Next Article in Journal
Technological Aspects and Evaluation Methods for Polymer Matrices as Dental Drug Carriers
Previous Article in Journal
Identification of New Key Genes and Their Association with Breast Cancer Occurrence and Poor Survival Using In Silico and In Vitro Methods
Previous Article in Special Issue
Muscle and Muscle-like Autoantigen Expression in Myasthenia Gravis Thymus: Possible Molecular Hint for Autosensitization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gut Dysbiosis and Blood-Brain Barrier Alteration in Hepatic Encephalopathy: From Gut to Brain

1
Cellular and Molecular Research Center, Iran University of Medical Sciences, Tehran 1449614535, Iran
2
Department of Neuroscience, Faculty of Advanced Technologies in Medicine, Iran University of Medical Sciences, Tehran 1449614535, Iran
3
Neuroscience Research Center, Mashhad University of Medical Sciences, Mashhad 9919191778, Iran
4
Department of Microbiology, Faculty of Medicine, Mazandaran University of Medical Sciences, Sari 4815733971, Iran
5
Research Development Center, Sina Hospital, Tehran University of Medical Sciences, Tehran 1417653761, Iran
6
Digestive Disease Research Institute, Tehran University of Medical Sciences, Tehran 1417653761, Iran
7
Department of Neuroscience, Faculty of Medicine, Mashhad University of Medical Sciences, Mashhad 9919191778, Iran
8
Shefa Neuroscience Research Center, Khatam Alanbia Hospital, Tehran 9815733169, Iran
9
Department of Intestinal Failure and Liver Diseases, Rigshospitalet, Inge Lehmanns Vej 5, 2100 Copenhagen, Denmark
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2023, 11(5), 1272; https://doi.org/10.3390/biomedicines11051272
Submission received: 4 February 2023 / Revised: 20 March 2023 / Accepted: 28 March 2023 / Published: 25 April 2023
(This article belongs to the Special Issue Molecular Mechanisms of Neurological Autoimmune Disorders)

Abstract

:
A common neuropsychiatric complication of advanced liver disease, hepatic encephalopathy (HE), impacts the quality of life and length of hospital stays. There is new evidence that gut microbiota plays a significant role in brain development and cerebral homeostasis. Microbiota metabolites are providing a new avenue of therapeutic options for several neurological-related disorders. For instance, the gut microbiota composition and blood-brain barrier (BBB) integrity are altered in HE in a variety of clinical and experimental studies. Furthermore, probiotics, prebiotics, antibiotics, and fecal microbiota transplantation have been shown to positively affect BBB integrity in disease models that are potentially extendable to HE by targeting gut microbiota. However, the mechanisms that underlie microbiota dysbiosis and its effects on the BBB are still unclear in HE. To this end, the aim of this review was to summarize the clinical and experimental evidence of gut dysbiosis and BBB disruption in HE and a possible mechanism.

1. Introduction

Hepatic encephalopathy (HE) is the main complication of acute liver failure (ALF) and advanced liver diseases; it is characterized as a set of complex spectra of brain dysfunctions and neuropsychiatric impairments, ranging between preclinical changes and deep coma [1,2]. In addition to economical burdens, HE severely affects daily routines and reduces the quality of life in patients and their caregivers [3]. According to the etiology of the disease, HE is divided into three types. Type A is associated with ALF, type B is associated with portosystemic shunts, and type C results from portal hypertension and cirrhosis. The European Association for the Study of the Liver (EASL) has categorized HE as being either covert HE and overt HE, according to the severity of the complications. Based on the West Haven Criteria, covert HE is concluded as minimal or grade I of HE, which is only diagnosed by neuropsychiatric tests, while overt HE is considered grade II-IV of HE [4]. The prevalence of HE is associated with the etiology of the disease, the severity of the complications, and the type of HE. The prevalence of covert or minimal HE is about 20–80% in patients with cirrhosis [5,6,7,8,9]. Approximately 30–40% of patients with cirrhosis experience overt HE during their course of the disease and it is associated with risk factors such as ascites, infections, esophageal bleeding, diabetes, and covert HE [10,11,12,13]. Although the pathogenesis of HE remains incompletely elucidated, the central role of accumulated gut-derived agents, particularly ammonia, has been confirmed [14]. An axial role for gut microbiota in the pathophysiology of HE, which known as the gut-brain axis, has also been suggested [15]. Furthermore, the impairment of the blood-brain barrier (BBB) has been reported in many experimental models, as well as in patients with HE. This review aims to demonstrate how gut dysbiosis disrupts the integrity of the BBB following HE and to discuss how alterations in gut microbiota diversity affect the BBB integrity in HE.

2. Gut Dysbiosis in Hepatic Encephalopathy

The term gut microbiome, previously known as gut normal flora, encompasses all microorganisms, including bacteria, fungi (fungal mycobiome), viruses (virome), and archaea that inhabit the small intestine and colon [16]. One to two years after birth, the microbiota community is created, and its diversity is altered under the influence of age, diet, lifestyle, and geography during one’s lifetime [17]. The gut microbiomes synthetize the enzymes that break down some indigestible polysaccharides and carbohydrates to absorbable glucose and short-chain fatty acids. The gut microbiomes also are responsible for the production of vitamins, including vitamin K, cobalamin (B12), biotin (B7), and folate (B9), as well as facilitating the absorption of calcium, magnesium, and iron by enterocytes [18]. Normal gut microorganisms compete for luminal contents with pathogens and prevent the colonization of these organisms within the gut. Furthermore, gut microbiomes are in close contact with the lamina propria where a huge number of intestinal immune cells are located that develop the host immune system and increase intestinal protection [19]. The intestinal tract is well-innervated by the enteric nervous system and vagus nerve, which comprise the transmission pathways for signals from the gut to the central nervous system (CNS) [20]. Some clinical and experimental evidence has confirmed that gut dysbiosis happens in cirrhotic patients with HE and experimental animal models (Table 1). These studies are mainly conducted through metagenomics and metatranscriptomics approaches (Box 1). Gut microbiota analysis in cirrhotic patients with HE revealed a significant decrease in Clostridiales XIV, Ruminococcaceae, and Lachnospiraceae as the autochthonous taxa with a significant enhancement in Enterococcaeae, Staphylococcaceae and Enterobacteriaceae as the pathogenic taxa compared to healthy controls [21]. Increased pathogenic taxa (i.e., Enterobacteriaceae and Enterococcaceae) and decreased autochthonous taxa (i.e., Lachnospiraceae, Ruminococcaceae and Clostridiales XIV) were also reported in fecal analysis in cirrhotic patients with HE compared to cirrhotic patients without HE [22]. In addition, dysbiosis was well correlated with systemic inflammation in these patients [22].
Box 1. Metagenomics and metatranscriptomics analyses in microbiome studies.
Studies in microbiota communities are a field of great interest due to changes in their composition associated with health and diseases. To identify the structure and functional dynamics of the gut microbiome, several multi-omics approaches such as metagenomics (DNAs interactions) and metatranscriptomics (RNAs interactions) are applicable. Each approach images different types of biological information from the microbiota communities. Current investigations combine more than one omics approach to image a high-resolution picture of the gut microbiota dynamics [41]. Genomics is concerned with identifying the genetic composition of a single bacterium, but metagenomics is concerned with identifying the genetic composition of an entire community of bacterial communities (for instance, the entire gut microbiome of humans) [42]. Metagenomics consists of several culture-independent techniques (i.e., experimental and bioinformatic approaches) to analyze extracted DNA directly from biological samples, such as saliva, stool, and sputum [43]. In summary, the DNA extracted from bacterial communities amplified the DNA containing 16S ribosomal RNA (rRNA as a housekeeping genetic marker) gene and was sequenced. Afterwards, similar sequences as operational taxonomic units (OTUs) were compared to several reference 16S databases (i.e., RDP, SILVA, and Greengenes) to identify the microbiota composition in the samples based on the OTUs similarities [44,45]. In shotgun metagenomic sequencing, all extracted community DNA is sequenced and compared to some reference genomes in several databases (i.e., KEGG, and BLAST) to specify the abundance, diversity, and function of the microbiome [45,46]. Metatranscriptomics is a powerful tool to specify the expression patterns of genes (functional profile; total mRNA) in sequenced genomes from communities of bacteria in a particular sample. This approach provides information about the diversity of active genes and also identifies differences in the gene expression patterns following health and disease in communities of bacteria [47]. In metatranscriptomics microbiome studies, first the total RNA is extracted from bacterial communities, followed by the purification of the desired RNA (i.e., mRNA, microRNA, and lincRNA) and the ribosomal RNA is removed. Then, the RNA is reverse transcribed into cDNA and is amplified to induce a cDNA library. After that, the library is sequenced for further comparing with reference genomes in some databases (i.e., KEGG, CARD, VFDB, and Kraken) [48,49]

