-producing bacteria [31]. Beyond cholesterol modulation, these natural bioactive components also exhibit some immunomodulatory and anti-inflammatory activities [32], playing a positive role in preventing and treating various diseases. Integrating these natural functional components into strategies for modulating the intestinal microbiota to regulate cholesterol metabolism opens a new avenue for therapeutic intervention.

2. Pathways of Cholesterol Metabolism

Epidemiological studies have evidenced a compelling association between elevated cholesterol levels and cardiovascular diseases (CVDs), the latter being a predominant cause of mortality and disability in developed countries, projected by health organizations to persist until 2030 [33,34]. Cholesterol in humans primarily derives from two sources: endogenous cholesterol synthesized in the liver and intestine, and exogenous cholesterol acquired through the consumption of animal-derived foods [35]. Cholesterol homeostasis is vital for the physiological balance among hepatic cholesterol synthesis, absorption, transport, and biliary excretion (Figure 1).

2.1. Cholesterol Synthesis

The liver is pivotal in cholesterol homeostasis, spearheading its synthesis and conversion into bile acids [36]. Cholesterol biosynthesis involves a cascade of enzymatic reactions [37]. Initially, acetyl-CoA is transformed into mevalonate (MVA) though several reactions. This transformation has two acetyl-CoA molecules merging into acetoacetyl-CoA, which then combines with another acetyl-CoA, facilitated by HMG-CoA synthase (HMG-CoA-S), to form HMG-CoA. This compound is subsequently reduced to MVA by HMG-CoA reductase (HMG-CoA-R), the rate-limiting enzyme crucial for averting excessive cholesterol synthesis and accumulation [38,39,40]. MVA then morphs into isopentenyl pyrophosphate (IPP), with mevalonate kinase (MK), phosphomevalonate kinase (PMK), and mevalonate diphosphate decarboxylase (MDD) governing this transition [41,42]. Through condensation, IPP evolves into squalene (SQ), a 30-carbon precursor to all steroids. SQ is then oxidized to 2,3-oxidosqualene (OS) by squalene monooxygenase (SM) and cyclized to lanosterol by oxidosqualene cyclase-lanosterol synthase (OSC) [43,44,45]. Lanosterol ultimately transforms into cholesterol after a series of complex reactions [46]. The lanosterol-to-cholesterol conversion stage is a complex process with numerous enzymes at play and still requires further elucidation regarding its structure and mechanisms.

2.2. Cholesterol Absorption

Cholesterol absorption in the small intestines is an intricately regulated physiological process governed at the cellular level by a series of proteins. Cholesterol within the intestinal tract primarily originates from dietary intake, biliary secretion, and the intestinal mucosal epithelium. Western diets are estimated to contribute approximately 300–500 mg of cholesterol daily, while the bile provides 800–1200 mg, and the intestinal mucosal epithelium adds around 300 mg [47]. Cholesterol absorption begins in the stomach, forming micelles after being emulsified by bile acids in the small intestine. It is important to note that only non-esterified cholesterol can form these micelles [48]. These micelles subsequently interact with the Niemann-Pick C1-like 1 protein (NPC1L1), a pivotal transporter in cholesterol absorption processes, facilitating the micelles’ transport into the intestinal epithelial cells [49]. NPC1L1 predominantly localizes at the apical membrane of enterocytes in the small intestine [50]. Entering the intestinal epithelial cells, cholesterol is esterified by acyl-coenzyme A: cholesterol acyltransferases 2 (ACAT2) in the endoplasmic reticulum, forming cholesterol ester (CE). Subsequently, microsomal triglyceride transfer protein (MTP) transfers CE into chylomicrons (CMs), which enter the lymphatic system and the bloodstream, and are transported to the liver [51,52]. Two forms of ACAT enzymes, namely, ACAT1 and ACAT2, have been identified in mammals. While ACAT1 is ubiquitously expressed in various tissues, ACAT2 is predominantly found in intestinal epithelial cells and hepatocytes [53]. MTP is a lipid transfer protein responsible for transporting CE from the endoplasmic reticulum to nascent apoB lipoproteins, further facilitating the assembly of CM [54]. ApoB lipoproteins primarily mediate the transportation and metabolism of cholesterol and triglycerides [55]. Non-esterified cholesterol, on the other hand, is transported back into the intestinal lumen by ATP-binding cassette transporters G5 and G8 (ABCG5/8) [51]. These transporters, functioning as heterodimers, are prominently expressed in the microvilli of intestinal cells and the canalicular membrane of hepatocytes, playing a collective role in cholesterol excretion [56].

2.3. Cholesterol Excretion

Human body excretes approximately one gram of cholesterol daily, half of which is transformed into bile acids (BAs) and eliminated through feces [57]. At the same time, the remainder exists unesterified within fecal matter. In the liver, cholesterol 7α-hydroxylase (CYP7A1) and sterol 27-hydroxylase (CYP27A1) catalyze the 7-α-hydroxylation and 27-hydroxylation of cholesterol, respectively, further synthesizing primary bile acids, cholic acid (CA), and chenodeoxycholic acid (CDCA). These primary bile acids accumulate in the bile, conjugated with glycine (Gly) or taurine [58,59,60]. Some primary bile acids undergo deconjugation and 7α-dehydroxylation in the intestinal tract, generating secondary bile acids, namely, deoxycholic acid (DCA) and lithocholic acid (LCA) [61,62]. Both primary and secondary bile acids are partially absorbed in the ileum and returned to the liver through the portal venous system [57]. Due to its insolubility, LCA is generally poorly reabsorbed [63]. Bile acids that are not absorbed are excreted as fecal acidic sterols [64]. Approximately 3–5 g of bile acids circulate within the intestine multiple times (between 6–10 cycles), a process under sophisticated feedback regulation [65]. Within this regulatory framework, CYP7A1, acting as the rate-limiting enzyme for bile acid biosynthesis, has its activity negatively modulated by the nuclear bile acid receptor, Farnesoid X receptor (FXR). When the bile acid pool within the enterohepatic circulation increases, FXR is activated, thereby inhibiting the transcriptional activity of the CYP7A1 gene [66]. CDCA is crucial in activating FXR, its most potent ligand [67].
A second pathway involves the excretion of cholesterol by intestinal cells, manifesting as fecal-neutral sterols (FNSs) [68]. Cholesterol that was unabsorbed by the small intestine is transported to the intestinal lumen by ABCG5/8, eventually excreted as fecal-neutral sterols [51]. This process is positively regulated by Liver X receptor α (LXRα), a principal regulator participating in the mRNA expression of ABCG5/8 [69]. Additionally, ABCG5/8 facilitates the secretion of cholesterol and phytosterol into the bile [70]. The overexpression of ABCG5/8 reduces the absorption of dietary cholesterol [71].

3. Mechanisms through Which Gut Microbiota Influences Cholesterol Metabolism

In recent years, abundant research has centered around understanding how gut microbiota communities influence human health. One emerging piece of evidence is that the gut microbiota can affect cholesterol metabolism. These microbes engage in cholesterol metabolism through various mechanisms to reduce plasma cholesterol levels. It has been elucidated that Lactobacillus alone embodies multiple distinct mechanisms for cholesterol removal [72]. In general, the gut microbiota achieves cholesterol reduction through the following mechanisms: transforming complex non-digestible polysaccahrides into monosaccharides and fermenting them to produce beneficial SCFAs [73]; generating BSH, which facilitates the deconjugation of conjugated bile acids and releases free bile acids [74]; participating in the regulation of gene expressions associated with cholesterol metabolism [75]; promoting the conversion of cholesterol into fecal sterols [76]; and influencing the production of LPS, which affects cholesterol levels [25].

3.1. Participation of Gut Microbiota in Modifying Conjugated Bile Acids

Some gut microbes can produce BSH and hydrolyze the conjugated bile acids into free bile acids, thus increasing the ratio of free bile acids to conjugated bile acids, leading to a greater excretion of bile acids, and resulting in a smaller pool of both cholesterol and bile acids [14]. This is because the free bile acids are mostly excreted into feces, whereas the latter ones are reabsorbed back into the liver [64]. A portion of hepatic cholesterol undergoes conversion into BA. In conjunction with glycine (Gly) and taurine, these BA molecules form conjugated bile acids (C-BAs), which enter the small intestine. Lactic acid bacteria (LAB) secrete BSH within the intestinal environment. Under the catalytic influence of BSH, C-BA deconjugates, giving rise to free BA, Gly, and Taurine. The generated free BA is subsequently excreted from the body, while the remaining C-BA recirculates to the liver via the portal vein (Figure 2). In addition, the liver plays a pivotal role in maintaining cholesterol homeostasis and orchestrating cholesterol synthesis and its conversion into bile acids for elimination [77]. The conjugation process results in a reduced acid dissociation constant (pKa) and the complete ionization of these acids, which exist in the form of anions [78,79,80]. Bile acids undergo various biochemical modifications in the human large intestine, including deconjugation, 7α/β-dehydroxylation, and epimerization [21]. Deconjugation is achieved through the enzymatic hydrolysis of the C24 N-acylamide bond that links bile acids with their conjugated amino acids. These deconjugated primary bile acids function as signaling molecules, reflecting the body’s total bile acid levels and increasing the concentrations of cholic acid (CA) and chenodeoxycholic acid (CDCA). Furthermore, glycine and taurine released during deconjugation serve as nutrient sources for the intestinal microbiota [81]. Approximately 26.03% of the total bacterial population in the large intestine exhibits BSH activity [82]. BSH with deconjugation capabilities primarily resides in Gram-positive bacteria, including Bifidobacterium, Lactobacillus, Clostridium, Enterococcus, and Listeria [83,84,85,86,87,88]. Nonetheless, BSH activity is not exclusive to Gram-positive bacteria; Gram-negative bacteria like Stenotrophomonas, Bacteroides, and Brucella also exhibit BSH activity [89,90,91]. All BSH deconjugation reactions depend on the hydrolysis of the N-acylamide bond, releasing taurine or glycine, with the reaction exhibiting maximum activity in neutral or mildly acidic environments (pH 5–7), with an optimal pH of approximately 6 [83,92].
Scientists have identified a gene cluster in Clostridium scindens (C. scindens), the bai operon, crucially implicated in bile acid dehydroxylation. This operon encodes for a multitude of enzymes essential for the dehydroxylation process [93]. The baiG gene within this cluster encodes bile acid transport proteins, facilitating the uptake of CA by bacterial strains and transporting CDCA and DCA [94]. Under the influence of baiB, bile acids are oxidized to form cholyl-coenzyme A (CoA), which is then further oxidized by baiA2 to produce 3-oxo-cholyl-CoA. Subsequently, baiCD catalyzes the oxidation of 3-oxo-cholyl-CoA to form 3-oxo-Δ4-cholyl-CoA. Then, baiF transfers CoA from 3-oxo-Δ4-cholyl-CoA to CA, resulting in the formation of 3-oxo-Δ4-CA and cholyl-CoA [95]. This 3-oxo-Δ4-CA, under the action of baiE, undergoes dehydroxylation to yield 3-oxo-Δ4,6-DCA, the rate-limiting step in the process [93]. Following this, continuous activity by baiN on 3-oxo-Δ4,6-DCA produces 3-oxo-DCA, which is then converted to DCA through the combined effort of baiO and baiA2, occurring at the C7 position, known as 7α-dehydroxylation [96,97]. 7β-dehydroxylation occurs similarly, with the primary distinction being the utilization of baiH instead of baiCD for the oxidation at the C4 position, with the activity of the 7β-dehydratase enzyme possibly serving as the rate-limiting step for 7β-dehydroxylation. Currently, bacteria such as C. scindens, C. hylemonae, C. perfringens, and P. hiranonis have been observed to produce enzymes capable of facilitating the 3α-dehydrogenation of hydroxysteroids, a crucial step within the 7α-dehydroxylation pathway [98,99,100].
In the metabolic pathways of bile acids, positional isomerization, a significant biochemical process, gives rise to many functional derivatives. This mechanism primarily hinges on the action of location-specific hydroxysteroid dehydrogenases (HSDHs), such as 7α-HSDH, which oxidize hydroxyl groups [101]. This process is then followed by the reduction facilitated by another location-specific hydroxysteroid dehydrogenase, 7β-HSDH. Enzymes analogous to these include 3α/β-HSDH and 12α/β-HSDH [102,103]. Through positional isomerization, CA can be transformed into various derivatives, including ursodeoxycholic acid (UCA), 12-epi-cholic acid (12-ECA), or iso-cholic acid (iCA). Similarly, CDCA can undergo isomerization to yield ursodeoxycholic acid (UDCA) or iso-chenodeoxycholic acid (iCDCA) [81]. These isomerization reactions enhance the diversity and metabolism of bile acids and further promote cholesterol metabolism. Existing studies corroborate that specific intestinal micro-organisms, such as Clostridium baratii, can isomerize CDCA to UDCA [104]. In addition, several other gut microbes—including Ruminococcus, Clostridium, Stenotrophomonas maltophilia, and Collinsella aerofaciens—have been verified to generate UDCA through the activity of 7α/β-HSDH.