3. Blood-Brain Barrier Structure and Transportation Systems

The BBB is a complex and highly selective border that forms an important part of the neurovascular unit in the CNS. This unique structure protects neuronal cells from direct contact with circulatory neurotoxic agents, pathogens, and peripheral inflammation [50]. The BBB is well organized in such a way that allows the passage of some small hydrophobic molecules (i.e., oxygen and carbon monoxide) by passive diffusion, and glucose and amino acids through active transport. The barrier is comprised of one single layer of muscle-loss endothelial cells that are covered by a basement membrane, pericytes, and astrocyte end-feet (Figure 1) [51]. The first obstacle against circulatory neurotoxic molecules is a thin monolayer of endothelial cells that are connected via tight junctions and adherent junctions. The structure of these cells is such that it distinguishes them from endothelial cells in the other tissues. They strongly seal the paracellular pathway through highly expressed tight junction proteins that allow them to severely regulate the flux of ions and molecules between the blood and brain tissue [52]. Cerebral endothelial cells also have lower caveolae on their luminal membrane, a higher number of mitochondria, lower levels of leukocyte adhesion molecule, and a restricted transcytosis rate compared to the endothelial cells in other tissues [52]. The tight junctions in place of endothelial junctional clefts are formed by many transmembrane and cytoplasmic proteins that are responsible for sealing the paracellular pathway and restricting the movement of solutes. Two major transmembrane proteins occludin and claudin (claudin-5), as well as one cytoplasmic protein zonula occludens-1 (ZO-1), are formed in the endothelial tight junctions (Figure 1) [53]. Occludin and claudin are integral membrane proteins that have a significant role in the sealing and barrier function of the endothelial junctional clefts, while ZO-1, as a membrane-associated guanylate kinase and a scaffold protein, anchors transmembrane proteins to the actin cytoskeleton, which intensifies the stability of the tight junctions and has a significant role in signaling between cells [53]. Another group of proteins, adherens proteins, are expressed in the junctional clefts that stabilize and regulate the function of tight junction transmembrane proteins. Some important adherens proteins that are present in the endothelial junctional clefts include junctional adhesion molecules (JAMs), platelet endothelial cell adhesion molecule 1 (PECAM-1), endothelial cell-selective adhesion molecule (ESAM), and catenins [54].
Only three transmission routes can be considered for solute transportation across the BBB (Figure 1); The first transmission route is a sealed paracellular pathway by tight junction that is allowed hydrated ions to pass in a limited way [55]. The second path is transcytosis transport (specifically endocytosis), which mediates the transportation of some macromolecules and drugs across the BBB. Transcytosis is most commonly observed in two forms through the BBB, such as receptor-mediated transport and absorptive-mediated transport [55,56]. Receptor-mediated transport or clathrin-mediated endocytosis is a transportation pathway that is mediated by clathrin proteins and requires the interaction of a specific substrate to its receptor on the endothelial membrane [57]. Ligand-receptor interaction on the plasma membrane triggers the invagination of the membrane through a clathrin-dependent mechanism [58]. Finally, a ligand in a clathrin-coated pit can enter the cell. Transporting low-density lipoprotein (LDL) by the LDL receptor, iron via the transferrin receptor, and the insulin-like growth factor receptor are three important examples of receptor-mediated transport through the BBB [55,59]. Absorptive-mediated transport or caveolae-mediated transport is another form of transportation for macromolecules and drugs across the BBB that are triggered by the shear stress that is induced by the substrate and its interaction with endothelial glycocalyx at the luminal side [60,61]. The interaction of the ligand with glycocalyx can trigger the oligomerization of the mechanosensory protein caveolins, caveolar invagination, and endocytosis of the ligand [58]. Beyond the aforementioned routes, the brain endothelial cells express a number of transporters to uptake essential substrates as the third transmission route through the BBB [62,63]. This type of transportation is categorized into three forms: active transporter, carrier-mediated transport, and ion transporters [63]. ATP-binding cassette (ABC) transporters such as P-glycoprotein 1 (P-gp) are common examples of active transport in the BBB that hydrolyze ATP to utilize its energy to efflux many substrates or drug metabolisms across the concentration gradient from the brain tissue [64]. Carrier-mediated transporters are responsible for transporting glucose, specific ions, amino acids, organic anions, cations, and other substances with specific properties that cannot directly pass through the cell membrane [65]. Carrier-mediated transporters may be categorized into three types: uniporter, symporter, and antiporter. Glucose transporter 1 (GLUT1) is an example of uniporter-mediated transport, which is mediated by the movement of glucose down its concentration gradient across the BBB [66]. Monocarboxylate transporter 1 (MCT1) is an example of symporter-mediated transport that passes both the proton and lactate in the same direction [67]. Organic anion transporting polypeptides (OATPs) such OATP1A2 are mainly expressed in the apical membrane, and OATP2B1 is expressed in the basal membrane of brain endothelial cells. These are other examples of carrier-mediated transport across the BBB, which mediate the uptake and efflux of neurosteroids and thyroid hormones [68]. L-type amino acid transporter 1 (LAT1) is a common example of antiporter-mediated transport in the BBB. These transporters regulate the influx of neutral essential amino acids (i.e., histidine, tryptophan, leucine, isoleucine, phenylalanine and tyrosine) into the brain endothelial cell and astrocytes in exchange with the efflux of glutamine [69]. The endothelial cells also highly express many ion transporters, such as Na+-K+ ATPase, Na+-K+-Cl cotransporter, K+ channels, Na+/Ca2+, Na+/H+, and Cl/HCO3 exchangers, Na+/HCO3 symporters and Ca2+ transporters severely regulate the sodium gradient concentration to ensure the uptake of sodium-dependent substrates, maintain the intracellular pH, and regulate the brain interstitial ions equilibrium and water content [70].
The second cellular components of the BBB are pericytes, which are located in the basement membrane in close contact with the abluminal side of the endothelial cells via gap junctions and peg-socket junctions [71]. These cells can be visualized by some specific markers, for example, platelet-derived growth factor receptor beta (PDGFRβ) and neuron-glial antigen 2 (NG2) in mice. Pericytes are important in the development of tight junctions and the formation of the BBB. These cells stabilize the brain endothelial cells and protect the brain from the invasion of immune cells [72]. Pericytes also synthetize some proteins that regulate the capillary blood flow and trigger angiogenesis [73]. The paracrine interaction of endothelial cells and pericytes is also well regulated through platelet-derived growth factor beta (PDGF-β), which is released from endothelial cells and binds to its receptors on pericytes. This ligand-receptor interaction triggers signaling cascades that mediate the proliferation and migration of pericytes [74]. Astrocytes are the most numerus glial cells in the CNS that play many important functions, including ions homeostasis, the regulation of the pH in the extracellular space, controlling the cerebral blood flow, providing nutrients to the neurons, CNS repairing following injuries, and the regulation of endothelial cells’ function and neurotransmission [75,76]. The processes of these cells ensheath the microvasculature as astrocyte end-feet that enable them to send signals from activated neurons to regulate the blood flow [77]. The high expression of aquaporin 4 water channels (AQP4) has been identified in astrocytes, indicating that these cells play a major role in water homeostasis and the clearance of waste substances from the brain interstitial space [78]. The dystroglycan–dystrophin complex is also expressed and localized in the astrocyte end-feet that connect the end-feet cytoskeleton to the basement membrane [79,80]. It is concluded that this complex facilitates the localization of AQP4 in the perivascular space to maintain water flow and form a part of the glymphatic system [80]. Therefore, the BBB not only acts as a sealed physical barrier and a transportation border, but also is a metabolic structure, and its constituent cells can affect each other and their vicinity through signaling molecules.
The intact integrity and normal functioning of the cellular components of the BBB ensure the homeostasis of CNS and proper neuronal function [81]. Any small changes in this restricting barrier and its properties can confront the neuronal cells to direct contact with circulatory neurotoxic agents [81,82].
BBB dysfunction and an increase in the CNS levels of neurotoxic substrates has been reported in patients with HE and in experimental models (Table 2). In cases of a disrupted BBB following liver insufficiency, the circulating ammonia, mercaptans, lipopolysaccharide (LPS), and other gut-derived neurotoxic agents pass the BBB and make close contact with the neurons, astrocytes, and microglia, which trigger neuroinflammation and produce reactive oxygen species [83,84,85,86,87]. The disrupted BBB also allows an increase in the entry of immune cells and circulatory pro-inflammatory cytokines to the brain parenchyma and causes neuroinflammation and damage [88,89]. The question is how the BBB is disrupted in HE and whether gut microbiota-derived molecules are implicated.

4. Altered Gut Microbial Metabolites and Molecules May Disrupt the Integrity of the BBB in HE

The intestinal microbiome seems to be of central importance in the progression of HE [106,107]. The decreased expression of tight junction proteins and the increased BBB permeabilization have been addressed in the brain of germ-free mice compared with normal gut flora [108]. The gut microbiota produces some molecules and metabolites that can exert beneficial or harmful effects on the host CNS. Short-chain fatty acids (SCFAs, i.e., propionate, butyrate and acetate) are the final product of the bacterial fermentation of non-digestible polysaccharides in the lower intestinal tract and act as signaling molecules, have anti-inflammatory properties, and protect colonic epithelial cells [109]. Some studies have shown that basal non-toxic levels of SCFAs can preserve the intestinal barrier integrity, protect the BBB from oxidative stress, and positively regulate the expression of thigh junction proteins [110,111,112]. The treatment of germ-free mice with sodium butyrate strongly recovered the destructed BBB after visualizing the level of Evans blue in the frontal cortex, striatum, and hippocampus [108]. Moreover, the treatment of rhesus monkeys with antibiotics altered the gut microbiota composition, and in particular, decreased the SCFAs-producing phyla and impaired the permeability of the BBB in the thalamus [113]. A lower amount of propionate, butyrate, and acetate was seen in the fecal samples of cirrhotic patients with HE compared to those without HE [114]. The family Ruminococcaceae is the main source for the production of SCFAs in the human intestinal tract [115]. As shown in Table 1, this SCFAs-producing family of gut bacteria have been decreased in the gut of patients with cirrhosis and HE [22,27]. Specifically, the decreased genera in Ruminococcaceae and Lachnospiraceae families have been reported in fecal samples taken from cirrhotic patients, which were associated with a low capacity to produce butyrate in the intestine of patients [116]. The decreased basal concentration of SCFAs may indirectly affect the integrity of the BBB through the impairment of the intestinal barrier following cirrhosis. A recent systematic review reported that the administration of SCFAs recovers the destructed intestinal barrier and improves the severity of hepatic injury following liver disease [117]. The disruption of the normal physical intestinal barrier may induce microbial and endotoxin translocation, which result in systemic inflammation, liver injury, BBB permeabilization, and neuroinflammation in cirrhosis and HE [118,119]. The intestinal microbiome also contributes to the metabolism of bile acids and mediates the synthesis of a small part of bile acids (as secondary bile acids), which maintain the integrity of the intestinal barrier and regulate the intestinal immune responses through farsenoid X receptor (FXR) and bile acid receptor GPBAR-1 (TGR5) [120,121,122,123]. Decreased concentrations of intestinal secondary bile acids following gut dysbiosis are shown in cirrhosis [124,125]. Studies have shown that the concentrations of circulatory primary bile acids produced by hepatocytes increased [100,105,126,127,128], while the levels of intestinal secondary bile acids produced by the gut microbiome reduced in cirrhosis [124,125]. Therefore, a bile acids imbalance may affect the BBB with two bile acid-dependent approaches in cirrhosis: (1) an increased circulatory concentration of bile acids directly correlated with the BBB permeabilization through a Rac1-dependent mechanism [100]; (2) the decreased intestinal concentration of bile acids indirectly affect the integrity of the BBB via the destruction of the intestinal barrier, bacterial translocation, and the induction of systemic inflammation (Figure 2). The microbe- or pathogen-associated molecular patterns (MAMPs or PAMPs) are other important products of gut microbiota that are identified by expressed pattern recognition receptors (PRRs) on innate immune cells (i.e., natural killer cells, dendritic cells and monocytes/macrophages) [129]. Toll-like receptors (TLRs) are considered as the main groups of PRRs that recognize both MAMPs and endogenous damage-associated molecular pattern molecules (DAMPs) in the intestinal tract [130]. The activation of TLRs leads to a change in the expression of the downstream signaling proteins that mediate the immune responses to initiator agents. Some bacterial endotoxins, such as LPS, flagellin, and bacterial DNA, are considered to be the main MAMPs that are recognized via TLR4, TLR5, and TLR9, respectively, and are responsible for the activation of the immune system and the production of inflammatory cytokines following gut dysbiosis and bacterial infections [131]. A destructed intestinal barrier and intestinal bacterial overgrowth may explain the cause of bacterial translocation in cirrhosis. Clinical and experimental studies have shown bacterial translocation in cirrhosis [132,133,134,135,136,137,138]. The circulatory levels of potent inflammatory MAMPs and LPS increased and were associated with the risk of hospitalization and death in patients with advanced liver disease [139,140]. The activation of TLRs by LPS and other MAMPs can trigger other downstream signaling molecules, such as Interleukin-1 receptor-associated kinases (IRAKs) and TNF receptor-associated factor 6 (TRAF6), which increase the expression of some important inflammatory transcription factors (i.e., nuclear factor Kappa-B, activator protein 1, interferon regulatory factor 3, p38 mitogen-activated protein kinase and c-Jun N-terminal kinase) [129,141,142]. These translocated transcription factors trigger the gene expression of proinflammatory cytokines and chemokines, such as interleukin-1 beta, interleukin-6, tumor necrosis factor alpha, CXCL8 CXC motif chemokine ligand 8, and CXCL10 CXC motif chemokine ligand 10 [143,144,145]. The induction of systemic inflammation via the activation of TLRs in response to MAMPs can activate hepatic kupffer cells, induce liver damage, and make the BBB susceptible to disruption, as seen in experimental studies [83,146,147]. Increased circulatory levels of LPS were correlated with systemic inflammation, the down-regulation of tight junction proteins (i.e., claudin-5 and ZO-1), the production of oxidative stress, and neuroinflammation in rodents [148,149]. It has been reported that circulatory LPS can disrupt the integrity of the BBB via an increase in the expression of brain α-synuclein [150]. In azoxymethane-induced HE mice, the injection of LPS strongly exacerbated the concentration of pro-inflammatory cytokines, worsened hyperammonemia, increased liver injury, up-regulated brain matrix metalloproteinase-9, and severely increased the permeability of the BBB [83]. Moreover, the injection of LPS in galactosamine-induced HE mice led to the shrinkage of the brain endothelial cells, decreased tight junction protein occludin, and the raised extravasation of Evans blue dye into the brain in a TNFα-dependent manner [91]. Other probable gut microbiota-derived products, such as mercaptans and phenol, were able to impair the BBB in acute liver injury [151]. The Kupffer cells are resident liver macrophages that have an important role in the defense against infections and endotoxins under normal conditions [152]. In cirrhosis and liver failure, the impaired intestinal barrier leads to a rise in the circulatory concentrations of gut-derived endotoxins, which can reach the liver and activate Kupffer cells. Activated Kupffer cells trigger cytokine and chemokine production, polymorphonuclear cells infiltration, the induction of oxidative stress, and tissue hypoxia in liver parenchyma, which leads to hepatocyte injury and the release of pro-inflammatory cytokines into the circulation [152,153,154]. Furthermore, shunting ad gut-derived endotoxins through varices will only exacerbate a proinflammatory drive in the systemic circulation and impair the permeability of the BBB by a change in the expression of tight junction proteins, the activation of brain glial cells, and the infiltration of peripheral immune cells [103,105,155,156,157,158,159].
Gut microbiota-derived metabolites also affect the CNS through the enteric nervous system (ENS), which normally connects to the CNS via the vagus nerve and sympathetic pathways [160,161]. Gut microbiota products (derived from the small intestine and large intestinal bacteria; Box 2) can directly regulate the development and function of the ENS through MAMPs or PAMPs-PRRs interactions. The cellular components of ENS (i.e., enteric glial cells and neurons) express several PRRs, such as TLR2, TLR3, TLR4, TLR7, and TLR9, which activate in response to LPS, polysaccharide A, and RNAs, and lead to the regulation of the intestinal motility and contractility, as well as the maintenance of the enteric plexuses [162,163,164]. Furthermore, SCFAs as the final product of the bacterial fermentation of non-digestible polysaccharides, which can activate free fatty acid receptor 3 (FFAR3) on enteric neurons and enteroendocrine cells [165]. SCFAs also regulate the gastrointestinal motility by affecting ENS in rats [166]. In addition to the direct effects of the gut microbiota-derived products on the ENS, these products indirectly regulate the function of ENS through several interface cells (i.e., enterochromaffin cells, enteric macrophages and dendritic cells, and intestinal stromal cells) [167,168,169,170,171,172]. The activation of these intermediate cells by gut microbiota metabolites leads to the production of many products, such as serotonin, neuropeptide Y, gastric inhibitory polypeptide, glucagon-like peptide-1, substance P, bone morphogenetic protein 2, glial cell-derived neurotrophic factor, and transforming growth factor β1, which affect the neurons of the ENS [164,167,173,174,175,176]. Gut dysbiosis following liver diseases may negatively regulate the function of the CNS through the differentially produced gut products that prevent the normal regulatory functions of the ENS that have previously been induced by normal gut microbiome metabolites.
Box 2. Small intestine versus large intestine microbiome composition.
Bacteria constitute a large part of the gut microbiome compared to other microorganisms (i.e., fungi, viruses, protozoa, and archaea). In addition, the composition of the gut bacteria throughout the gastrointestinal tract (GI) is not the same [177,178]. The communities of bacteria increased from the upper GI (104–106 cells/gram feces) to the lower GI (109–1012 cells/gram feces) as a result of the lower pH, rapid movement of luminal contents, and the presence of antibacterial agents in the stomach and duodenum compared to the lower GI [177,179,180]. The family of Lactobacillaceae and Enterobacteriaceae are dominant bacteria in the stomach and small intestine (i.e., duodenum, jejunum, and ileum), while Bacteroidaceae, Prevotellaceae, Rikenellacea, Lachnospiraceae, and Ruminococcaceae are major dominant families of bacteria in the large intestine [179]. In microbiome studies, it is important to note that a new experimental comparative study by Ji-Seon Ahn et al. revealed that the fecal microbiome contents are not acceptable to be generalized to the entire gut microbiome [181].