3.2. Microbial Production of SCFAs and Their Effects on Cholesterol Metabolism

SCFAs in intestines are pivotal in sustaining human health. Specifically, various bacteria, including Alloprevotella, Bacteroides, Clostridium, Eubacterium, Faecalibacterium, and Roseburia, are known to produce these beneficial SCFAs, with Bu being a prominent member [105]. Bu has demonstrated its therapeutic potential in various diseases, including gastrointestinal disorders, the regulation of carbohydrate metabolism, and an improvement in obesity [106]. Further research indicates a connection between Bu and cholesterol metabolism. Previous studies have revealed that Bu can reduce the serum low-density lipoprotein cholesterol (LDL-C) level, a crucial risk factor for cardiovascular diseases [107]. Currently, statins are the preferred treatment for lowering LDL-C, primarily by inhibiting HMG-CoA-R, consequently upregulating the expression of LDL receptors (LDLRs), which enhances LDL uptake from the circulation, ultimately reducing LDL-C levels in the plasma [108,109]. Furthermore, the sterol-regulatory element binding protein-2, a key regulator of cholesterol metabolism and homeostasis, increases LDLR expression upon activation [110]. SCFAs, such as Bu as exemplified in Figure 3, generated by the gut microbiota participate in cholesterol metabolism through two distinct pathways. Firstly, Bu acts to inhibit the expression of HMG-CoA-R, thus further suppressing cholesterol synthesis, ultimately leading to reduced cholesterol levels. Secondly, Bu influences the activity of SREBP-2, thereby promoting the expression of LDL-R. The upregulation of LDL-R expression accelerates the uptake of LDL from the bloodstream, ultimately resulting in lowered levels of LDL-C (Figure 3) [22]. It is noteworthy that the mechanism of Bu differs significantly from that of statins.
Beyond Bu, other SCFAs have also exhibited cholesterol-lowering properties. For instance, it has been shown that injecting Pr into the ceca of rats fed with a casein-based diet results in a noticeable reduction in plasma cholesterol levels [111]. Furthermore, Ac could inhibit hepatic lipid synthesis and reduce TC and TG levels in mice given a high-fat diet [112]. The supplementation of SCFAs with two to four carbons into the diet reduces blood cholesterol in hamsters [113]. As other SCFAs like valerate (Va), caproate (Ca), and isobutyrate are quantitatively very low in the colon, no sufficient research data support their cholesterol-lowering activities, and this warrants further investigation.

3.3. Gene Expression Involvement of Lactobacillus in Cholesterol Metabolism

Research has been shown that lactic acid bacteria (LAB) can remarkably mitigate cholesterol levels via several mechanisms, including assimilation, absorption, and co-precipitation [114,115]. One study has unveiled that Lactobacillus fermentum SM-7 can absorb and co-precipitate up to 38.5% of cholesterol and assimilate an additional 60% [116]. LAB also plays a crucial role in cholesterol reduction by regulating the gene expression of enzymes involved in cholesterol synthesis, absorption, and excretion. The phosphorylation activity of AMPK governs various regulators and transcription factors implicated in lipid metabolism [117]. In this regard, Lactiplantibacillus plantarum DR7 can downregulate the mRNA of HMG-CoA-R by mediating AMPK phosphorylation, subsequently lowering cholesterol levels [118]. Moreover, approximately 50% of daily dietary cholesterol is absorbed through the intestines, with the remainder being excreted through feces [23]. Dietary cholesterol requires specific binding with NPC1L1 in intestinal epithelial cells for absorption, whereas it requires ABCG5/G8 to shuttle cholesterol back to the lumen of the intestine for elimination [119,120]. Noteworthy is the discovery that LAB, through the activation of PPAR and LXR, influences the expression of ABCG5/G8 and NPC1L1, playing a significant role in cholesterol excretion and absorption processes [121].
SREBPs, expressed principally in the liver, encompass three subtypes: SREBP-1a, SREBP-1c, and SREBP-2 [122]. SREBP-1a can effectively activate all SREBP-responsive genes, inclusive of those involved in the synthesis of cholesterol, fatty acids, and triglycerides. On the other hand, SREBP-1c prioritizes the transcription activation of genes necessary for fatty acid synthesis without activating cholesterol-related genes. In contrast, SREBP-2 primarily activates LDL-R genes and those requisite for cholesterol synthesis [123]. Both Lactobacillus plantarum NCU116 and L. brevis SBC8803 have been demonstrated to impede cholesterol accumulation by influencing SREBP expressions, ultimately reducing the cholesterol concentration [124,125]. Lastly, CYP7A1, an enzyme facilitating bile acid synthesis, is integral in maintaining mammal cholesterol homeostasis [126]. Notably, LXRα and FXR act as positive and negative regulators in cholesterol metabolism, modulating the expression of CYP7A1 mRNA [127]. It has been shown that Lactobacillus plantarum H6 could increase bile acid synthesis and CYP7A1 expression by suppressing FXR target gene expression [128]. Furthermore, the transcription of CYP7A1 is negatively regulated by FGF15 signaling. The research conducted by Kim et al. found that Lactobacillus rhamnosus GG could suppress FGF15 expression, promoting an increase in CYP7A1 expression in the liver, and reducing total cholesterol levels [129].

3.4. Probiotic Conversion of Cholesterol to Coprostanol

Coprostanol possesses a distinctive cis A/B ring configuration in its chemical structure, prompting the shift of 3-OH from the axial to the equatorial position. This unique structural adjustment potentially hinders the incorporation of coprostanol into mucosal cells, consequently limiting its absorption in the intestines [130]. Hence, it is perceived that this transformative process is an effective approach to reduce plasma TC levels because the gut microbiota could transform cholesterol to coprostanol [24,131]. It transpires that the microbial conversion of cholesterol to coprostanol in the intestine is mediated by three primary pathways. The process of converting cholesterol into coprostanol can be categorized into one direct pathway and two indirect pathways. In the direct pathway, cholesterol undergoes reduction, specifically targeting the 5–6 double bond, resulting in the formation of coprostanol. The first indirect pathway involves a series of reactions catalyzed by various enzymes, including cholesterol oxidase, encompassing oxidation, isomerization, and reduction processes that ultimately lead to the production of coprostanol. The second indirect pathway is through the allocholesterol pathway, distinct from the standard cholesterol pathway, leading to the reduction of cholesterol into coprostanol. (Figure 4). Two of these are indirect: Initially, cholesterol is oxidized to intermediary 5-alpha-cholestan-3-one under the influence of cholesterol oxidase; subsequently, 5-alpha-cholestan-3-one undergoes isomerization to form 4-cholesten-3-one, which is then reduced to coprostanone; and, finally, coprostanone is further reduced to coprostanol [132,133]. Another route involves the isomerization of cholesterol to allocholesterol, followed by the reduction of allocholesterol to coprostanol [134,135,136]. Additionally, a direct pathway exists, where cholesterol is transformed into coprostanol through the direct reduction of the 5–6 double bond; however, this pathway has been less extensively researched [137,138]. In this regard, Eubacterium coprostanoligenes ATCC 51222 could convert 90% of cholesterol into coprostanol in the medium [133] and E. ATCC 21408 could directly convert cholesterol into coprostanol through intermediary steps involving 4-cholesten-3-one and coprostanone [134]. Furthermore, a mixed culture of Lactobacillus acidophilus 43121, Lactobacillus casei, and Bifidobacterium could reduce TC levels and augment coprostanol excretion [130,139]. Studies have shown that probiotics like Bifidobacterium, Lactobacillus, and Clostridium can convert cholesterol to coprostanol under in vitro conditions [140,141]. Although the probiotic-mediated conversion of cholesterol to coprostanol is substantiated as an effective cholesterol-lowering mechanism, the challenge still exists, including the identification of specific microbial strains and enzymes involved in the process. Given the high oxygen sensitivity of these micro-organisms, the number of strains successfully isolated for the purpose of cholesterol reduction is considerably limited. Therefore, future work necessitates a continuation in the research on microbial strain isolation and the comprehensive genomic analysis of these strains to elucidate their precise roles in cholesterol reduction.

3.5. Lipopolysaccharides’ Involvement in Cholesterol Metabolism

A healthy gut microbiota is associated with a low production of LPS. In general, LPS is a component embedded within the outer membrane of Gram-negative bacteria [142]. Existing literature illustrates LPS’s dynamic engagement with lipids in the bloodstream through various mechanisms. Earlier studies had shown that LPS could increase LDL cholesterol, whereas it decreased HDL cholesterol, presumably by promoting HMG-CoA reductase [143]. In addition, the concentration of triacylglycerols in the blood can also be modulated through distinct pathways activated by LPS. One study has delineated that lower doses of LPS could stimulate the hepatic synthesis of very-low-density lipoprotein (VLDL), whereas higher doses inhibited lipoprotein metabolic degradation [144]. It is most likely that the gut microbiota modulates plasma cholesterol and triacylglycerol partially by affecting the production of LPS.
LPS is also proinflammatory. Numerous studies have demonstrated that LPS possesses a significant binding affinity with TC. Upon binding, these LPS-TC complexes are transported via the lymphatic system, potentially inducing inflammatory responses [145]. Further research indicates that LPS can activate toll-like receptors 4 and 9 (TLR4 and TLR9), subsequently triggering the NLR family pyrin domain-containing 3 (NLRP3) inflammasome, a process believed to be involved in the fibrotic progression of non-alcoholic fatty liver disease (NAFLD) [146]. Studies by Yoshida et al. discovered that strains Bacteroides vulgatus and Bacteroides dorei could reduce the concentration of LPS produced by the intestinal microbiota [25]. This function might positively contribute to the alleviation of atherosclerosis [25]. However, research exploring cholesterol reduction through the modulation of intestinal LPS levels remains scant.

4. Natural Functional Constituents Influencing Gut Microbiota in the Regulation of Cholesterol Metabolism

Natural functional components, encompassing indigestible polysaccharides, phenolic compounds, unsaturated fatty acids, and phytosterols, have manifested as being capable of fostering the proliferation of probiotics in the intestine. They fortify the human immune system by activating or modulating immune cells and responses. Furthermore, these functional constituents can be deployed as adjunctive measures to prevent cardiovascular diseases and certain inflammatory conditions. Regarding their effect on plasma cholesterol mediated by the gut microbiota, these constituents primarily exert their influence by enhancing the proliferation of SCFA-producing strains, modulating strains involved in cholesterol metabolism, promoting BSH-producing strains, and facilitating the conversion of cholesterol to coprostanol.

4.1. Indigestible Polysaccharides

Indigestible polysaccharides, abundant natural prebiotics, positively sway the gut microbiota and their metabolism. They not only adjust the microbiota composition but also promote the growth of beneficial bacteria. Upon consumption, indigestible polysaccharides could reach the large colon where they are fermented by the gut microbiota, producing SCFAs, primarily by Bacteroides and Firmicutes members. Bacteroides Thetaiotaomicron produce propionate (Pr) and acetate (Ac), which are subsequently transformed into butyrate by Eubacterium rectale [147]. Furthermore, an increase in butyrate production is observed with F. prausnitzii [148]. It is imperative to highlight that butyrate can be produced by the intestinal microflora through fiber fermentation or the Wood–Ljungdahl pathway [149]. Studies corroborate the efficacy of SCFAs in the reduction of plasma cholesterol. Seaweed polysaccharides have been demonstrated to be capable of alleviating gut microbiota dysbiosis and reducing cholesterol [150]. For instance, polysaccharides derived from red algae could increase the production of SCFAs, favorably modulate the gut microbiota, and reduce cholesterol [151]. Additionally, polysaccharides from Porphyra could alleviate the diet-induced intestinal dysbiosis by enhancing the population of Eubacterium xylanophilum, a known butyrate producer [152]. Alginate oligosaccharides derived from brown algae could elevate the BSH activity and enhance the CYP7A1 activity, facilitating bile acid synthesis and cholesterol reduction [153]. BSH is predominantly produced in the intestine by a consortium of bacteria, including Bacteroides, Bifidobacterium, Clostridium, Enterobacter, Enterococcus, and Lactobacillus [154]. Beyond seaweeds, recent findings emphasize the ability of polysaccharides from edible fungi to regulate the gut microbiota composition. Research has shown that polysaccharides from Auricularia auricula could stimulate the growth of SCFA-producing bacteria like Oscillibacter and Lactobacillus, ultimately enhancing the production of intestinal SCFAs and thereby reducing cholesterol [155]. Similar functions have been also observed with polysaccharides derived from mushrooms and Pleurotus eryngii [156,157].