5. Therapeutic Targets

Given that gut dysbiosis may contribute to the pathophysiology of BBB permeabilization in HE, therapeutic interventions that target gut dysbiosis and their metabolites may be useful for restoring the destructed BBB and its consequent immune cells infiltrations, oxidative stress, microglial activation, and neuroinflammation in HE. Antibiotic therapy (mainly rifaximin, neomycin, and metronidazole) targets the gut microbiota and decreases the production and absorption of intestinal ammonia, resulting in alleviated hyperammonemia, decreased hospitalization for overt HE, and an improved quality of life for patients with HE [182,183]. The efficacy of antibiotics on the integrity of the BBB is very limited and only one study has indicated that the administration of rifaximin in hyperammonemic BDL rats significantly decreased the florescent intensity of fluorochrome (an indicator of BBB permeabilization) in the brain tissue [104]. The knowledge of how the alternation and normalization of the gut dysbiosis influence the structure of the BBB in HE remains unknown. However, the modulation of the gut dysbiosis using some therapeutic options, such as probiotics and prebiotics, antibiotic therapy, and fecal microbiota transplantation (FMT), have shown beneficial effects on the BBB integrity in several experimental models, including Parkinson’s disease (AD), Alzheimer’s disease, depression, Gulf war illness, toxicity, inflammation, and aging (Table 3).
Probiotic supplements contain live microorganisms that modulate the composition of the gut microflora, while prebiotics are types of nutrients or agents (i.e., dietary fibers and non-absorbable disaccharide) that are beneficial for the growth and activity of the gut microbiota [199,200]. A combination of probiotic and prebiotic therapy in a chronic mice model of AD prevented striatal neuronal apoptosis, improved the integrity of the BBB, and showed neuroprotection effects [184]. In an APP/PS1 mice model of Alzheimer’s disease, the administration of a combination of beneficial bacteria decreased the signal intensity of Evans blue in hippocampus, increased the tight junction proteins ZO-1 and occludin, and reduced the cerebral concentration of LPS [185]. An enhanced cerebral expression of ZO-1 and occludin was also seen in a rat model of traumatic brain injury after the administration of probiotic Bifidobacterium lactis [201]. The decreased extravasation of Evans blue dye was shown in the prefrontal cortex of mice models of depressive-like behavior after the administration of a new probiotic Komagataella pastoris, KM71H [186]. In the mice model of stress, the administration of Lactobacillus plantarum MTCC 9510 significantly decreased the cerebral levels of Evans blue, increased brain-derived neurotrophic factor (BDNF), and decreased neuroinflammation and cerebral oxidative stress [188]. In addition, the treatment of the senescence-accelerated mouse prone 8 (SAMP8) mouse strain model of aging with some species of Lactobacillus and Bifidobacterium, as well as ProBiotic-4, reduced the BBB damage, decreased the LPS concentrations, and mitigated pro-inflammatory signaling cascades in the brain tissue [190,202]. In mouse models of Gulf war illness as a chronic and multi-symptomatic disorder, the administration of andrographolide decreased intestinal pro-inflammatory cytokines, restored the claudin-5 protein level, increased the BDNF, and decreased the microglial activation in cerebral tissues [191].
FMT is a procedure to collect feces from healthy individuals and transfer them into the recipient’s intestine, mainly in patients with gut dysbiosis [203]. Overall, FMT may improve hyperammonemia through the normalization of the gut microbiome composition to low urease bacteria, increasing the hepatocyte clearance of ammonia, and the improvement of the intestinal barrier structure in liver diseases [204]. The effect of FMT on the structure of the BBB in cirrhosis and HE has been less investigated. In a chronic rotenone administration-induced Parkinson’s disease mouse model, FMT significantly restored the tight junction proteins (i.e., occludin, claudin-5, and ZO-1), decreased the astrocyte reactivity, reduced the microglial activation, decreased the concentration of LPS, and suppressed the inflammatory cascade TLR4/MyD88/NF-κB in substantia nigra [195]. In addition, FMT attenuated the florescent intensity of Evans blue and increased the protein expression of occludin-5 in the spinal cord, as well as reducing the astrocyte reactivity and decreasing the microglial activation in the brain tissue of an experimental autoimmune encephalomyelitis mouse model of multiple sclerosis [196]. In a mice model of spinal cord injury, FMT attenuated the extravasation of Evans blue, increased the expression of ZO-1 and occludin, and decreased astrogliosis and microglial activation in the spinal cord [197].

6. Conclusions and Perspective

The gut microbiota contributes to the development and homeostasis of the nervous system. Gut dysbiosis is associated with cognitive impairments, cerebral abnormalities, and other complications in patients with advanced liver diseases. Gut-originated metabolites released from altered pathologic bacteria impair the integrity of BBB, devastate the brain microenvironment, and lead to the development of HE. The modulation of the gut microbiome composition by different agents has been revealed to have beneficial effects in the management of HE. Moreover, the manipulation of the intestinal microbiome to protect the beneficial species and decrease the harmful microbiome can improve and control HE through changes in the levels of circulating neurotoxic metabolites and products. Probiotics and prebiotics, antibiotics, and FMT improve the pathological changes in the structure of the BBB in numerous diseases, such as Alzheimer’s disease, Parkinson’s disease, depression, chronic stress, and stroke. The efficacy of these interventions still remains an open challenge in HE. However, the majority of animal and human studies have only focused on gut bacteria communities and underestimated the mycobiome diversity and virome composition of the gut microbiota in HE. In terms of future perspectives, it may be important to investigate the mycobiome and virome composition of the gut microbiome, along with bacteria, through multi-omics comprehensive metagenomics and metatranscriptomics studies to identify the specific microbial taxa and metabolites that are associated with HE and potentially serve as biomarkers for the diagnosis and monitoring of the disease. In conclusion, the relationship between gut dysbiosis and HE is a complex and dynamic process that requires further investigation. However, the growing body of evidence suggests that targeting gut dysbiosis represents a promising strategy for the prevention and treatment of HE.

Author Contributions

A.S. (Ali Shahbazi) and A.S. (Ali Sepehrinezhad) designed the study, performed the literature review, and drafted the manuscript. A.S. (Ali Sepehrinezhad) also drew and prepared the figures and corrected grammatical errors. S.S.N. performed the literature review, supervised the work, corrected grammatical errors, and drafted the manuscript. E.V., M.G. and S.M. drafted the manuscript. R.J. drafted the manuscript, and clinically edited the manuscript. F.S.L. critically and scientifically edited the revised manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Häussinger, D.; Dhiman, R.K.; Felipo, V.; Görg, B.; Jalan, R.; Kircheis, G.; Merli, M.; Montagnese, S.; Romero-Gomez, M.; Schnitzler, A.; et al. Hepatic encephalopathy. Nat. Rev. Dis. Prim. 2022, 8, 43. [Google Scholar] [CrossRef]
  2. Legaz, I.; Bolarin, J.M.; Campillo, J.A.; Moya, R.M.; Luna, A.; Osuna, E.; Minguela, A.; Sanchez-Bueno, F.; Alvarez, M.R.; Muro, M. Pretransplant ascites and encephalopathy and their influence on survival and liver graft rejection in alcoholic cirrhosis disease. Arch. Med. Sci. AMS 2021, 17, 682–693. [Google Scholar] [CrossRef] [PubMed]
  3. Montagnese, S.; Bajaj, J.S. Impact of Hepatic Encephalopathy in Cirrhosis on Quality-of-Life Issues. Drugs 2019, 79, 11–16. [Google Scholar] [CrossRef]
  4. Montagnese, S.; Rautou, P.-E.; Romero-Gómez, M.; Larsen, F.S.; Shawcross, D.L.; Thabut, D.; Vilstrup, H.; Weissenborn, K. EASL Clinical Practice Guidelines on the management of hepatic encephalopathy. J. Hepatol. 2022, 77, 807–824. [Google Scholar] [CrossRef]
  5. Hartmann, I.J.; Groeneweg, M.; Quero, J.C.; Beijeman, S.J.; de Man, R.A.; Hop, W.C.; Schalm, S.W. The prognostic significance of subclinical hepatic encephalopathy. Am. J. Gastroenterol. 2000, 95, 2029–2034. [Google Scholar] [CrossRef] [PubMed]
  6. Jesús Maldonado-Garza, H.; Vázquez-Elizondo, G.; Obed Gaytán-Torres, J.; Ricardo Flores-Rendón, Á.; Graciela Cárdenas-Sandoval, M.; Javier Bosques-Padilla, F. Prevalence of minimal hepatic encephalopathy in cirrhotic patients. Ann. Hepatol. 2011, 10, S40–S44. [Google Scholar] [CrossRef]
  7. Groeneweg, M.; Moerland, W.; Quero, J.C.; Hop, W.C.J.; Krabbe, P.F.; Schalm, S.W. Screening of subclinical hepatic encephalopathy. J. Hepatol. 2000, 32, 748–753. [Google Scholar] [CrossRef]
  8. Sharma, P.; Sharma, B.C.; Puri, V.; Sarin, S.K. Critical flicker frequency: Diagnostic tool for minimal hepatic encephalopathy. J. Hepatol. 2007, 47, 67–73. [Google Scholar] [CrossRef] [PubMed]
  9. Legaz, I.; Navarro-Noguera, E.; Bolarín, J.M.; García-Alonso, A.M.; Luna Maldonado, A.; Mrowiec, A.; Campillo, J.A.; Gimeno, L.; Moya-Quiles, R.; Álvarez-López, M.d.R.; et al. Epidemiology, Evolution, and Long-Term Survival of Alcoholic Cirrhosis Patients Submitted to Liver Transplantation in Southeastern Spain. Alcohol Clin. Exp. Res. 2016, 40, 794–805. [Google Scholar] [CrossRef]
  10. Romero-Gómez, M.; Boza, F.; García-Valdecasas, M.S.; García, E.; Aguilar-Reina, J. Subclinical hepatic encephalopathy predicts the development of overt hepatic encephalopathy. Am. J. Gastroenterol. 2001, 96, 2718–2723. [Google Scholar] [CrossRef] [PubMed]
  11. Watson, H.; Jepsen, P.; Wong, F.; Ginès, P.; Córdoba, J.; Vilstrup, H. Satavaptan treatment for ascites in patients with cirrhosis: A meta-analysis of effect on hepatic encephalopathy development. Metab. Brain Dis. 2013, 28, 301–305. [Google Scholar] [CrossRef]
  12. Vilstrup, H.; Amodio, P.; Bajaj, J.; Cordoba, J.; Ferenci, P.; Mullen, K.D.; Weissenborn, K.; Wong, P. Hepatic encephalopathy in chronic liver disease: 2014 Practice Guideline by the American Association for the Study of Liver Diseases and the European Association for the Study of the Liver. Hepatology 2014, 60, 715–735. [Google Scholar] [CrossRef] [PubMed]
  13. Amodio, P.; Del Piccolo, F.; Pettenò, E.; Mapelli, D.; Angeli, P.; Iemmolo, R.; Muraca, M.; Musto, C.; Gerunda, G.; Rizzo, C.; et al. Prevalence and prognostic value of quantified electroencephalogram (EEG) alterations in cirrhotic patients. J. Hepatol. 2001, 35, 37–45. [Google Scholar] [CrossRef] [PubMed]
  14. Bass, N.M.; Mullen, K.D.; Sanyal, A.; Poordad, F.; Neff, G.; Leevy, C.B.; Sigal, S.; Sheikh, M.Y.; Beavers, K.; Frederick, T.; et al. Rifaximin Treatment in Hepatic Encephalopathy. N. Engl. J. Med. 2010, 362, 1071–1081. [Google Scholar] [CrossRef]
  15. Chen, Z.; Ruan, J.; Li, D.; Wang, M.; Han, Z.; Qiu, W.; Wu, G. The Role of Intestinal Bacteria and Gut-Brain Axis in Hepatic Encephalopathy. Front. Cell. Infect. Microbiol. 2021, 10, 595759. [Google Scholar] [CrossRef]
  16. Fan, Y.; Pedersen, O. Gut microbiota in human metabolic health and disease. Nat. Rev. Microbiol. 2021, 19, 55–71. [Google Scholar] [CrossRef] [PubMed]
  17. Huttenhower, C.; Gevers, D.; Knight, R.; Abubucker, S.; Badger, J.H.; Chinwalla, A.T.; Creasy, H.H.; Earl, A.M.; FitzGerald, M.G.; Fulton, R.S.; et al. Structure, function and diversity of the healthy human microbiome. Nature 2012, 486, 207–214. [Google Scholar]
  18. Rowland, I.; Gibson, G.; Heinken, A.; Scott, K.; Swann, J.; Thiele, I.; Tuohy, K. Gut microbiota functions: Metabolism of nutrients and other food components. Eur. J. Nutr. 2018, 57, 1–24. [Google Scholar] [CrossRef]
  19. Zheng, D.; Liwinski, T.; Elinav, E. Interaction between microbiota and immunity in health and disease. Cell Res. 2020, 30, 492–506. [Google Scholar] [CrossRef]
  20. Cho, I.; Blaser, M.J. The human microbiome: At the interface of health and disease. Nat. Rev. Genet. 2012, 13, 260–270. [Google Scholar] [CrossRef]
  21. Bajaj, J.S.; Heuman, D.M.; Hylemon, P.B.; Sanyal, A.J.; White, M.B.; Monteith, P.; Noble, N.A.; Unser, A.B.; Daita, K.; Fisher, A.R.; et al. Altered profile of human gut microbiome is associated with cirrhosis and its complications. J. Hepatol. 2014, 60, 940–947. [Google Scholar] [CrossRef] [PubMed]
  22. Bajaj, J.S.; Betrapally, N.S.; Hylemon, P.B.; Heuman, D.M.; Daita, K.; White, M.B.; Unser, A.; Thacker, L.R.; Sanyal, A.J.; Kang, D.J.; et al. Salivary microbiota reflects changes in gut microbiota in cirrhosis with hepatic encephalopathy. Hepatology 2015, 62, 1260–1271. [Google Scholar] [CrossRef]
  23. Bajaj, J.S.; Hylemon, P.B.; Ridlon, J.M.; Heuman, D.M.; Daita, K.; White, M.B.; Monteith, P.; Noble, N.A.; Sikaroodi, M.; Gillevet, P.M. Colonic mucosal microbiome differs from stool microbiome in cirrhosis and hepatic encephalopathy and is linked to cognition and inflammation. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 303, G675–G685. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Zhai, H.; Geng, J.; Yu, R.; Ren, H.; Fan, H.; Shi, P. Large-scale survey of gut microbiota associated with MHE Via 16S rRNA-based pyrosequencing. Am. J. Gastroenterol. 2013, 108, 1601–1611. [Google Scholar] [CrossRef]
  25. Sung, C.M.; Lin, Y.F.; Chen, K.F.; Ke, H.M.; Huang, H.Y.; Gong, Y.N.; Tsai, W.S.; You, J.F.; Lu, M.J.; Cheng, H.T.; et al. Predicting Clinical Outcomes of Cirrhosis Patients with Hepatic Encephalopathy From the Fecal Microbiome. Cell. Mol. Gastroenterol. Hepatol. 2019, 8, 301–318.e2. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, J.Y.; Bajaj, J.S.; Wang, J.B.; Shang, J.; Zhou, X.M.; Guo, X.L.; Zhu, X.; Meng, L.N.; Jiang, H.X.; Mi, Y.Q.; et al. Lactulose improves cognition, quality of life, and gut microbiota in minimal hepatic encephalopathy: A multicenter, randomized controlled trial. J. Dig. Dis. 2019, 20, 547–556. [Google Scholar] [CrossRef]
  27. Bajaj, J.S.; Ridlon, J.M.; Hylemon, P.B.; Thacker, L.R.; Heuman, D.M.; Smith, S.; Sikaroodi, M.; Gillevet, P.M. Linkage of gut microbiome with cognition in hepatic encephalopathy. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G168–G175. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, Y.; Guo, J.; Qian, G.; Fang, D.; Shi, D.; Guo, L.; Li, L. Gut dysbiosis in acute-on-chronic liver failure and its predictive value for mortality. J. Gastroenterol. Hepatol. 2015, 30, 1429–1437. [Google Scholar] [CrossRef]
  29. Yukawa-Muto, Y.; Kamiya, T.; Fujii, H.; Mori, H.; Toyoda, A.; Sato, I.; Konishi, Y.; Hirayama, A.; Hara, E.; Fukuda, S.; et al. Distinct responsiveness to rifaximin in patients with hepatic encephalopathy depends on functional gut microbial species. Hepatol. Commun. 2022, 6, 2090–2104. [Google Scholar] [CrossRef] [PubMed]