4.2. Polyphenolic Compounds

Polyphenolic compounds are significantly active entities characterized by antioxidant and anti-inflammatory properties found extensively within various plants. These compounds, classified into phenolic acids, flavonoids, tannins, and lignans based on their unique compositional and structural characteristics, play vital roles in human health [29]. Tea leaves are rich sources of various polyphenolic compounds, including catechins, epicatechins, and quercetin (QR) [158]. Tzounis et al. discovered that catechins enhance the proliferation of the Blautia coccoides-Eubacterium rectale group and Bifidobacterium in the gut, with the former being recognized for increasing the concentration of SCFAs within the intestinal environment [159,160]. Furthermore, their research unveils the significant role of flavanols in cocoa: these compounds not only elevate the levels of Bifidobacterium and Lactobacillus in the gut but also inhibit the growth of certain pathogenic micro-organisms [161]. Moreover, red wine is a substantial source of polyphenolic compounds, including but not limited to resveratrol, proanthocyanidins, and flavanols [162]. An analysis conducted by MI Queipo-Ortuño and colleagues on the impact of red wine extracts on the intestinal microbiota revealed that individuals who consumed red wine over a continuous four-week period exhibited significant increments in the levels of Enterococcus, Prevotella, Bifidobacterium, Bacteroides uniformis, Eggerthella lenta, and Blautia coccoides-Eubacterium rectale within their gut. Simultaneously, there was a discernible decrease in the levels of TC, TG, and HDL, a trend closely correlated with the presence of SCFA-producing bacterial species [163]. However, it is crucial to acknowledge that not all plant-derived polyphenols yield positive effects on the regulation of the intestinal microbiota. Researchers found that QR mildly inhibits the growth of Bifidobacterium and Enterococcus while myricetin suppresses the growth of all LAB without adversely affecting harmful bacteria like Salmonella [164,165]. In conclusion, the interactions between polyphenolic compounds and the gut microbiota are multifaceted, encapsulating both positive and negative effects. These complex interactions necessitate further exploration and research for a deeper understanding of their true impact on human health.

4.3. Unsaturated Fatty Acids

Improving dietary fat quality by increasing the intake of polyunsaturated fatty acids (PUFAs) while reducing saturated fatty acids (SFAs) significantly decreases serum cholesterol levels [30]. Recent research data illuminate the function of dietary fats as potential modulators of the human gut microbiota composition, with their total amounts and quality acting as pivotal factors in shaping microbial communities in the gut [166,167]. Studies reveal that a higher PUFA intake not only amplifies the total bacterial count within the gut flora but also fosters the proliferation of beneficial bacterial species [168]. Principal sources of PUFAs encompass aquatic species, micro-organisms, algae, and oil crops [169,170,171,172]. Notably, alpha-linolenic acid (ALA), gamma-linolenic acid (GLA), linoleic acid (LA), eicosapentaenoic acid (EPA), and docosahexaenoic acid (DHA) are deemed beneficial for health [172]. It is imperative to acknowledge that PUFAs are categorized into two primary families, ω 6 (n-6) and ω 3 (n-3), with EPA and DHA belonging to omega-3 unsaturated fatty acids [173,174]. Within the human gut, Lachnospiraceae and Bifidobacterium are identified as beneficial bacteria. The abundance of Lachnospiraceae and Bifidobacterium negatively correlates with LDL levels [175,176]. These bacterial classes contribute to cholesterol reduction by transforming it into coprostanol [177]. Research conducted by Watson et al. observed a notable increase in the abundance of both Lachnospiraceae and Bifidobacterium in healthy individuals upon omega-3 PUFA intake [178]. Similarly, a study by Tindall et al. demonstrated that consuming ALA-rich walnuts increases Lachnospiraceae [179]. Moreover, studies by Wan et al. found that both EPA and DHA increase Lachnospiraceae abundance and positively correlate with the proliferation of various lactic acid-producing bacteria [180]. Research by Li et al. disclosed that Spirulina, rich in LA and GLA, could enhances the abundance of several beneficial bacterial groups in the gut, including Prevotella, Porphyromonadaceae, Barnesiella, and Parasutterella [181,182]. Particularly, Prevotella, negatively correlated with serum biochemical indicators, promotes bile acid synthesis, further regulating cholesterol metabolism [183].

4.4. Phytosterol

Phytosterol (PS) are renowned for their potent cholesterol-lowering effects [184,185]. The primary components of PS include β-sitosterol, stigmasterol, campesterol, and brassicasterol, among others [186]. Studies suggest a daily intake of 2 g of PS can effectively reduce cholesterol levels, particularly TC and LDL-C, by 6–15% [187]. It is noteworthy that lotus seeds, being rich in various bioactive compounds including alkaloids, flavonoid compounds, and PS, are considered excellent food and medicinal sources. Research conducted by Liu et al. revealed that PS in lotus seed cores significantly enhances the abundance of beneficial bacterial phyla in the gut, including Firmicutes, Bacteroides, Actinobacteria, and Proteobacteria [188]. Firmicutes suppress Clostridium perfringens growth, thus maintaining intestinal homeostasis [189]. Bacteroides are involved in the metabolism of bile acids and the bioconversion of steroidal compounds. At the same time, certain bacteria within the Actinobacteria phylum are known to lower blood sugar and lipid levels. Additionally, soy is a commendable source of PS due to its high content, availability, and safety [190]. Research indicates that upon soy PS intake, there is an increase in the abundance of beneficial gut microbes like Lactobacillus, Oscillibacter, and Ackermanella [191]. Importantly, an increase in Ackermanella correlates positively with significant improvements in lipid metabolism and the restoration of colonic mucosal barrier functions [192].

4.5. Dicaffeoylquinic Acid

Kuding Tea (KDC), popular in China and Southeast Asian nations like Singapore and Malaysia, is recognized as a functional tea beverage known for its multiple pharmacological activities including dispelling wind-heat, quenching thirst, eliminating phlegm, and boosting alertness [193]. KDC is rich in caffeoylquinic acid derivatives with antioxidant activities, such as 3-CQA, 5-CQA, 3,4-diCQA, 3,5-diCQA, and 4,5-diCQA [194]. Research conducted by Xie et al. discovered that dicaffeoylquinic acid (DCQA) in Kuding tea modulates cholesterol metabolism in mice and promotes the growth of beneficial gut microbes like Bifidobacterium and Akkermansia muciniphila [31]. These microbial populations’ alterations subsequently influence microbial community functions, including bile acid biosynthesis. Notably, the genus Odoribacter, belonging to the Porphyromonadaceae family, is identified as a primary producer of Ac, Pr, and Bu, which are SCFAs proven to lower cholesterol levels effectively [195,196]. DCQA adjusts the relative abundance of gut microbes like Odoribacter, Prevotella, Bacteroides, Parasutterella, and Lachnospiraceae, effectively ameliorating gut dysbiosis [194]. Furthermore, DCQA alters the functional characteristics of the gut microbial community, providing a potential mechanism foundation for maintaining gut health and regulating cholesterol levels.

5. Conclusions

Gut microbiota dysbiosis is a risk factor in the pathophysiological processes related to cholesterol-associated diseases, constituting a subtle and potential mechanism of disease onset. This mechanism, directly or indirectly, influences human health. Notably, in cardiovascular diseases, abnormal cholesterol levels facilitate the formation and development of atherosclerotic plaques by inducing the generation of oxidized LDL. Increasing evidence suggests that a healthy gut microbiota engages in cholesterol reduction through various pathways, making it imperative to explore the precise mechanisms by which it achieves this. The conduction and findings of clinical trials will offer deeper insights into treating the cardiovascular diseases induced by high blood cholesterol. Furthermore, the interaction between natural functional ingredients and the cholesterol-lowering actions of the gut microbiota also represents a significant focus of research. This focus is poised to profoundly impact the development of novel therapeutic strategies for drug treatment. In summary, a deeper understanding of the mechanisms through which the gut microbiota reduces cholesterol is scientifically essential and opens a new avenue for the prevention of cardiovascular diseases. Prospective studies should further deepen the understanding of the connection between the gut microbiota and cholesterol reduction while exploring and identifying more effective prevention and treatment strategies.

Author Contributions

C.D. and J.P. contributed equally to this manuscript. Conceptualization, data curation, investigation, writing—original draft preparation, validation, and visualization: C.D. and J.P.; conceptualization, funding acquisition, writing—review and editing, supervision, and project administration: H.Z. and Z.-Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Project Number 32001709), the Guangdong Provincial Basic and Applied Basic Research Fund Regional Joint Fund (Project Number 22202107190000486), the National Key Research and Development Program of China (Project number 2022YFE0139500) and the Hong Kong Research Grants Council General Research Fund (Project Number CUHK 14102321 and 14104923).

Data Availability Statement

The data are openly available in a public repository. To collect the articles cited in this review, we conducted a comprehensive literature search in databases such as PubMed, Web of Science, and Google Scholar. We used keywords related to our research topic, like ‘cholesterol metabolism’ and ‘gut microbiota’, for the search. During the search process, we did not set any specific time constraints to ensure the inclusion of as many relevant studies as possible. Our only criterion for selection was language, considering only articles written in English. This approach was taken to ensure that we could accurately understand and analyze the content of the selected literature. No other specific inclusion or exclusion criteria were set, allowing us to comprehensively evaluate all relevant research, thereby ensuring the breadth and depth of the review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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] [PubMed]
  2. Mai, V.; Draganov, P.V. Recent Advances and Remaining Gaps in Our Knowledge of Associations between Gut Microbiota and Human Health. World J. Gastroenterol. 2009, 15, 81–85. [Google Scholar] [CrossRef] [PubMed]
  3. Santacruz, A.; Marcos, A.; Wärnberg, J.; Martí, A.; Martin-Matillas, M.; Campoy, C.; Moreno, L.A.; Veiga, O.; Redondo-Figuero, C.; Garagorri, J.M.; et al. Interplay between Weight Loss and Gut Microbiota Composition in Overweight Adolescents. Obesity 2009, 17, 1906–1915. [Google Scholar] [CrossRef] [PubMed]
  4. Ley, R.E.; Lozupone, C.A.; Hamady, M.; Knight, R.; Gordon, J.I. Worlds within Worlds: Evolution of the Vertebrate Gut Microbiota. Nat. Rev. Microbiol. 2008, 6, 776–788. [Google Scholar] [CrossRef] [PubMed]
  5. Gibson, G.R.; McCartney, A.L.; Rastall, R.A. Prebiotics and Resistance to Gastrointestinal Infections. Br. J. Nutr. 2005, 93 (Suppl. S1), S31–S34. [Google Scholar] [CrossRef] [PubMed]
  6. Campbell, J.M.; Fahey, G.C.; Wolf, B.W. Selected Indigestible Oligosaccharides Affect Large Bowel Mass, Cecal and Fecal Short-Chain Fatty Acids, pH and Microflora in Rats. J. Nutr. 1997, 127, 130–136. [Google Scholar] [CrossRef]
  7. Kurokawa, K.; Itoh, T.; Kuwahara, T.; Oshima, K.; Toh, H.; Toyoda, A.; Takami, H.; Morita, H.; Sharma, V.K.; Srivastava, T.P.; et al. Comparative Metagenomics Revealed Commonly Enriched Gene Sets in Human Gut Microbiomes. DNA Res. 2007, 14, 169–181. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Zeng, H.; Umar, S.; Rust, B.; Lazarova, D.; Bordonaro, M. Secondary Bile Acids and Short Chain Fatty Acids in the Colon: A Focus on Colonic Microbiome, Cell Proliferation, Inflammation, and Cancer. Int. J. Mol. Sci. 2019, 20, 1214. [Google Scholar] [CrossRef]
  10. Cook, S.I.; Sellin, J.H. Short Chain Fatty Acids in Health and Disease. Aliment Pharmacol. Ther. 1998, 12, 499–507. [Google Scholar] [CrossRef]
  11. Saito, T.; Hayashida, H.; Furugen, R. Comment on: Cani et al. Metabolic Endotoxemia Initiates Obesity and Insulin Resistance: Diabetes 56:1761–1772. Diabetes 2007, 56, e20. [Google Scholar] [CrossRef] [PubMed]
  12. Huang, Z.; Kraus, V.B. Does Lipopolysaccharide-Mediated Inflammation Have a Role in OA? Nat. Rev. Rheumatol. 2016, 12, 123–129. [Google Scholar] [CrossRef] [PubMed]
  13. Maher, J.J.; Leon, P.; Ryan, J.C. Beyond Insulin Resistance: Innate Immunity in Nonalcoholic Steatohepatitis. Hepatology 2008, 48, 670–678. [Google Scholar] [CrossRef] [PubMed]
  14. Patel, A.K.; Singhania, R.R.; Pandey, A.; Chincholkar, S.B. Probiotic Bile Salt Hydrolase: Current Developments and Perspectives. Appl. Biochem. Biotechnol. 2010, 162, 166–180. [Google Scholar] [CrossRef]
  15. Li, T.; Chiang, J.Y.L. Regulation of Bile Acid and Cholesterol Metabolism by PPARs. PPAR Res. 2009, 2009, 501739. [Google Scholar] [CrossRef] [PubMed]
  16. Durstine, J.L.; Grandjean, P.W.; Cox, C.A.; Thompson, P.D. Lipids, Lipoproteins, and Exercise. J. Cardiopulm. Rehabil. 2002, 22, 385–398. [Google Scholar] [CrossRef]
  17. Cox-York, K.A. The Effects of Moderate Exercise on Measures of Postprandial Lipemia; Colorado State University: Fort Collins, CO, USA, 2009. [Google Scholar]
  18. Björkbacka, H.; Kunjathoor, V.V.; Moore, K.J.; Koehn, S.; Ordija, C.M.; Lee, M.A.; Means, T.; Halmen, K.; Luster, A.D.; Golenbock, D.T.; et al. Reduced Atherosclerosis in MyD88-Null Mice Links Elevated Serum Cholesterol Levels to Activation of Innate Immunity Signaling Pathways. Nat. Med. 2004, 10, 416–421. [Google Scholar] [CrossRef] [PubMed]
  19. Law, M.R.; Thompson, S.G.; Wald, N.J. Assessing Possible Hazards of Reducing Serum Cholesterol. BMJ 1994, 308, 373–379. [Google Scholar] [CrossRef]
  20. Horwich, T.B.; Hamilton, M.A.; Maclellan, W.R.; Fonarow, G.C. Low Serum Total Cholesterol Is Associated with Marked Increase in Mortality in Advanced Heart Failure. J. Card. Fail. 2002, 8, 216–224. [Google Scholar] [CrossRef]
  21. Ridlon, J.M.; Kang, D.J.; Hylemon, P.B. Bile salt biotransformations by human intestinal bacteria. J. Lipid Res. 2006, 47, 241–259. [Google Scholar] [CrossRef]
  22. Bridgeman, S.; Woo, H.C.; Newsholme, P.; Mamotte, C. Butyrate Lowers Cellular Cholesterol through HDAC Inhibition and Impaired SREBP-2 Signalling. Int. J. Mol. Sci. 2022, 23, 15506. [Google Scholar] [CrossRef]
  23. Wilson, M.D.; Rudel, L.L. Review of Cholesterol Absorption with Emphasis on Dietary and Biliary Cholesterol. J. Lipid Res. 1994, 35, 943–955. [Google Scholar] [CrossRef]
  24. Li, L.; Batt, S.M.; Wannemuehler, M.; Dispirito, A.; Beitz, D.C. Effect of Feeding of a Cholesterol-Reducing Bacterium, Eubacterium Coprostanoligenes, to Germ-Free Mice. Lab. Anim. Sci. 1998, 48, 253–255. [Google Scholar] [PubMed]
  25. Yoshida, N.; Emoto, T.; Yamashita, T.; Watanabe, H.; Hayashi, T.; Tabata, T.; Hoshi, N.; Hatano, N.; Ozawa, G.; Sasaki, N.; et al. Bacteroides vulgatus and Bacteroides dorei Reduce Gut Microbial Lipopolysaccharide Production and Inhibit Atherosclerosis. Circulation 2018, 138, 2486–2498. [Google Scholar] [CrossRef] [PubMed]
  26. Ishimwe, N.; Daliri, E.B.; Lee, B.H.; Fang, F.; Du, G. The Perspective on Cholesterol-Lowering Mechanisms of Probiotics. Mol. Nutr. Food Res. 2015, 59, 94–105. [Google Scholar] [CrossRef] [PubMed]
  27. Peng, M.; Tabashsum, Z.; Anderson, M.; Truong, A.; Houser, A.K.; Padilla, J.; Akmel, A.; Bhatti, J.; Rahaman, S.O.; Biswas, D. Effectiveness of Probiotics, Prebiotics, and Prebiotic-like Components in Common Functional Foods. Compr. Rev. Food Sci. Food Saf. 2020, 19, 1908–1933. [Google Scholar] [CrossRef] [PubMed]
  28. Sudheer, S.; Gangwar, P.; Usmani, Z.; Sharma, M.; Sharma, V.K.; Sana, S.S.; Almeida, F.; Dubey, N.K.; Singh, D.P.; Dilbaghi, N.; et al. Shaping the Gut Microbiota by Bioactive Phytochemicals: An Emerging Approach for the Prevention and Treatment of Human Diseases. Biochimie 2022, 193, 38–63. [Google Scholar] [CrossRef]
  29. Colomer, R.; Sarrats, A.; Lupu, R.; Puig, T. Natural Polyphenols and Their Synthetic Analogs as Emerging Anticancer Agents. Curr. Drug Targets 2017, 18, 147–159. [Google Scholar] [CrossRef] [PubMed]
  30. Mozaffarian, D.; Micha, R.; Wallace, S. Effects on Coronary Heart Disease of Increasing Polyunsaturated Fat in Place of Saturated Fat: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. PLoS Med. 2010, 7, e1000252. [Google Scholar] [CrossRef] [PubMed]
  31. Xie, M.; Chen, G.; Wan, P.; Dai, Z.; Zeng, X.; Sun, Y. Effects of Dicaffeoylquinic Acids from Ilex kudingcha on Lipid Metabolism and Intestinal Microbiota in High-Fat-Diet-Fed Mice. J. Agric. Food Chem. 2019, 67, 171–183. [Google Scholar] [CrossRef]
  32. Francini, A.; Sebastiani, L. Phenolic Compounds in Apple (Malus x Domestica Borkh.): Compounds Characterization and Stability during Postharvest and after Processing. Antioxidants 2013, 2, 181–193. [Google Scholar] [CrossRef] [PubMed]
  33. Manson, J.E.; Tosteson, H.; Ridker, P.M.; Satterfield, S.; Hebert, P.; O’Connor, G.T.; Buring, J.E.; Hennekens, C.H. The Primary Prevention of Myocardial Infarction. N. Engl. J. Med. 1992, 326, 1406–1416. [Google Scholar] [CrossRef] [PubMed]
  34. Kunnen, S.; Van Eck, M. Lecithin:Cholesterol Acyltransferase: Old Friend or Foe in Atherosclerosis? J. Lipid Res. 2012, 53, 1783–1799. [Google Scholar] [CrossRef] [PubMed]
  35. Lecerf, J.-M.; de Lorgeril, M. Dietary Cholesterol: From Physiology to Cardiovascular Risk. Br. J. Nutr. 2011, 106, 6–14. [Google Scholar] [CrossRef]
  36. Trapani, L.; Segatto, M.; Pallottini, V. Regulation and Deregulation of Cholesterol Homeostasis: The Liver as a Metabolic “Power Station”. World J. Hepatol. 2012, 4, 184–190. [Google Scholar] [CrossRef] [PubMed]
  37. Robichon, C.; Dugail, I. De Novo Cholesterol Synthesis at the Crossroads of Adaptive Response to Extracellular Stress through SREBP. Biochimie 2007, 89, 260–264. [Google Scholar] [CrossRef] [PubMed]
  38. Gesto, D.S.; Pereira, C.M.S.; Cerqueira, N.M.F.S.; Sousa, S.F. An Atomic-Level Perspective of HMG-CoA-Reductase: The Target Enzyme to Treat Hypercholesterolemia. Molecules 2020, 25, 3891. [Google Scholar] [CrossRef]
  39. Jo, Y.; Debose-Boyd, R.A. Control of Cholesterol Synthesis through Regulated ER-Associated Degradation of HMG CoA Reductase. Crit. Rev. Biochem. Mol. Biol. 2010, 45, 185–198. [Google Scholar] [CrossRef]
  40. Goldstein, J.L.; Brown, M.S. Progress in Understanding the LDL Receptor and HMG-CoA Reductase, Two Membrane Proteins That Regulate the Plasma Cholesterol. J. Lipid Res. 1984, 25, 1450–1461. [Google Scholar] [CrossRef]
  41. Olivier, L.M.; Chambliss, K.L.; Gibson, K.M.; Krisans, S.K. Characterization of Phosphomevalonate Kinase: Chromosomal Localization, Regulation, and Subcellular Targeting. J. Lipid Res. 1999, 40, 672–679. [Google Scholar] [CrossRef]
  42. Kovacs, W.J.; Olivier, L.M.; Krisans, S.K. Central Role of Peroxisomes in Isoprenoid Biosynthesis. Prog. Lipid Res. 2002, 41, 369–391. [Google Scholar] [CrossRef] [PubMed]
  43. Brusselmans, K.; Timmermans, L.; Van de Sande, T.; Van Veldhoven, P.P.; Guan, G.; Shechter, I.; Claessens, F.; Verhoeven, G.; Swinnen, J.V. Squalene Synthase, a Determinant of Raft-Associated Cholesterol and Modulator of Cancer Cell Proliferation. J. Biol. Chem. 2007, 282, 18777–18785. [Google Scholar] [CrossRef] [PubMed]