  30. Hua, X.; Feng, H. Changes in intestinal microbiota of HBV-associated liver cirrhosis with/without hepatic encephalopathy. Medicine 2022, 101, e29935. [Google Scholar] [CrossRef]
  31. Lin, Y.; Yan, G.; Feng, F.; Wang, M.; Long, F. Characterization of intestinal microbiota and serum metabolites in patients with mild hepatic encephalopathy. Open Life Sci. 2022, 17, 139–154. [Google Scholar] [CrossRef] [PubMed]
  32. Bajaj, J.S.; Sikaroodi, M.; Shamsaddini, A.; Henseler, Z.; Santiago-Rodriguez, T.; Acharya, C.; Fagan, A.; Hylemon, P.B.; Fuchs, M.; Gavis, E.; et al. Interaction of bacterial metagenome and virome in patients with cirrhosis and hepatic encephalopathy. Gut 2021, 70, 1162–1173. [Google Scholar] [CrossRef]
  33. Kang, D.J.; Betrapally, N.S.; Ghosh, S.A.; Sartor, R.B.; Hylemon, P.B.; Gillevet, P.M.; Sanyal, A.J.; Heuman, D.M.; Carl, D.; Zhou, H.; et al. Gut microbiota drive the development of neuroinflammatory response in cirrhosis in mice. Hepatology 2016, 64, 1232–1248. [Google Scholar] [CrossRef]
  34. Yang, B.; Sun, T.; Chen, Y.; Xiang, H.; Xiong, J.; Bao, S. The Role of Gut Microbiota in Mice with Bile Duct Ligation-Evoked Cholestatic Liver Disease-Related Cognitive Dysfunction. Front. Microbiol. 2022, 13, 909461. [Google Scholar] [CrossRef] [PubMed]
  35. Đurašević, S.; Pejić, S.; Grigorov, I.; Nikolić, G.; Mitić-Ćulafić, D.; Dragićević, M.; Đorđević, J.; Todorović Vukotić, N.; Đorđević, N.; Todorović, A.; et al. Effects of C60 Fullerene on Thioacetamide-Induced Rat Liver Toxicity and Gut Microbiome Changes. Antioxidants 2021, 10, 911. [Google Scholar] [CrossRef]
  36. Zheng, Y.; Wang, J.; Wang, J.; Jiang, R.; Zhao, T. Gut microbiota combined with metabolomics reveal the mechanism of curcumol on liver fibrosis in mice. Biomed. Pharmacother. 2022, 152, 113204. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, J.; Zhang, D.; Zhu, B.; Wang, S.; Xu, Y.; Zhang, C.; Yang, H.; Wang, S.; Liu, P.; Qin, L.; et al. Rubus chingii Hu. unripe fruits extract ameliorates carbon tetrachloride-induced liver fibrosis and improves the associated gut microbiota imbalance. Chin. Med. 2022, 17, 56. [Google Scholar] [CrossRef] [PubMed]
  38. Cabrera-Rubio, R.; Patterson, A.M.; Cotter, P.D.; Beraza, N. Cholestasis induced by bile duct ligation promotes changes in the intestinal microbiome in mice. Sci. Rep. 2019, 9, 12324. [Google Scholar] [CrossRef]
  39. De Minicis, S.; Rychlicki, C.; Agostinelli, L.; Saccomanno, S.; Candelaresi, C.; Trozzi, L.; Mingarelli, E.; Facinelli, B.; Magi, G.; Palmieri, C.; et al. Dysbiosis contributes to fibrogenesis in the course of chronic liver injury in mice. Hepatology 2014, 59, 1738–1749. [Google Scholar] [CrossRef]
  40. Li, Y.; Lv, L.; Ye, J.; Fang, D.; Shi, D.; Wu, W.; Wang, Q.; Wu, J.; Yang, L.; Bian, X.; et al. Bifidobacterium adolescentis CGMCC 15058 alleviates liver injury, enhances the intestinal barrier and modifies the gut microbiota in d-galactosamine-treated rats. Appl. Microbiol. Biotechnol. 2019, 103, 375–393. [Google Scholar] [CrossRef]
  41. Terrón-Camero, L.C.; Gordillo-González, F.; Salas-Espejo, E.; Andrés-León, E. Comparison of Metagenomics and Metatranscriptomics Tools: A Guide to Making the Right Choice. Genes 2022, 13, 2280. [Google Scholar] [CrossRef]
  42. Liu, J.; Zhang, Q.; Dong, Y.-Q.; Yin, J.; Qiu, Y.-Q. Diagnostic accuracy of metagenomic next-generation sequencing in diagnosing infectious diseases: A meta-analysis. Sci. Rep. 2022, 12, 21032. [Google Scholar] [CrossRef] [PubMed]
  43. Albertsen, M. Long-read metagenomics paves the way toward a complete microbial tree of life. Nat. Methods 2023, 20, 30–31. [Google Scholar] [CrossRef]
  44. Madison, J.; LaBumbard, B.; Woodhams, D. Shotgun Metagenomic and 16S rRNA Gene Sequencing of Museum-Derived Rana pipiens Gut Microbiota Generate Differing Community Diversity Metrics Based on Commonly Used Analysis Pipelines. Authorea Prepr. 2023. [Google Scholar]
  45. Zuo, W.; Wang, B.; Bai, X.; Luan, Y.; Fan, Y.; Michail, S.; Sun, F. 16S rRNA and metagenomic shotgun sequencing data revealed consistent patterns of gut microbiome signature in pediatric ulcerative colitis. Sci. Rep. 2022, 12, 6421. [Google Scholar] [CrossRef]
  46. Morgan, X.C.; Huttenhower, C. Chapter 12: Human microbiome analysis. PLoS Comput. Biol. 2012, 8, e1002808. [Google Scholar] [CrossRef] [PubMed]
  47. Zampieri, G.; Campanaro, S.; Angione, C.; Treu, L. Metatranscriptomics-guided genome-scale metabolic modeling of microbial communities. Cell Rep. Methods 2023, 3, 100383. [Google Scholar] [CrossRef] [PubMed]
  48. Heravi, F.S.; Zakrzewski, M.; Vickery, K.; Malone, M.; Hu, H. Metatranscriptomic Analysis Reveals Active Bacterial Communities in Diabetic Foot Infections. Front. Microbiol. 2020, 11. [Google Scholar] [CrossRef] [PubMed]
  49. Aguiar-Pulido, V.; Huang, W.; Suarez-Ulloa, V.; Cickovski, T.; Mathee, K.; Narasimhan, G. Metagenomics, Metatranscriptomics, and Metabolomics Approaches for Microbiome Analysis. Evol. Bioinform. Online 2016, 12, 5–16. [Google Scholar] [CrossRef]
  50. Abbott, N.J.; Patabendige, A.A.; Dolman, D.E.; Yusof, S.R.; Begley, D.J. Structure and function of the blood-brain barrier. Neurobiol. Dis. 2010, 37, 13–25. [Google Scholar] [CrossRef]
  51. Obermeier, B.; Verma, A.; Ransohoff, R.M. The blood–brain barrier. Handb. Clin. Neurol. 2016, 133, 39–59. [Google Scholar]
  52. Tietz, S.; Engelhardt, B. Brain barriers: Crosstalk between complex tight junctions and adherens junctions. J. Cell Biol. 2015, 209, 493–506. [Google Scholar] [CrossRef] [PubMed]
  53. Luissint, A.-C.; Artus, C.; Glacial, F.; Ganeshamoorthy, K.; Couraud, P.-O. Tight junctions at the blood brain barrier: Physiological architecture and disease-associated dysregulation. Fluids Barriers CNS 2012, 9, 23. [Google Scholar] [CrossRef] [PubMed]
  54. Stamatovic, S.M.; Johnson, A.M.; Keep, R.F.; Andjelkovic, A.V. Junctional proteins of the blood-brain barrier: New insights into function and dysfunction. Tissue Barriers 2016, 4, e1154641. [Google Scholar] [CrossRef] [PubMed]
  55. Sahin, O.; Thompson, H.P.; Goodman, G.W.; Li, J.; Urayama, A. Mucopolysaccharidoses and the blood–brain barrier. Fluids Barriers CNS 2022, 19, 76. [Google Scholar] [CrossRef] [PubMed]
  56. Villaseñor, R.; Lampe, J.; Schwaninger, M.; Collin, L. Intracellular transport and regulation of transcytosis across the blood–brain barrier. Cell. Mol. Life Sci. 2019, 76, 1081–1092. [Google Scholar] [CrossRef]
  57. Lajoie, J.M.; Shusta, E.V. Targeting receptor-mediated transport for delivery of biologics across the blood-brain barrier. Annu. Rev. Pharmacol. Toxicol. 2015, 55, 613. [Google Scholar] [CrossRef]
  58. Ayloo, S.; Gu, C. Transcytosis at the blood–brain barrier. Curr. Opin. Neurobiol. 2019, 57, 32–38. [Google Scholar] [CrossRef]
  59. Deng, H.; Dutta, P.; Liu, J. Stochastic simulations of nanoparticle internalization through transferrin receptor dependent clathrin-mediated endocytosis. Biochim. Biophys. Acta-Gen. Subj. 2018, 1862, 2104–2111. [Google Scholar] [CrossRef]
  60. Sorets, A.G.; Rosch, J.C.; Duvall, C.L.; Lippmann, E.S. Caveolae-mediated transport at the injured blood-brain barrier as an underexplored pathway for central nervous system drug delivery. Curr. Opin. Chem. Eng. 2020, 30, 86–95. [Google Scholar] [CrossRef]
  61. Zhou, M.; Shi, S.X.; Liu, N.; Jiang, Y.; Karim, M.S.; Vodovoz, S.J.; Wang, X.; Zhang, B.; Dumont, A.S. Caveolae-mediated endothelial transcytosis across the blood-brain barrier in acute ischemic stroke. J. Clin. Med. 2021, 10, 3795. [Google Scholar] [CrossRef]
  62. De Boer, A.; van der Sandt, I.; Gaillard, P. The role of drug transporters at the blood-brain barrier. Annu. Rev. Pharmacol. Toxicol. 2003, 43, 629–656. [Google Scholar] [CrossRef] [PubMed]
  63. Pardridge, W.M. Blood–brain barrier endogenous transporters as therapeutic targets: A new model for small molecule CNS drug discovery. Expert Opin. Ther. Targets 2015, 19, 1059–1072. [Google Scholar] [CrossRef] [PubMed]
  64. Eng, M.E.; Imperio, G.E.; Bloise, E.; Matthews, S.G. ATP-binding cassette (ABC) drug transporters in the developing blood-brain barrier: Role in fetal brain protection. Cell. Mol. Life Sci. 2022, 79, 415. [Google Scholar] [CrossRef]