  44. Dominiczak, M.H.; Wallace, A.M. Medical Biochemistry: Biosynthesis of Cholesterol and Steroids; Mosby Elseviers: Philadelphia, PA, USA, 2009. [Google Scholar]
  45. Astruc, M.; Tabacik, C.; Descomps, B.; de Paulet, A.C. Squalene Epoxidase and Oxidosqualene Lanosterol-Cyclase Activities in Cholesterogenic and Non-Cholesterogenic Tissues. Biochim. Biophys. Acta 1977, 487, 204–211. [Google Scholar] [CrossRef] [PubMed]
  46. Cabrera-Vivas, B.M.; Ramírez, J.C.; Martínez-Aguilera, L.M.R.; Kubfi-Garfias, C. Theoretical Assessment of the Mechanisms Involved in the Cholesterol Biosynthesis from Lanosterol. Theochem J. Mol. Struct. 2002, 584, 5–14. [Google Scholar] [CrossRef]
  47. Wang, D.Q.-H. Regulation of Intestinal Cholesterol Absorption. Annu. Rev. Physiol. 2007, 69, 221–248. [Google Scholar] [CrossRef]
  48. Carey, M.C.; Small, D.M.; Bliss, C.M. Lipid Digestion and Absorption. Annu. Rev. Physiol. 1983, 45, 651–677. [Google Scholar] [CrossRef]
  49. Chen, Z.-Y.; Ma, K.Y.; Liang, Y.; Peng, C.; Zuo, Y. Role and Classification of Cholesterol-Lowering Functional Foods. J. Funct. Food 2011, 3, 61–69. [Google Scholar] [CrossRef]
  50. Jia, L.; Betters, J.L.; Yu, L. Niemann-Pick C1-like 1 (NPC1L1) Protein in Intestinal and Hepatic Cholesterol Transport. Annu. Rev. Physiol. 2011, 73, 239–259. [Google Scholar] [CrossRef]
  51. Alphonse, P.A.S.; Jones, P.J.H. Revisiting Human Cholesterol Synthesis and Absorption: The Reciprocity Paradigm and Its Key Regulators. Lipids 2016, 51, 519–536. [Google Scholar] [CrossRef]
  52. Trautwein, E.A.; Duchateau, G.; Lin, Y.G.; Mel’nikov, S.M.; Molhuizen, H.O.F.; Ntanios, F.Y. Proposed Mechanisms of Cholesterol-Lowering Action of Plant Sterols. Eur. J. Lipid Sci. Technol. 2003, 105, 171–185. [Google Scholar] [CrossRef]
  53. Chang, T.Y.; Chang, C.C.; Lin, S.; Yu, C.; Li, B.L.; Miyazaki, A. Roles of Acyl-Coenzyme A:Cholesterol Acyltransferase-1 and -2. Curr. Opin. Lipidol. 2001, 12, 289–296. [Google Scholar] [CrossRef] [PubMed]
  54. Berriot-Varoqueaux, N.; Aggerbeck, L.P.; Samson-Bouma, M.; Wetterau, J.R. The Role of the Microsomal Triglygeride Transfer Protein in Abetalipoproteinemia. Annu. Rev. Nutr. 2000, 20, 663–697. [Google Scholar] [CrossRef] [PubMed]
  55. Feingold, K.R. Introduction to Lipids and Lipoproteins; MDText.com, Inc.: South Dartmouth, MA, USA, 2015. [Google Scholar]
  56. Graf, G.A.; Li, W.-P.; Gerard, R.D.; Gelissen, I.; White, A.; Cohen, J.C.; Hobbs, H.H. Coexpression of ATP-Binding Cassette Proteins ABCG5 and ABCG8 Permits Their Transport to the Apical Surface. J. Clin. Investig. 2002, 110, 659–669. [Google Scholar] [CrossRef] [PubMed]
  57. Mayes, P.A. Cholesterol Synthesis, Transport, and Excretion. In Harper’s Biochemistry; Murray, R.K., Mayes, P.A., Granner, D.K., Rodwell, V.W., Eds.; Appleton & Lange: East Norwalk, CT, USA, 1990; pp. 253–255. [Google Scholar]
  58. Ramírez-Pérez, O.; Cruz-Ramón, V.; Chinchilla-López, P.; Méndez-Sánchez, N. The Role of the Gut Microbiota in Bile Acid Metabolism. Ann. Hepatol. 2017, 16, s15–s20. [Google Scholar] [CrossRef]
  59. Xie, G.; Jiang, R.; Wang, X.; Liu, P.; Zhao, A.; Wu, Y.; Huang, F.; Liu, Z.; Rajani, C.; Zheng, X.; et al. Conjugated Secondary 12α-Hydroxylated Bile Acids Promote Liver Fibrogenesis. EBioMedicine 2021, 66, 103290. [Google Scholar] [CrossRef]
  60. Ridlon, J.M.; Devendran, S.; Alves, J.M.; Doden, H.; Wolf, P.G.; Pereira, G.V.; Ly, L.; Volland, A.; Takei, H.; Nittono, H.; et al. The “in Vivo Lifestyle” of Bile Acid 7α-Dehydroxylating Bacteria: Comparative Genomics, Metatranscriptomic, and Bile Acid Metabolomics Analysis of a Defined Microbial Community in Gnotobiotic Mice. Gut Microbes 2020, 11, 381–404. [Google Scholar] [CrossRef]
  61. Winston, J.A.; Theriot, C.M. Diversification of Host Bile Acids by Members of the Gut Microbiota. Gut Microbes 2020, 11, 158–171. [Google Scholar] [CrossRef]
  62. Kühn, T.; Stepien, M.; López-Nogueroles, M.; Damms-Machado, A.; Sookthai, D.; Johnson, T.; Roca, M.; Hüsing, A.; Maldonado, S.G.; Cross, A.J.; et al. Prediagnostic Plasma Bile Acid Levels and Colon Cancer Risk: A Prospective Study. J. Natl. Cancer Inst. 2020, 112, 516–524. [Google Scholar] [CrossRef]
  63. Hofmann, A.F. Detoxification of Lithocholic Acid, a Toxic Bile Acid: Relevance to Drug Hepatotoxicity. Drug Metab. Rev. 2004, 36, 703–722. [Google Scholar] [CrossRef]
  64. Gérard, P. Metabolism of Cholesterol and Bile Acids by the Gut Microbiota. Pathogens 2013, 3, 14–24. [Google Scholar] [CrossRef]
  65. Dowling, R.H.; Mack, E.; Small, D.M. Effects of Controlled Interruption of the Enterohepatic Circulation of Bile Salts by Biliary Diversion and by Ileal Resection on Bile Salt Secretion, Synthesis, and Pool Size in the Rhesus Monkey. J. Clin. Investig. 1970, 49, 232–242. [Google Scholar] [CrossRef]
  66. Chiang, J.Y.L.; Ferrell, J.M. Discovery of Farnesoid X Receptor and Its Role in Bile Acid Metabolism. Mol. Cell Endocrinol. 2022, 548, 111618. [Google Scholar] [CrossRef]
  67. Rizzo, G.; Renga, B.; Mencarelli, A.; Pellicciari, R.; Fiorucci, S. Role of FXR in Regulating Bile Acid Homeostasis and Relevance for Human Diseases. Curr. Drug Targets Immune Endocr. Metabol. Disord. 2005, 5, 289–303. [Google Scholar] [CrossRef] [PubMed]
  68. Reeskamp, L.F.; Meessen, E.C.E.; Groen, A.K. Transintestinal Cholesterol Excretion in Humans. Curr. Opin. Lipidol 2018, 29, 10–17. [Google Scholar] [CrossRef] [PubMed]
  69. Yu, X.-H.; Qian, K.; Jiang, N.; Zheng, X.-L.; Cayabyab, F.S.; Tang, C.-K. ABCG5/ABCG8 in Cholesterol Excretion and Atherosclerosis. Clin. Chim. Acta 2014, 428, 82–88. [Google Scholar] [CrossRef] [PubMed]
  70. Brown, J.M.; Yu, L. Opposing Gatekeepers of Apical Sterol Transport: Niemann-Pick C1-Like 1 (NPC1L1) and ATP-Binding Cassette Transporters G5 and G8 (ABCG5/ABCG8). Immunol. Endocr. Metab. Agents Med. Chem. 2009, 9, 18–29. [Google Scholar] [CrossRef] [PubMed]
  71. Yu, L.; Li-Hawkins, J.; Hammer, R.E.; Berge, K.E.; Horton, J.D.; Cohen, J.C.; Hobbs, H.H. Overexpression of ABCG5 and ABCG8 Promotes Biliary Cholesterol Secretion and Reduces Fractional Absorption of Dietary Cholesterol. J. Clin. Investig. 2002, 110, 671–680. [Google Scholar] [CrossRef]
  72. Miremadi, F.; Ayyash, M.; Sherkat, F.; Stojanovska, L. Cholesterol Reduction Mechanisms and Fatty Acid Composition of Cellular Membranes of Probiotic Lactobacilli and Bifidobacteria. J. Funct. Food. 2014, 9, 295–305. [Google Scholar] [CrossRef]
  73. Blaak, E.E.; Canfora, E.E.; Theis, S.; Frost, G.; Groen, A.K.; Mithieux, G.; Nauta, A.; Scott, K.; Stahl, B.; van Harsselaar, J.; et al. Short Chain Fatty Acids in Human Gut and Metabolic Health. Benef. Microbes 2020, 11, 411–455. [Google Scholar] [CrossRef]
  74. Ridlon, J.M.; Harris, S.C.; Bhowmik, S.; Kang, D.-J.; Hylemon, P.B. Consequences of Bile Salt Biotransformations by Intestinal Bacteria. Gut Microbes 2016, 7, 22–39. [Google Scholar] [CrossRef]
  75. Cao, K.; Zhang, K.; Ma, M.; Ma, J.; Tian, J.; Jin, Y. Lactobacillus Mediates the Expression of NPC1L1, CYP7A1, and ABCG5 Genes to Regulate Cholesterol. Food Sci. Nutr. 2021, 9, 6882–6891. [Google Scholar] [CrossRef] [PubMed]
  76. Hu, H.; Shao, W.; Liu, Q.; Liu, N.; Wang, Q.; Xu, J.; Zhang, X.; Weng, Z.; Lu, Q.; Jiao, L.; et al. Gut Microbiota Promotes Cholesterol Gallstone Formation by Modulating Bile Acid Composition and Biliary Cholesterol Secretion. Nat. Commun. 2022, 13, 252. [Google Scholar] [CrossRef] [PubMed]
  77. Tsai, M.-J.; Chang, W.-A.; Liao, S.-H.; Chang, K.-F.; Sheu, C.-C.; Kuo, P.-L. The Effects of Epigallocatechin Gallate (EGCG) on Pulmonary Fibroblasts of Idiopathic Pulmonary Fibrosis (IPF)A Next-Generation Sequencing and Bioinformatic Approach. Int. J. Mol. Sci. 2019, 20, 1958. [Google Scholar] [CrossRef]
  78. Bortolini, O.; Medici, A.; Poli, S. Biotransformations on Steroid Nucleus of Bile Acids. Steroids 1997, 62, 564–577. [Google Scholar] [CrossRef] [PubMed]
  79. Russell, D.W. The Enzymes, Regulation, and Genetics of Bile Acid Synthesis. Annu. Rev. Biochem. 2003, 72, 137–174. [Google Scholar] [CrossRef] [PubMed]
  80. Moini, J. Epidemiology of Diet and Diabetes Mellitus-ScienceDirect. Epidemiol. Diabetes 2019, 57–73. [Google Scholar]
  81. Guzior, D.V.; Quinn, R.A. Review: Microbial Transformations of Human Bile Acids. Microbiome 2021, 9, 140. [Google Scholar] [CrossRef]
  82. Song, Z.; Cai, Y.; Lao, X.; Wang, X.; Lin, X.; Cui, Y.; Kalavagunta, P.K.; Liao, J.; Jin, L.; Shang, J.; et al. Taxonomic Profiling and Populational Patterns of Bacterial Bile Salt Hydrolase (BSH) Genes Based on Worldwide Human Gut Microbiome. Microbiome 2019, 7, 9. [Google Scholar] [CrossRef]
  83. Kim, G.-B.; Yi, S.-H.; Lee, B.H. Purification and Characterization of Three Different Types of Bile Salt Hydrolases from Bifidobacterium Strains. J. Dairy Sci. 2004, 87, 258–266. [Google Scholar] [CrossRef]
  84. Elkins, C.A.; Moser, S.A.; Savage, D.C. Genes Encoding Bile Salt Hydrolases and Conjugated Bile Salt Transporters in Lactobacillus johnsonii 100-100 and Other Lactobacillus Species. Microbiology 2001, 147, 3403–3412. [Google Scholar] [CrossRef]
  85. Corzo, G.; Gilliland, S.E. Bile Salt Hydrolase Activity of Three Strains of Lactobacillus acidophilus. J. Dairy Sci. 1999, 82, 472–480. [Google Scholar] [CrossRef] [PubMed]
  86. Coleman, J.P.; Hudson, L.L. Cloning and Characterization of a Conjugated Bile Acid Hydrolase Gene from Clostridium perfringens. Appl. Environ. Microbiol. 1995, 61, 2514–2520. [Google Scholar] [CrossRef] [PubMed]
  87. Wijaya, A.; Hermann, A.; Abriouel, H.; Specht, I.; Yousif, N.M.K.; Holzapfel, W.H.; Franz, C.M.A.P. Cloning of the Bile Salt Hydrolase (Bsh) Gene from Enterococcus faecium FAIR-E 345 and Chromosomal Location of Bsh Genes in Food Enterococci. J. Food Prot. 2004, 67, 2772–2778. [Google Scholar] [CrossRef]
  88. Dussurget, O.; Cabanes, D.; Dehoux, P.; Lecuit, M.; Buchrieser, C.; Glaser, P.; Cossart, P.; European Listeria Genome Consortium. Listeria monocytogenes Bile Salt Hydrolase Is a PrfA-Regulated Virulence Factor Involved in the Intestinal and Hepatic Phases of Listeriosis. Mol. Microbiol. 2002, 45, 1095–1106. [Google Scholar] [CrossRef] [PubMed]
  89. Dean, M.; Cervellati, C.; Casanova, E.; Squerzanti, M.; Lanzara, V.; Medici, A.; Polverino De Laureto, P.; Bergamini, C.M. Characterization of Cholylglycine Hydrolase from a Bile-Adapted Strain of Xanthomonas maltophilia and Its Application for Quantitative Hydrolysis of Conjugated Bile Salts. Appl. Environ. Microbiol. 2002, 68, 3126–3128. [Google Scholar] [CrossRef] [PubMed]
  90. Kawamoto, K.; Horibe, I.; Uchida, K. Purification and Characterization of a New Hydrolase for Conjugated Bile Acids, Chenodeoxycholyltaurine Hydrolase, from Bacteroides vulgatus. J. Biochem. 1989, 106, 1049–1053. [Google Scholar] [CrossRef] [PubMed]
  91. Delpino, M.V.; Marchesini, M.I.; Estein, S.M.; Comerci, D.J.; Cassataro, J.; Fossati, C.A.; Baldi, P.C. A Bile Salt Hydrolase of Brucella abortus Contributes to the Establishment of a Successful Infection through the Oral Route in Mice. Infect. Immun. 2007, 75, 299–305. [Google Scholar] [CrossRef]
  92. Percy-Robb, I.W.; Collee, J.G. Bile Acids: A pH Dependent Antibacterial System in the Gut? Br. Med. J. 1972, 3, 813–815. [Google Scholar] [CrossRef]
  93. Ridlon, J.M.; Hylemon, P.B. Identification and Characterization of Two Bile Acid Coenzyme A Transferases from Clostridium scindens, a Bile Acid 7α-Dehydroxylating Intestinal Bacterium. J. Lipid Res. 2012, 53, 66–76. [Google Scholar] [CrossRef]
  94. Mallonee, D.H.; Hylemon, P.B. Sequencing and Expression of a Gene Encoding a Bile Acid Transporter from Eubacterium sp. Strain VPI 12708. J. Bacteriol. 1996, 178, 7053–7058. [Google Scholar] [CrossRef]
  95. Heinken, A.; Ravcheev, D.A.; Baldini, F.; Heirendt, L.; Fleming, R.M.T.; Thiele, I. Systematic Assessment of Secondary Bile Acid Metabolism in Gut Microbes Reveals Distinct Metabolic Capabilities in Inflammatory Bowel Disease. Microbiome 2019, 7, 75. [Google Scholar] [CrossRef]
  96. Harris, S.C.; Devendran, S.; Méndez-García, C.; Mythen, S.M.; Wright, C.L.; Fields, C.J.; Hernandez, A.G.; Cann, I.; Hylemon, P.B.; Ridlon, J.M. Bile Acid Oxidation by Eggerthella Lenta Strains C592 and DSM 2243T. Gut Microbes 2018, 9, 523–539. [Google Scholar] [PubMed]
  97. Funabashi, M.; Grove, T.L.; Wang, M.; Varma, Y.; McFadden, M.E.; Brown, L.C.; Guo, C.; Higginbottom, S.; Almo, S.C.; Fischbach, M.A. A Metabolic Pathway for Bile Acid Dehydroxylation by the Gut Microbiome. Nature 2020, 582, 566–570. [Google Scholar] [CrossRef] [PubMed]
  98. Bhowmik, S.; Jones, D.H.; Chiu, H.-P.; Park, I.-H.; Chiu, H.-J.; Axelrod, H.L.; Farr, C.L.; Tien, H.J.; Agarwalla, S.; Lesley, S.A. Structural and Functional Characterization of BaiA, an Enzyme Involved in Secondary Bile Acid Synthesis in Human Gut Microbe. Proteins 2014, 82, 216–229. [Google Scholar] [CrossRef] [PubMed]
  99. Kang, D.-J.; Ridlon, J.M.; Moore, D.R.; Barnes, S.; Hylemon, P.B. Clostridium scindens baiCD and baiH Genes Encode Stereo-Specific 7alpha/7beta-Hydroxy-3-Oxo-Delta4-Cholenoic Acid Oxidoreductases. Biochim. Biophys. Acta 2008, 1781, 16–25. [Google Scholar] [CrossRef] [PubMed]
  100. Doden, H.; Sallam, L.A.; Devendran, S.; Ly, L.; Doden, G.; Daniel, S.L.; Alves, J.M.P.; Ridlon, J.M. Metabolism of Oxo-Bile Acids and Characterization of Recombinant 12α-Hydroxysteroid Dehydrogenases from Bile Acid 7α-Dehydroxylating Human Gut Bacteria. Appl. Environ. Microbiol. 2018, 84, e00235-18. [Google Scholar] [CrossRef]
  101. Hirano, S.; Masuda, N. Epimerization of the 7-Hydroxy Group of Bile Acids by the Combination of Two Kinds of Microorganisms with 7 Alpha- and 7 Beta-Hydroxysteroid Dehydrogenase Activity, Respectively. J. Lipid Res. 1981, 22, 1060–1068. [Google Scholar] [CrossRef] [PubMed]
  102. Garcia, C.J.; Kosek, V.; Beltrán, D.; Tomás-Barberán, F.A.; Hajslova, J. Production of New Microbially Conjugated Bile Acids by Human Gut Microbiota. Biomolecules 2022, 12, 687. [Google Scholar] [CrossRef]
  103. Franco, P.; Porru, E.; Fiori, J.; Gioiello, A.; Cerra, B.; Roda, G.; Caliceti, C.; Simoni, P.; Roda, A. Identification and Quantification of Oxo-Bile Acids in Human Faeces with Liquid Chromatography-Mass Spectrometry: A Potent Tool for Human Gut Acidic Sterolbiome Studies. J. Chromatogr. A 2019, 1585, 70–81. [Google Scholar] [CrossRef]
  104. Lepercq, P.; Gérard, P.; Béguet, F.; Raibaud, P.; Grill, J.-P.; Relano, P.; Cayuela, C.; Juste, C. Epimerization of Chenodeoxycholic Acid to Ursodeoxycholic Acid by Clostridium baratii Isolated from Human Feces. FEMS Microbiol. Lett. 2004, 235, 65–72. [Google Scholar] [CrossRef]
  105. Martin-Gallausiaux, C.; Marinelli, L.; Blottière, H.M.; Larraufie, P.; Lapaque, N. SCFA: Mechanisms and Functional Importance in the Gut. Proc. Nutr. Soc. 2021, 80, 37–49. [Google Scholar] [CrossRef] [PubMed]
  106. Du, Y.; Li, X.; Su, C.; Xi, M.; Zhang, X.; Jiang, Z.; Wang, L.; Hong, B. Butyrate Protects against High-Fat Diet-Induced Atherosclerosis via up-Regulating ABCA1 Expression in Apolipoprotein E-Deficiency Mice. Br. J. Pharmacol. 2020, 177, 1754–1772. [Google Scholar] [CrossRef] [PubMed]
  107. Hara, H.; Haga, S.; Aoyama, Y.; Kiriyama, S. Short-Chain Fatty Acids Suppress Cholesterol Synthesis in Rat Liver and Intestine. J. Nutr. 1999, 129, 942–948. [Google Scholar] [CrossRef] [PubMed]
  108. Johansen, M.E.; Green, L.A.; Sen, A.; Kircher, S.; Richardson, C.R. Cardiovascular Risk and Statin Use in the United States. Ann. Fam. Med. 2014, 12, 215–223. [Google Scholar] [CrossRef] [PubMed]
  109. Ikonen, E. Mechanisms for Cellular Cholesterol Transport: Defects and Human Disease. Physiol. Rev. 2006, 86, 1237–1261. [Google Scholar] [CrossRef] [PubMed]
  110. Le Lay, S.; Krief, S.; Farnier, C.; Lefrère, I.; Le Liepvre, X.; Bazin, R.; Ferré, P.; Dugail, I. Cholesterol, a Cell Size-Dependent Signal That Regulates Glucose Metabolism and Gene Expression in Adipocytes. J. Biol. Chem. 2001, 276, 16904–16910. [Google Scholar] [CrossRef] [PubMed]
  111. Ebihara, K.; Miyada, T.; Nakajima, A. Hypocholesterolemic Effect of Cecally Infused Propionic Acid in Rats Fed a Cholesterol-Free, Casein Diet. Nutr. Res. 1993, 13, 209–217. [Google Scholar] [CrossRef]
  112. Fushimi, T.; Suruga, K.; Oshima, Y.; Fukiharu, M.; Tsukamoto, Y.; Goda, T. Dietary Acetic Acid Reduces Serum Cholesterol and Triacylglycerols in Rats Fed a Cholesterol-Rich Diet. Br. J. Nutr. 2006, 95, 916–924. [Google Scholar] [CrossRef]
  113. Zhao, Y.; Liu, J.; Hao, W.; Zhu, H.; Liang, N.; He, Z.; Ma, K.Y.; Chen, Z.-Y. Structure-Specific Effects of Short-Chain Fatty Acids on Plasma Cholesterol Concentration in Male Syrian Hamsters. J. Agric. Food Chem. 2017, 65, 10984–10992. [Google Scholar] [CrossRef]
  114. Pereira, D.I.A.; Gibson, G.R. Cholesterol Assimilation by Lactic Acid Bacteria and Bifidobacteria Isolated from the Human Gut. Appl. Environ. Microbiol. 2002, 68, 4689–4693. [Google Scholar] [CrossRef]
  115. Klaver, F.A.; van der Meer, R. The Assumed Assimilation of Cholesterol by Lactobacilli and Bifidobacterium bifidum Is Due to Their Bile Salt-Deconjugating Activity. Appl. Environ. Microbiol. 1993, 59, 1120–1124. [Google Scholar] [CrossRef] [PubMed]
  116. Pan, D.D.; Zeng, X.Q.; Yan, Y.T. Characterisation of Lactobacillus fermentum SM-7 Isolated from Koumiss, a Potential Probiotic Bacterium with Cholesterol-Lowering Effects. J. Sci. Food Agric. 2011, 91, 512–518. [Google Scholar] [CrossRef] [PubMed]
  117. Srivastava, R.A.K.; Pinkosky, S.L.; Filippov, S.; Hanselman, J.C.; Cramer, C.T.; Newton, R.S. AMP-Activated Protein Kinase: An Emerging Drug Target to Regulate Imbalances in Lipid and Carbohydrate Metabolism to Treat Cardio-Metabolic Diseases. J. Lipid Res. 2012, 53, 2490–2514. [Google Scholar] [CrossRef] [PubMed]
  118. Lew, L.-C.; Choi, S.-B.; Khoo, B.-Y.; Sreenivasan, S.; Ong, K.-L.; Liong, M.-T. Lactobacillus plantarum DR7 Reduces Cholesterol via Phosphorylation of AMPK That Down-Regulated the mRNA Expression of HMG-CoA Reductase. Korean J. Food Sci. Anim. Resour. 2018, 38, 350–361. [Google Scholar]
  119. Betters, J.L.; Yu, L. NPC1L1 and Cholesterol Transport. FEBS Lett. 2010, 584, 2740–2747. [Google Scholar] [CrossRef]
  120. Temel, R.E.; Tang, W.; Ma, Y.; Rudel, L.L.; Willingham, M.C.; Ioannou, Y.A.; Davies, J.P.; Nilsson, L.-M.; Yu, L. Hepatic Niemann-Pick C1-like 1 Regulates Biliary Cholesterol Concentration and Is a Target of Ezetimibe. J. Clin. Investig. 2007, 117, 1968–1978. [Google Scholar] [CrossRef]
  121. Repa, J.J.; Berge, K.E.; Pomajzl, C.; Richardson, J.A.; Hobbs, H.; Mangelsdorf, D.J. Regulation of ATP-Binding Cassette Sterol Transporters ABCG5 and ABCG8 by the Liver X Receptors Alpha and Beta. J. Biol. Chem. 2002, 277, 18793–18800. [Google Scholar] [CrossRef]
  122. Le Lay, S.; Lefrère, I.; Trautwein, C.; Dugail, I.; Krief, S. Insulin and Sterol-Regulatory Element-Binding Protein-1c (SREBP-1C) Regulation of Gene Expression in 3T3-L1 Adipocytes. Identification of CCAAT/Enhancer-Binding Protein Beta as an SREBP-1C Target. J. Biol. Chem. 2002, 277, 35625–35634. [Google Scholar] [CrossRef]
  123. Horton, J.D.; Goldstein, J.L.; Brown, M.S. SREBPs: Activators of the Complete Program of Cholesterol and Fatty Acid Synthesis in the Liver. J. Clin. Investig. 2002, 109, 1125–1131. [Google Scholar] [CrossRef]
  124. Li, C.; Nie, S.-P.; Ding, Q.; Zhu, K.-X.; Wang, Z.-J.; Xiong, T.; Gong, J.; Xie, M.-Y. Cholesterol-Lowering Effect of Lactobacillus plantarum NCU116 in a Hyperlipidaemic Rat Model. J. Funct. Food. 2014, 8, 340–347. [Google Scholar] [CrossRef]
  125. Segawa, S.; Wakita, Y.; Hirata, H.; Watari, J. Oral Administration of Heat-Killed Lactobacillus brevis SBC8803 Ameliorates Alcoholic Liver Disease in Ethanol-Containing Diet-Fed C57BL/6N Mice. Int. J. Food Microbiol. 2008, 128, 371–377. [Google Scholar] [CrossRef] [PubMed]
  126. Norlin, M.; Wikvall, K. Enzymes in the Conversion of Cholesterol into Bile Acids. Curr. Mol. Med. 2007, 7, 199–218. [Google Scholar] [CrossRef] [PubMed]
  127. 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] [PubMed]
  128. Qu, T.; Yang, L.; Wang, Y.; Jiang, B.; Shen, M.; Ren, D. Reduction of Serum Cholesterol and Its Mechanism by Lactobacillus plantarum H6 Screened from Local Fermented Food Products. Food Funct. 2020, 11, 1397–1409. [Google Scholar] [CrossRef] [PubMed]
  129. Kim, B.; Park, K.-Y.; Ji, Y.; Park, S.; Holzapfel, W.; Hyun, C.-K. Protective Effects of Lactobacillus rhamnosus GG against Dyslipidemia in High-Fat Diet-Induced Obese Mice. Biochem. Biophys. Res. Commun. 2016, 473, 530–536. [Google Scholar] [CrossRef]