  65. Ohtsuki, S.; Terasaki, T. Contribution of carrier-mediated transport systems to the blood–brain barrier as a supporting and protecting interface for the brain; importance for CNS drug discovery and development. Pharm. Res. 2007, 24, 1745–1758. [Google Scholar] [CrossRef]
  66. Veys, K.; Fan, Z.; Ghobrial, M.; Bouché, A.; García-Caballero, M.; Vriens, K.; Conchinha, N.V.; Seuwen, A.; Schlegel, F.; Gorski, T. Role of the GLUT1 glucose transporter in postnatal CNS angiogenesis and blood-brain barrier integrity. Circ. Res. 2020, 127, 466–482. [Google Scholar] [CrossRef] [PubMed]
  67. Lee, N.-Y.; Kang, Y.-S. In vivo and in vitro evidence for brain uptake of 4-phenylbutyrate by the monocarboxylate transporter 1 (MCT1). Pharm. Res. 2016, 33, 1711–1722. [Google Scholar] [CrossRef]
  68. Schäfer, A.M.; Meyer zu Schwabedissen, H.E.; Grube, M. Expression and Function of Organic Anion Transporting Polypeptides in the Human Brain: Physiological and Pharmacological Implications. Pharmaceutics 2021, 13, 834. [Google Scholar] [CrossRef]
  69. Häfliger, P.; Charles, R.P. The L-Type Amino Acid Transporter LAT1-An Emerging Target in Cancer. Int. J. Mol. Sci. 2019, 20. [Google Scholar] [CrossRef]
  70. O’Donnell, M.E. Blood-brain barrier Na transporters in ischemic stroke. Adv. Pharmacol. 2014, 71, 113–146. [Google Scholar]
  71. Gerhardt, H.; Wolburg, H.; Redies, C. N-cadherin mediates pericytic-endothelial interaction during brain angiogenesis in the chicken. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2000, 218, 472–479. [Google Scholar] [CrossRef]
  72. Armulik, A.; Genové, G.; Mäe, M.; Nisancioglu, M.H.; Wallgard, E.; Niaudet, C.; He, L.; Norlin, J.; Lindblom, P.; Strittmatter, K. Pericytes regulate the blood–brain barrier. Nature 2010, 468, 557–561. [Google Scholar] [CrossRef] [PubMed]
  73. Ramsauer, M.; Krause, D.; Dermietzel, R. Angiogenesis of the blood-brain barrier in vitro and the function of cerebral pericytes. FASEB J. 2002, 16, 1274–1276. [Google Scholar] [CrossRef] [PubMed]
  74. Sweeney, M.D.; Ayyadurai, S.; Zlokovic, B.V. Pericytes of the neurovascular unit: Key functions and signaling pathways. Nat. Neurosci. 2016, 19, 771–783. [Google Scholar] [CrossRef] [PubMed]
  75. Abbott, N.J.; Rönnbäck, L.; Hansson, E. Astrocyte–endothelial interactions at the blood–brain barrier. Nat. Rev. Neurosci. 2006, 7, 41–53. [Google Scholar] [CrossRef]
  76. Kruyer, A.; Kalivas, P.W.; Scofield, M.D. Astrocyte regulation of synaptic signaling in psychiatric disorders. Neuropsychopharmacology 2023, 48, 21–36. [Google Scholar] [CrossRef]
  77. Gordon, G.R.; Howarth, C.; MacVicar, B.A. Bidirectional control of arteriole diameter by astrocytes. Exp. Physiol. 2011, 96, 393–399. [Google Scholar] [CrossRef]
  78. Bihlmaier, R.; Deffner, F.; Mattheus, U.; Neckel, P.H.; Hirt, B.; Mack, A.F. Aquaporin-1 and Aquaporin-4 Expression in Ependyma, Choroid Plexus and Surrounding Transition Zones in the Human Brain. Biomolecules 2023, 13, 212. [Google Scholar] [CrossRef]
  79. Jahncke, J.N.; Wright, K.M. The many roles of dystroglycan in nervous system development and function: Dystroglycan and neural circuit development. Dev. Dyn. 2023, 252, 61–80. [Google Scholar] [CrossRef]
  80. Nicchia, G.P.; Cogotzi, L.; Rossi, A.; Basco, D.; Brancaccio, A.; Svelto, M.; Frigeri, A. Expression of multiple AQP4 pools in the plasma membrane and their association with the dystrophin complex. J. Neurochem. 2008, 105, 2156–2165. [Google Scholar] [CrossRef]
  81. Weiss, N.; Miller, F.; Cazaubon, S.; Couraud, P.-O. The blood-brain barrier in brain homeostasis and neurological diseases. Biochim. Biophys. Acta-Biomembr. 2009, 1788, 842–857. [Google Scholar] [CrossRef] [PubMed]
  82. Segarra, M.; Aburto, M.R.; Acker-Palmer, A. Blood-brain barrier dynamics to maintain brain homeostasis. Trends Neurosci. 2021, 44, 393–405. [Google Scholar] [CrossRef] [PubMed]
  83. 83. Peng, X.; Luo, Z.; He, S.; Zhang, L.; Li, Y. Blood-Brain Barrier Disruption by Lipopolysaccharide and Sepsis-Associated Encephalopathy. Front. Cell Infect Microbiol. 2021, 11, 768108. [Google Scholar] [CrossRef] [PubMed]
  84. Li, Y.; Zhang, J.; Xu, P.; Sun, B.; Zhong, Z.; Liu, C.; Ling, Z.; Chen, Y.; Shu, N.; Zhao, K. Acute liver failure impairs function and expression of breast cancer-resistant protein (BCRP) at rat blood–brain barrier partly via ammonia-ROS-ERK 1/2 activation. J. Neurochem. 2016, 138, 282–294. [Google Scholar] [CrossRef]
  85. Ott, P.; Larsen, F.S. Blood-brain barrier permeability to ammonia in liver failure: A critical reappraisal. Neurochem. Int. 2004, 44, 185–198. [Google Scholar] [CrossRef] [PubMed]
  86. Cui, W.; Sun, C.-M.; Liu, P. Alterations of blood-brain barrier and associated factors in acute liver failure. Gastroenterol. Res. Pract. 2013, 2013, 841707. [Google Scholar] [CrossRef]
  87. Sepehrinezhad, A.; Zarifkar, A.; Namvar, G.; Shahbazi, A.; Williams, R. Astrocyte swelling in hepatic encephalopathy: Molecular perspective of cytotoxic edema. Metab. Brain Dis. 2020, 35, 559–578. [Google Scholar] [CrossRef] [PubMed]
  88. Erickson, M.A.; Dohi, K.; Banks, W.A. Neuroinflammation: A common pathway in CNS diseases as mediated at the blood-brain barrier. Neuroimmunomodulation 2012, 19, 121–130. [Google Scholar] [CrossRef]
  89. Banks, W.A.; Gray, A.M.; Erickson, M.A.; Salameh, T.S.; Damodarasamy, M.; Sheibani, N.; Meabon, J.S.; Wing, E.E.; Morofuji, Y.; Cook, D.G. Lipopolysaccharide-induced blood-brain barrier disruption: Roles of cyclooxygenase, oxidative stress, neuroinflammation, and elements of the neurovascular unit. J. Neuroinflammation 2015, 12, 1–15. [Google Scholar] [CrossRef]
  90. Kato, M.; Hughes, R.D.; Keays, R.T.; Williams, R. Electron microscopic study of brain capillaries in cerebral edema from fulminant hepatic failure. Hepatology 1992, 15, 1060–1066. [Google Scholar] [CrossRef]
  91. Lv, S.; Song, H.L.; Zhou, Y.; Li, L.X.; Cui, W.; Wang, W.; Liu, P. Tumour necrosis factor-alpha affects blood-brain barrier permeability and tight junction-associated occludin in acute liver failure. Liver Int. 2010, 30, 1198–1210. [Google Scholar] [CrossRef]
  92. Livingstone, A.S.; Potvin, M.; Goresky, C.A.; Finlayson, M.H.; Hinchey, E.J. Changes in the Blood-Brain Barrier in Hepatic Coma after Hepatectomy in the Rat. Gastroenterology 1977, 73, 697–704. [Google Scholar] [CrossRef] [PubMed]
  93. Zaki, A.E.; Ede, R.J.; Davis, M.; Williams, R. Experimental studies of blood brain barrier permeability in acute hepatic failure. Hepatology 1984, 4, 359–363. [Google Scholar] [CrossRef] [PubMed]
  94. Mossakowski, M.J.; Borowicz, J. Early electron-microscopic changes in hepatogenic encephalopathy induced by thioacetamide intoxication in rats. Neuropatol. Pol. 1985, 23, 375–387. [Google Scholar] [PubMed]
  95. Traber, P.G.; Dal Canto, M.; Ganger, D.R.; Blei, A.T. Electron microscopic evaluation of brain edema in rabbits with galactosamine-induced fulminant hepatic failure: Ultrastructure and integrity of the blood-brain barrier. Hepatology 1987, 7, 1272–1277. [Google Scholar] [CrossRef] [PubMed]
  96. Nguyen, J.H.; Yamamoto, S.; Steers, J.; Sevlever, D.; Lin, W.; Shimojima, N.; Castanedes-Casey, M.; Genco, P.; Golde, T.; Richelson, E.; et al. Matrix metalloproteinase-9 contributes to brain extravasation and edema in fulminant hepatic failure mice. J. Hepatol. 2006, 44, 1105–1114. [Google Scholar] [CrossRef]
  97. Chen, F.; Ohashi, N.; Li, W.; Eckman, C.; Nguyen, J.H. Disruptions of occludin and claudin-5 in brain endothelial cells in vitro and in brains of mice with acute liver failure. Hepatology 2009, 50, 1914–1923. [Google Scholar] [CrossRef]
  98. Kristiansen, R.G.; Lindal, S.; Myreng, K.; Revhaug, A.; Ytrebø, L.M.; Rose, C.F. Neuropathological changes in the brain of pigs with acute liver failure. Scand. J. Gastroenterol. 2010, 45, 935–943. [Google Scholar] [CrossRef]
  99. Wang, W.; Lv, S.; Zhou, Y.; Fu, J.; Li, C.; Liu, P. Tumor necrosis factor-α affects blood-brain barrier permeability in acetaminophen-induced acute liver failure. Eur. J. Gastroenterol. Hepatol. 2011, 23, 552–558. [Google Scholar] [CrossRef]
  100. Quinn, M.; McMillin, M.; Galindo, C.; Frampton, G.; Pae, H.Y.; DeMorrow, S. Bile acids permeabilize the blood brain barrier after bile duct ligation in rats via Rac1-dependent mechanisms. Dig. Liver Dis. 2014, 46, 527–534. [Google Scholar] [CrossRef]