  130. Freier, T.A.; Beitz, D.C.; Li, L.; Hartman, P.A. Characterization of Eubacterium coprostanoligenes sp. nov., a Cholesterol-Reducing Anaerobe. Int. J. Syst. Evol. Microbiol. 1994, 44, 137–142. [Google Scholar] [CrossRef]
  131. Li, L.; Buhman, K.K.; Hartman, P.A.; Beitz, D.C. Hypocholesterolemic Effect of Eubacterium Coprostanoligenes ATCC 51222 in Rabbits. Lett. Appl. Microbiol. 1995, 20, 137–140. [Google Scholar] [CrossRef]
  132. Parmentier, G.; Eyssen, H. Mechanism of Biohydrogenation of Cholesterol to Coprostanol by Eubacterium ATCC 21408. Biochim. Biophys. Acta 1974, 348, 279–284. [Google Scholar] [CrossRef]
  133. Ren, D.W.; Li, L.; Schwabacher, A.W.; Young, J.W.; Beitz, D.C. Mechanism of Cholesterol Reduction to Coprostanol by Eubacterium Coprostanoligenes ATCC 51222. Steroids 1996, 61, 33–40. [Google Scholar] [CrossRef]
  134. Eyssen, H.J.; Parmentier, G.G.; Compernolle, F.C.; De Pauw, G.; Piessens-Denef, M. Biohydrogenation of Sterols by Eubacterium ATCC 21,408--Nova Species. Eur. J. Biochem. 1973, 36, 411–421. [Google Scholar] [CrossRef]
  135. Mott, G.E.; Brinkley, A.W.; Mersinger, C.L. Biochemical Characterization of Cholesterol-Reducing Eubacterium. Appl. Environ. Microbiol. 1980, 40, 1017–1022. [Google Scholar] [CrossRef] [PubMed]
  136. Cuevas-Tena, M.; Alegria, A.; Lagarda, M.J. Relationship Between Dietary Sterols and Gut Microbiota: A Review. Eur. J. Lipid Sci. Technol. 2018, 120, 1800054. [Google Scholar] [CrossRef]
  137. Rosenfeld, R.S.; Gallagher, T.F. Further Studies of the Biotransformation of Cholesterol to Coprostanol. Steroids 1964, 4, 515–520. [Google Scholar] [CrossRef]
  138. Björkhem, I.; Gustafsson, J.A. Mechanism of Microbial Transformation of Cholesterol into Coprostanol. Eur. J. Biochem. 1971, 21, 428–432. [Google Scholar] [CrossRef] [PubMed]
  139. Park, Y.H.; Kim, J.G.; Shin, Y.W.; Kim, H.S.; Kim, Y.-J.; Chun, T.; Kim, S.H.; Whang, K.Y. Effects of Lactobacillus acidophilus 43121 and a Mixture of Lactobacillus casei and Bifidobacterium longum on the Serum Cholesterol Level and Fecal Sterol Excretion in Hypercholesterolemia-Induced Pigs. Biosci. Biotechnol. Biochem. 2008, 72, 595–600. [Google Scholar] [CrossRef]
  140. Snog-Kjaer, A.; Prange, I.; Dam, H. Conversion of Cholesterol into Coprosterol by Bacteria in Vitro. J. Gen. Microbiol. 1956, 14, 256–260. [Google Scholar] [CrossRef]
  141. Crowther, J.S.; Drasar, B.S.; Goddard, P.; Hill, M.J.; Johnson, K. The Effect of a Chemically Defined Diet on the Faecal Flora and Faecal Steroid Concentration. Gut 1973, 14, 790–793. [Google Scholar] [CrossRef] [PubMed]
  142. Heumann, D.; Roger, T. Initial Responses to Endotoxins and Gram-Negative Bacteria. Clin. Chim. Acta 2002, 323, 59–72. [Google Scholar] [CrossRef]
  143. Feingold, K.R.; Hardardottir, I.; Memon, R.; Krul, E.J.; Moser, A.H.; Taylor, J.M.; Grunfeld, C. Effect of Endotoxin on Cholesterol Biosynthesis and Distribution in Serum Lipoproteins in Syrian Hamsters. J. Lipid Res. 1993, 34, 2147–2158. [Google Scholar] [CrossRef]
  144. Read, T.E.; Harris, H.W.; Grunfeld, C.; Feingold, K.R.; Kane, J.P.; Rapp, J.H. The Protective Effect of Serum Lipoproteins against Bacterial Lipopolysaccharide. Eur. Heart J. 1993, 14 (Suppl. K), 125–129. [Google Scholar]
  145. Levels, J.H.M.; Lemaire, L.C.J.M.; van den Ende, A.E.; van Deventer, S.J.H.; van Lanschot, J.J.B. Lipid Composition and Lipopolysaccharide Binding Capacity of Lipoproteins in Plasma and Lymph of Patients with Systemic Inflammatory Response Syndrome and Multiple Organ Failure. Crit. Care Med. 2003, 31, 1647–1653. [Google Scholar] [CrossRef] [PubMed]
  146. Wree, A.; McGeough, M.D.; Peña, C.A.; Schlattjan, M.; Li, H.; Inzaugarat, M.E.; Messer, K.; Canbay, A.; Hoffman, H.M.; Feldstein, A.E. NLRP3 Inflammasome Activation Is Required for Fibrosis Development in NAFLD. J. Mol. Med. 2014, 92, 1069–1082. [Google Scholar] [CrossRef]
  147. Mahowald, M.A.; Rey, F.E.; Seedorf, H.; Turnbaugh, P.J.; Fulton, R.S.; Wollam, A.; Shah, N.; Wang, C.; Magrini, V.; Wilson, R.K.; et al. Characterizing a Model Human Gut Microbiota Composed of Members of Its Two Dominant Bacterial Phyla. Proc. Natl. Acad. Sci. USA 2009, 106, 5859–5864. [Google Scholar] [CrossRef] [PubMed]
  148. Rios-Covian, D.; Gueimonde, M.; Duncan, S.H.; Flint, H.J.; de los Reyes-Gavilan, C.G. Enhanced Butyrate Formation by Cross-Feeding between Faecalibacterium prausnitzii and Bifidobacterium adolescentis. FEMS Microbiol. Lett. 2015, 362, fnv176. [Google Scholar] [CrossRef] [PubMed]
  149. Miller, T.L.; Wolin, M.J. Pathways of Acetate, Propionate, and Butyrate Formation by the Human Fecal Microbial Flora. Appl. Environ. Microbiol. 1996, 62, 1589–1592. [Google Scholar] [CrossRef] [PubMed]
  150. Xu, S.-Y.; Huang, X.; Cheong, K.-L. Recent Advances in Marine Algae Polysaccharides: Isolation, Structure, and Activities. Mar. Drugs 2017, 15, 388. [Google Scholar] [CrossRef]
  151. Sun, X.; Duan, M.; Liu, Y.; Luo, T.; Na, S. The Beneficial Effects of Gracilaria Lemaneiformis Polysaccharides on Obesity and the Gut Microbiota in High Fat Diet-Fed Mice. J. Funct. Foods 2018, 46, 48–56. [Google Scholar] [CrossRef]
  152. Wei, J.; Zhao, Y.; Zhou, C.; Zhao, Q.; Zhong, H.; Zhu, X.; Fu, T.; Pan, L.; Shang, Q.; Yu, G. Dietary Polysaccharide from Enteromorpha Clathrata Attenuates Obesity and Increases the Intestinal Abundance of Butyrate-Producing Bacterium, Eubacterium Xylanophilum, in Mice Fed a High-Fat Diet. Polymers 2021, 13, 3286. [Google Scholar] [CrossRef]
  153. Yang, Z.; Liu, G.; Wang, Y.; Yin, J.; Wang, J.; Xia, B.; Li, T.; Yang, X.; Hou, P.; Hu, S.; et al. Fucoidan A2 from the Brown Seaweed Ascophyllum Nodosum Lowers Lipid by Improving Reverse Cholesterol Transport in C57BL/6J Mice Fed a High-Fat Diet. J. Agric. Food Chem. 2019, 67, 5782–5791. [Google Scholar] [CrossRef]
  154. Busnelli, M.; Manzini, S.; Chiesa, G. The Gut Microbiota Affects Host Pathophysiology as an Endocrine Organ: A Focus on Cardiovascular Disease. Nutrients 2019, 12, 79. [Google Scholar] [CrossRef]
  155. Zhang, T.; Zhao, W.; Xie, B.; Liu, H. Effects of Auricularia Auricula and Its Polysaccharide on Diet-Induced Hyperlipidemia Rats by Modulating Gut Microbiota. J. Funct. Food. 2020, 72, 104038. [Google Scholar] [CrossRef]
  156. Shimizu, T.; Mori, K.; Ouchi, K.; Kushida, M.; Tsuduki, T. Effects of Dietary Intake of Japanese Mushrooms on Visceral Fat Accumulation and Gut Microbiota in Mice. Nutrients 2018, 10, 610. [Google Scholar] [CrossRef] [PubMed]
  157. Nakahara, D.; Nan, C.; Mori, K.; Hanayama, M.; Kikuchi, H.; Hirai, S.; Egashira, Y. Effect of Mushroom Polysaccharides from Pleurotus eryngii on Obesity and Gut Microbiota in Mice Fed a High-Fat Diet. Eur. J. Nutr. 2020, 59, 3231–3244. [Google Scholar] [CrossRef] [PubMed]
  158. Del Rio, D.; Stewart, A.J.; Mullen, W.; Burns, J.; Lean, M.E.J.; Brighenti, F.; Crozier, A. HPLC-MSn Analysis of Phenolic Compounds and Purine Alkaloids in Green and Black Tea. J. Agric. Food Chem. 2004, 52, 2807–2815. [Google Scholar] [CrossRef]
  159. Tzounis, X.; Vulevic, J.; Kuhnle, G.G.C.; George, T.; Leonczak, J.; Gibson, G.R.; Kwik-Uribe, C.; Spencer, J.P.E. Flavanol Monomer-Induced Changes to the Human Faecal Microflora. Br. J. Nutr. 2008, 99, 782–792. [Google Scholar] [CrossRef]
  160. Koh, A.; De Vadder, F.; Kovatcheva-Datchary, P.; Bäckhed, F. From Dietary Fiber to Host Physiology: Short-Chain Fatty Acids as Key Bacterial Metabolites. Cell 2016, 165, 1332–1345. [Google Scholar] [CrossRef]
  161. Tzounis, X.; Rodriguez-Mateos, A.; Vulevic, J.; Gibson, G.R.; Kwik-Uribe, C.; Spencer, J.P.E. Prebiotic Evaluation of Cocoa-Derived Flavanols in Healthy Humans by Using a Randomized, Controlled, Double-Blind, Crossover Intervention Study. Am. J. Clin. Nutr. 2011, 93, 62–72. [Google Scholar] [CrossRef]
  162. Requena, T.; Monagas, M.; Pozo-Bayon, M.A.; Martin-Alvarez, P.J.; Bartolome, B.; del Campo, R.; Avila, M.; Martinez-Cuesta, M.C.; Pelaez, C.; Moreno-Arribas, M.V. Perspectives of the Potential Implications of Wine Polyphenols on Human Oral and Gut Microbiota. Trends Food Sci. Technol. 2010, 21, 332–344. [Google Scholar] [CrossRef]
  163. Queipo-Ortuño, M.I.; Boto-Ordóñez, M.; Murri, M.; Gomez-Zumaquero, J.M.; Clemente-Postigo, M.; Estruch, R.; Cardona Diaz, F.; Andrés-Lacueva, C.; Tinahones, F.J. Influence of Red Wine Polyphenols and Ethanol on the Gut Microbiota Ecology and Biochemical Biomarkers. Am. J. Clin. Nutr. 2012, 95, 1323–1334. [Google Scholar] [CrossRef]
  164. Duda-Chodak, A. The Inhibitory Effect of Polyphenols on Human Gut Microbiota. J. Physiol. Pharmacol. 2012, 63, 497–503. [Google Scholar]
  165. Puupponen-Pimiä, R.; Nohynek, L.; Meier, C.; Kähkönen, M.; Heinonen, M.; Hopia, A.; Oksman-Caldentey, K.M. Antimicrobial Properties of Phenolic Compounds from Berries. J. Appl. Microbiol. 2001, 90, 494–507. [Google Scholar] [CrossRef] [PubMed]
  166. Schoeler, M.; Caesar, R. Dietary Lipids, Gut Microbiota and Lipid Metabolism. Rev. Endocr. Metab. Disord. 2019, 20, 461–472. [Google Scholar] [CrossRef] [PubMed]
  167. Caesar, R.; Tremaroli, V.; Kovatcheva-Datchary, P.; Cani, P.D.; Bäckhed, F. Crosstalk between Gut Microbiota and Dietary Lipids Aggravates WAT Inflammation through TLR Signaling. Cell Metab. 2015, 22, 658–668. [Google Scholar] [CrossRef] [PubMed]
  168. Mujico, J.R.; Baccan, G.C.; Gheorghe, A.; Díaz, L.E.; Marcos, A. Changes in Gut Microbiota Due to Supplemented Fatty Acids in Diet-Induced Obese Mice. Br. J. Nutr. 2013, 110, 711–720. [Google Scholar] [CrossRef] [PubMed]
  169. Pirestani, S.; Sahari, M.A.; Barzegar, M.; Nikoopour, H. Lipid, Cholesterol and Fatty Acid Profile of Some Commercially Important Fish Species from South Caspian Sea. J. Food Biochem. 2010, 34, 886–895. [Google Scholar] [CrossRef]
  170. Nazemroaya, S.; Sahari, M.A.; Rezaei, M. Identification of Fatty Acid in Mackerel (Scomberomorus commersoni) and Shark (Carcharhinus dussumieri) Fillets and Their Changes during Six Months of Frozen Storage at −18 °C. J. Agric. Sci. Technol. 2011, 13, 553–566. [Google Scholar]
  171. Ratledge, C.; Wynn, J.P. The biochemistry and molecular biology of lipid accumulation in oleaginous microorganisms. Adv. Appl. Microbiol. 2002, 51, 1–51. [Google Scholar]
  172. Abedi, E.; Sahari, M.A. Long-Chain Polyunsaturated Fatty Acid Sources and Evaluation of Their Nutritional and Functional Properties. Food Sci. Nutr. 2014, 2, 443–463. [Google Scholar] [CrossRef]
  173. Ahn, D.U.; Lutz, S.; Sim, J.S. Effects of Dietary α-Linolenic Acid on the Fatty Acid Composition, Storage Stability and Sensory Characteristics of Pork Loin. Meat Sci. 1996, 43, 291–299. [Google Scholar] [CrossRef]
  174. Calder, P.C.; Yaqoob, P. Omega-3 Polyunsaturated Fatty Acids and Human Health Outcomes. Biofactors 2009, 35, 266–272. [Google Scholar] [CrossRef]
  175. Liu, Y.; Song, X.; Zhou, H.; Zhou, X.; Xia, Y.; Dong, X.; Zhong, W.; Tang, S.; Wang, L.; Wen, S.; et al. Gut Microbiome Associates With Lipid-Lowering Effect of Rosuvastatin in Vivo. Front. Microbiol. 2018, 9, 530. [Google Scholar] [CrossRef]
  176. Tong, A.-J.; Hu, R.-K.; Wu, L.-X.; Lv, X.-C.; Li, X.; Zhao, L.-N.; Liu, B. Ganoderma Polysaccharide and Chitosan Synergistically Ameliorate Lipid Metabolic Disorders and Modulate Gut Microbiota Composition in High Fat Diet-Fed Golden Hamsters. J. Food Biochem. 2020, 44, e13109. [Google Scholar] [CrossRef] [PubMed]
  177. Antharam, V.C.; McEwen, D.C.; Garrett, T.J.; Dossey, A.T.; Li, E.C.; Kozlov, A.N.; Mesbah, Z.; Wang, G.P. An Integrated Metabolomic and Microbiome Analysis Identified Specific Gut Microbiota Associated with Fecal Cholesterol and Coprostanol in Clostridium difficile Infection. PLoS ONE 2016, 11, e0148824. [Google Scholar] [CrossRef] [PubMed]
  178. Watson, H.; Mitra, S.; Croden, F.C.; Taylor, M.; Wood, H.M.; Perry, S.L.; Spencer, J.A.; Quirke, P.; Toogood, G.J.; Lawton, C.L.; et al. A Randomised Trial of the Effect of Omega-3 Polyunsaturated Fatty Acid Supplements on the Human Intestinal Microbiota. Gut 2018, 67, 1974–1983. [Google Scholar] [CrossRef] [PubMed]
  179. Tindall, A.M.; McLimans, C.J.; Petersen, K.S.; Kris-Etherton, P.M.; Lamendella, R. Walnuts and Vegetable Oils Containing Oleic Acid Differentially Affect the Gut Microbiota and Associations with Cardiovascular Risk Factors: Follow-up of a Randomized, Controlled, Feeding Trial in Adults at Risk for Cardiovascular Disease. J. Nutr. 2020, 150, 806–817. [Google Scholar] [CrossRef] [PubMed]
  180. Wan, J.; Hu, S.; Jacoby, J.J.; Liu, J.; Zhang, Y.; Yu, L.L. The Impact of Dietary Sn-2 Palmitic Triacylglycerols in Combination with Docosahexaenoic Acid or Arachidonic Acid on Lipid Metabolism and Host Faecal Microbiota Composition in Sprague Dawley Rats. Food Funct. 2017, 8, 1793–1802. [Google Scholar] [CrossRef]
  181. Li, T.-T.; Tong, A.-J.; Liu, Y.-Y.; Huang, Z.-R.; Wan, X.-Z.; Pan, Y.-Y.; Jia, R.-B.; Liu, B.; Chen, X.-H.; Zhao, C. Polyunsaturated Fatty Acids from Microalgae Spirulina platensis Modulates Lipid Metabolism Disorders and Gut Microbiota in High-Fat Diet Rats. Food Chem. Toxicol. 2019, 131, 110558. [Google Scholar] [CrossRef]
  182. Fawzy, M.; Mohamed, M. Functional Bioactive Compounds and Biological Activities of Spirulina platensis Lipids. Czech J. Food Sci. 2008, 26, 55–64. [Google Scholar]
  183. Nakayama, J.; Watanabe, K.; Jiang, J.; Matsuda, K.; Chao, S.-H.; Haryono, P.; La-Ongkham, O.; Sarwoko, M.-A.; Sujaya, I.N.; Zhao, L.; et al. Diversity in Gut Bacterial Community of School-Age Children in Asia. Sci. Rep. 2015, 5, 8397. [Google Scholar] [CrossRef]
  184. Miras-Moreno, B.; Sabater-Jara, A.B.; Pedreño, M.A.; Almagro, L. Bioactivity of Phytosterols and Their Production in Plant in Vitro Cultures. J. Agric. Food Chem. 2016, 64, 7049–7058. [Google Scholar] [CrossRef]
  185. Wang, X.; Guan, L.; Zhao, Y.; Lei, L.; Liu, Y.; Ma, K.Y.; Wang, L.; Man, S.W.; Wang, J.; Huang, Y.; et al. Plasma Cholesterol-Lowering Activity of Dietary Dihydrocholesterol in Hypercholesterolemia Hamsters. Atherosclerosis 2015, 242, 77–86. [Google Scholar] [CrossRef] [PubMed]
  186. Bortolomeazzi, R.; De Zan, M.; Pizzale, L.; Conte, L.S. Mass Spectrometry Characterization of the 5alpha-, 7alpha-, and 7beta-Hydroxy Derivatives of Beta-Sitosterol, Campesterol, Stigmasterol, and Brassicasterol. J. Agric. Food Chem. 1999, 47, 3069–3074. [Google Scholar] [CrossRef] [PubMed]
  187. Pasternak, R.C. Report of the Adult Treatment Panel III: The 2001 National Cholesterol Education Program Guidelines on the Detection, Evaluation and Treatment of Elevated Cholesterol in Adults. Cardiol. Clin. 2003, 21, 393–398. [Google Scholar] [CrossRef] [PubMed]
  188. Liu, D.; Pi, J.; Zhang, B.; Zeng, H.; Li, C.; Xiao, Z.; Fang, F.; Liu, M.; Deng, N.; Wang, J. Phytosterol of Lotus Seed Core Powder Alleviates Hypercholesterolemia by Regulating Gut Microbiota in High-Cholesterol Diet-Induced C57BL/6J Mice. Food Biosci. 2023, 51, 102279. [Google Scholar] [CrossRef]
  189. Hatziioanou, D.; Mayer, M.J.; Duncan, S.H.; Flint, H.J.; Narbad, A. A Representative of the Dominant Human Colonic Firmicutes, Roseburia faecis M72/1, Forms a Novel Bacteriocin-like Substance. Anaerobe 2013, 23, 5–8. [Google Scholar] [CrossRef] [PubMed]
  190. Ramdath, D.D.; Padhi, E.M.T.; Sarfaraz, S.; Renwick, S.; Duncan, A.M. Beyond the Cholesterol-Lowering Effect of Soy Protein: A Review of the Effects of Dietary Soy and Its Constituents on Risk Factors for Cardiovascular Disease. Nutrients 2017, 9, 324. [Google Scholar] [CrossRef] [PubMed]
  191. Li, X.; Zhang, Z.; Cheng, J.; Diao, C.; Yan, Y.; Liu, D.; Wang, H.; Zheng, F. Dietary Supplementation of Soybean-Derived Sterols Regulates Cholesterol Metabolism and Intestinal Microbiota in Hamsters. J. Funct. Food. 2019, 59, 242–250. [Google Scholar] [CrossRef]
  192. Everard, A.; Belzer, C.; Geurts, L.; Ouwerkerk, J.P.; Druart, C.; Bindels, L.B.; Guiot, Y.; Derrien, M.; Muccioli, G.G.; Delzenne, N.M.; et al. Cross-Talk between Akkermansia muciniphila and Intestinal Epithelium Controls Diet-Induced Obesity. Proc. Natl. Acad. Sci. USA 2013, 110, 9066–9071. [Google Scholar] [CrossRef]
  193. Li, L.; Xu, L.J.; Ma, G.Z.; Dong, Y.M.; Peng, Y.; Xiao, P.G. The Large-Leaved Kudingcha (Ilex latifolia Thunb and Ilex kudingcha C.J. Tseng): A Traditional Chinese Tea with Plentiful Secondary Metabolites and Potential Biological Activities. J. Nat. Med. 2013, 67, 425–437. [Google Scholar] [CrossRef]
  194. Wan, P.; Peng, Y.; Chen, G.; Xie, M.; Dai, Z.; Huang, K.; Dong, W.; Zeng, X.; Sun, Y. Dicaffeoylquinic Acids from Ilex Kudingcha Attenuate Dextran Sulfate Sodium-Induced Colitis in C57BL/6 Mice in Association with the Modulation of Gut Microbiota. J. Funct. Food. 2019, 61, 103468. [Google Scholar] [CrossRef]
  195. Morgan, X.C.; Tickle, T.L.; Sokol, H.; Gevers, D.; Devaney, K.L.; Ward, D.V.; Reyes, J.A.; Shah, S.A.; LeLeiko, N.; Snapper, S.B.; et al. Dysfunction of the Intestinal Microbiome in Inflammatory Bowel Disease and Treatment. Genome Biol. 2012, 13, R79. [Google Scholar] [CrossRef] [PubMed]
  196. Jakobsdottir, G.; Xu, J.; Molin, G.; Ahrné, S.; Nyman, M. High-Fat Diet Reduces the Formation of Butyrate, but Increases Succinate, Inflammation, Liver Fat and Cholesterol in Rats, While Dietary Fibre Counteracts These Effects. PLoS ONE 2013, 8, e80476. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Absorption, biosynthesis, and metabolic process of cholesterol. ABCG5/8, ATP binding transporter protein 5/8; ACAT2, Acyl-coenzyme: cholesterol acyltransferases 2; BA, Bile acid; CA, Cholic acid; CDCA, Chenodeoxycholic acid; CM, Chylomicron; DCA, Deoxycholic acid; HMG-CoA-R, HMG-CoA reductase; HMG-CoA-S, HMG-CoA synthase; IPP, Isopentoyl disphosphate; LCA, Lithocholic acid; MTP, Microsomal triglyceride transfer protein; MVA, Mevalonate; NPC1L1, Niemann-Pick C1-like 1 protein; SQ, squalene. Foods 12 04308 i001, cholesterol; Foods 12 04308 i002, bile acids.
Figure 1. Absorption, biosynthesis, and metabolic process of cholesterol. ABCG5/8, ATP binding transporter protein 5/8; ACAT2, Acyl-coenzyme: cholesterol acyltransferases 2; BA, Bile acid; CA, Cholic acid; CDCA, Chenodeoxycholic acid; CM, Chylomicron; DCA, Deoxycholic acid; HMG-CoA-R, HMG-CoA reductase; HMG-CoA-S, HMG-CoA synthase; IPP, Isopentoyl disphosphate; LCA, Lithocholic acid; MTP, Microsomal triglyceride transfer protein; MVA, Mevalonate; NPC1L1, Niemann-Pick C1-like 1 protein; SQ, squalene. Foods 12 04308 i001, cholesterol; Foods 12 04308 i002, bile acids.
Foods 12 04308 g001
Figure 2. BSH, an advantageous enzyme synthesized by the gut microbiota, participates in cholesterol metabolism through the hydrolysis of conjugated bile acids.
Figure 2. BSH, an advantageous enzyme synthesized by the gut microbiota, participates in cholesterol metabolism through the hydrolysis of conjugated bile acids.
Foods 12 04308 g002
Figure 3. The role of SCFAs, metabolic products of gut microbiota, in cholesterol metabolism. Bu, Butyrate; CETP, Cholesteryl ester transfer protein; HDL-C, High-density lipoproteins cholesterol; HMG-CoA-R, HMG-CoA reductase; IDL, Intermediate-density lipoproteins; IPP, Isopentoyl disphosphate; LDL-C, Low-density lipoprotein cholesterol; LDLR, LDL receptors; MVA, Mevalonate; SQ, Squalene; SREBP2, Sterol-regulatory element binding protein-2; VLDL-C, Very-low-density lipoproteins cholesterol.
Figure 3. The role of SCFAs, metabolic products of gut microbiota, in cholesterol metabolism. Bu, Butyrate; CETP, Cholesteryl ester transfer protein; HDL-C, High-density lipoproteins cholesterol; HMG-CoA-R, HMG-CoA reductase; IDL, Intermediate-density lipoproteins; IPP, Isopentoyl disphosphate; LDL-C, Low-density lipoprotein cholesterol; LDLR, LDL receptors; MVA, Mevalonate; SQ, Squalene; SREBP2, Sterol-regulatory element binding protein-2; VLDL-C, Very-low-density lipoproteins cholesterol.
Foods 12 04308 g003
Figure 4. Metabolic pathway of cholesterol conversion to coprostanol.
Figure 4. Metabolic pathway of cholesterol conversion to coprostanol.
Foods 12 04308 g004
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

Deng, C.; Pan, J.; Zhu, H.; Chen, Z.-Y. Effect of Gut Microbiota on Blood Cholesterol: A Review on Mechanisms. Foods 2023, 12, 4308. https://doi.org/10.3390/foods12234308

AMA Style

Deng C, Pan J, Zhu H, Chen Z-Y. Effect of Gut Microbiota on Blood Cholesterol: A Review on Mechanisms. Foods. 2023; 12(23):4308. https://doi.org/10.3390/foods12234308

Chicago/Turabian Style

Deng, Chuanling, Jingjin Pan, Hanyue Zhu, and Zhen-Yu Chen. 2023. "Effect of Gut Microbiota on Blood Cholesterol: A Review on Mechanisms" Foods 12, no. 23: 4308. https://doi.org/10.3390/foods12234308

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