  101. Chastre, A.; Bélanger, M.; Nguyen, B.N.; Butterworth, R.F. Lipopolysaccharide precipitates hepatic encephalopathy and increases blood-brain barrier permeability in mice with acute liver failure. Liver Int. 2014, 34, 353–361. [Google Scholar] [CrossRef]
  102. Faleiros, B.E.; Miranda, A.S.; Campos, A.C.; Gomides, L.F.; Kangussu, L.M.; Guatimosim, C.; Camargos, E.R.S.; Menezes, G.B.; Rachid, M.A.; Teixeira, A.L. Up-regulation of brain cytokines and chemokines mediates neurotoxicity in early acute liver failure by a mechanism independent of microglial activation. Brain Res. 2014, 1578, 49–59. [Google Scholar] [CrossRef]
  103. McMillin, M.A.; Frampton, G.A.; Seiwell, A.P.; Patel, N.S.; Jacobs, A.N.; DeMorrow, S. TGFβ1 exacerbates blood-brain barrier permeability in a mouse model of hepatic encephalopathy via upregulation of MMP9 and downregulation of claudin-5. Lab. Investig. 2015, 95, 903–913. [Google Scholar] [CrossRef]
  104. Thabut, D.; Mouri, S.; El Mourabit, H.; Morichon, R.; Wandum, D.; Lasnier, E.; Housset, C.; Weiss, N. Sodium benzoate and rifaximin are able to restore blood-brain barrier integrity in the cirrhotic rats. Intensive Care Med. Exp. 2015, 3, A691. [Google Scholar] [CrossRef]
  105. Grant, S.; McMillin, M.; Frampton, G.; Petrescu, A.D.; Williams, E.; Jaeger, V.; Kain, J.; DeMorrow, S. Direct Comparison of the Thioacetamide and Azoxymethane Models of Type A Hepatic Encephalopathy in Mice. Gene Expr. 2018, 18, 171–185. [Google Scholar] [CrossRef] [PubMed]
  106. Won, S.-M.; Oh, K.K.; Gupta, H.; Ganesan, R.; Sharma, S.P.; Jeong, J.-J.; Yoon, S.J.; Jeong, M.K.; Min, B.H.; Hyun, J.Y.; et al. The Link between Gut Microbiota and Hepatic Encephalopathy. Int. J. Mol. Sci. 2022, 23, 8999. [Google Scholar] [CrossRef] [PubMed]
  107. Saboo, K.; Petrakov, N.V.; Shamsaddini, A.; Fagan, A.; Gavis, E.A.; Sikaroodi, M.; McGeorge, S.; Gillevet, P.M.; Iyer, R.K.; Bajaj, J.S. Stool microbiota are superior to saliva in distinguishing cirrhosis and hepatic encephalopathy using machine learning. J. Hepatol. 2022, 76, 600–607. [Google Scholar] [CrossRef]
  108. Braniste, V.; Al-Asmakh, M.; Kowal, C.; Anuar, F.; Abbaspour, A.; Tóth, M.; Korecka, A.; Bakocevic, N.; Ng, L.G.; Kundu, P.; et al. The gut microbiota influences blood-brain barrier permeability in mice. Sci. Transl. Med. 2014, 6, 263ra158. [Google Scholar] [CrossRef] [PubMed]
  109. 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]
  110. Fröhlich, E.E.; Farzi, A.; Mayerhofer, R.; Reichmann, F.; Jačan, A.; Wagner, B.; Zinser, E.; Bordag, N.; Magnes, C.; Fröhlich, E.; et al. Cognitive impairment by antibiotic-induced gut dysbiosis: Analysis of gut microbiota-brain communication. Brain Behav. Immun. 2016, 56, 140–155. [Google Scholar] [CrossRef] [PubMed]
  111. Hoyles, L.; Snelling, T.; Umlai, U.-K.; Nicholson, J.K.; Carding, S.R.; Glen, R.C.; McArthur, S. Microbiome–host systems interactions: Protective effects of propionate upon the blood–brain barrier. Microbiome 2018, 6, 55. [Google Scholar] [CrossRef]
  112. Xu, Q.; Zhang, R.; Mu, Y.; Song, Y.; Hao, N.; Wei, Y.; Wang, Q.; Mackay, C.R. Propionate Ameliorates Alcohol-Induced Liver Injury in Mice via the Gut-Liver Axis: Focus on the Improvement of Intestinal Permeability. J. Agric. Food Chem. 2022, 70, 6084–6096. [Google Scholar] [CrossRef] [PubMed]
  113. Wu, Q.; Zhang, Y.; Zhang, Y.; Xia, C.; Lai, Q.; Dong, Z.; Kuang, W.; Yang, C.; Su, D.; Li, H.; et al. Potential effects of antibiotic-induced gut microbiome alteration on blood-brain barrier permeability compromise in rhesus monkeys. Ann. N. Y. Acad. Sci. 2020, 1470, 14–24. [Google Scholar] [CrossRef]
  114. Bloom, P.P.; Luévano, J.M.; Miller, K.J.; Chung, R.T. Deep stool microbiome analysis in cirrhosis reveals an association between short-chain fatty acids and hepatic encephalopathy. Ann. Hepatol. 2021, 25, 100333. [Google Scholar] [CrossRef] [PubMed]
  115. Xie, J.; Li, L.F.; Dai, T.Y.; Qi, X.; Wang, Y.; Zheng, T.Z.; Gao, X.Y.; Zhang, Y.J.; Ai, Y.; Ma, L.; et al. Short-Chain Fatty Acids Produced by Ruminococcaceae Mediate α-Linolenic Acid Promote Intestinal Stem Cells Proliferation. Mol. Nutr. Food Res. 2022, 66, e2100408. [Google Scholar] [CrossRef]
  116. Jin, M.; Kalainy, S.; Baskota, N.; Chiang, D.; Deehan, E.C.; McDougall, C.; Tandon, P.; Martínez, I.; Cervera, C.; Walter, J.; et al. Faecal microbiota from patients with cirrhosis has a low capacity to ferment non-digestible carbohydrates into short-chain fatty acids. Liver Int. 2019, 39, 1437–1447. [Google Scholar] [CrossRef] [PubMed]
  117. Pohl, K.; Moodley, P.; Dhanda, A. The effect of increasing intestinal short-chain fatty acid concentration on gut permeability and liver injury in the context of liver disease: A systematic review. J Gastroenterol. Hepatol. 2022, 37, 1498–1506. [Google Scholar] [CrossRef]
  118. Tsiaoussis, G.I.; Assimakopoulos, S.F.; Tsamandas, A.C.; Triantos, C.K.; Thomopoulos, K.C. Intestinal barrier dysfunction in cirrhosis: Current concepts in pathophysiology and clinical implications. World J. Hepatol. 2015, 7, 2058–2068. [Google Scholar] [CrossRef]
  119. Fukui, H. Leaky Gut and Gut-Liver Axis in Liver Cirrhosis: Clinical Studies Update. Gut Liver 2021, 15, 666–676. [Google Scholar] [CrossRef]
  120. Philipp, B. Bacterial degradation of bile salts. Appl. Microbiol. Biotechnol. 2011, 89, 903–915. [Google Scholar] [CrossRef]
  121. Guo, X.; Okpara, E.S.; Hu, W.; Yan, C.; Wang, Y.; Liang, Q.; Chiang, J.Y.; Han, S. Interactive Relationships between Intestinal Flora and Bile Acids. Int. J. Mol. Sci. 2022, 23, 8343. [Google Scholar] [CrossRef] [PubMed]
  122. Cipriani, S.; Mencarelli, A.; Chini, M.G.; Distrutti, E.; Renga, B.; Bifulco, G.; Baldelli, F.; Donini, A.; Fiorucci, S. The bile acid receptor GPBAR-1 (TGR5) modulates integrity of intestinal barrier and immune response to experimental colitis. PLoS ONE 2011, 6, e25637. [Google Scholar] [CrossRef]
  123. Gadaleta, R.M.; van Erpecum, K.J.; Oldenburg, B.; Willemsen, E.C.; Renooij, W.; Murzilli, S.; Klomp, L.W.; Siersema, P.D.; Schipper, M.E.; Danese, S.; et al. Farnesoid X receptor activation inhibits inflammation and preserves the intestinal barrier in inflammatory bowel disease. Gut 2011, 60, 463–472. [Google Scholar] [CrossRef]
  124. Acharya, C.; Bajaj, J.S. Chronic Liver Diseases and the Microbiome-Translating Our Knowledge of Gut Microbiota to Management of Chronic Liver Disease. Gastroenterology 2021, 160, 556–572. [Google Scholar] [CrossRef]
  125. Úbeda, M.; Lario, M.; Muñoz, L.; Borrero, M.J.; Rodríguez-Serrano, M.; Sánchez-Díaz, A.M.; Del Campo, R.; Lledó, L.; Pastor, Ó.; García-Bermejo, L.; et al. Obeticholic acid reduces bacterial translocation and inhibits intestinal inflammation in cirrhotic rats. J. Hepatol. 2016, 64, 1049–1057. [Google Scholar] [CrossRef]
  126. Neale, G.; Lewis, B.; Weaver, V.; Panveliwalla, D. Serum bile acids in liver disease. Gut 1971, 12, 145–152. [Google Scholar] [CrossRef] [PubMed]
  127. Kim, M.J.; Suh, D.J. Profiles of serum bile acids in liver diseases. Korean J. Intern. Med. 1986, 1, 37–42. [Google Scholar] [CrossRef] [PubMed]
  128. Balazs, I.; Horvath, A.; Leber, B.; Feldbacher, N.; Sattler, W.; Rainer, F.; Fauler, G.; Vermeren, S.; Stadlbauer, V. Serum bile acids in liver cirrhosis promote neutrophil dysfunction. Clin. Transl. Med. 2022, 12, e735. [Google Scholar] [CrossRef]
  129. Kumar, H.; Kawai, T.; Akira, S. Pathogen recognition by the innate immune system. Int. Rev. Immunol. 2011, 30, 16–34. [Google Scholar] [CrossRef]
  130. Negi, S.; Das, D.K.; Pahari, S.; Nadeem, S.; Agrewala, J.N. Potential Role of Gut Microbiota in Induction and Regulation of Innate Immune Memory. Front. Immunol. 2019, 10, 2441. [Google Scholar] [CrossRef]
  131. Kawai, T.; Akira, S. The role of pattern-recognition receptors in innate immunity: Update on Toll-like receptors. Nat. Immunol. 2010, 11, 373–384. [Google Scholar] [CrossRef]
  132. Gómez-Hurtado, I.; Such, J.; Francés, R. Microbiome and bacterial translocation in cirrhosis. Gastroenterol. Y Hepatol. 2016, 39, 687–696. [Google Scholar] [CrossRef] [PubMed]
  133. Rayes, N.; Seehofer, D.; Müller, A.R.; Hansen, S.; Bengmark, S.; Neuhaus, P. Influence of probiotics and fibre on the incidence of bacterial infections following major abdominal surgery—Results of a prospective trial. Z. Gastroenterol. 2002, 40, 869–876. [Google Scholar] [CrossRef]
  134. Such, J.; Francés, R.; Muñoz, C.; Zapater, P.; Casellas, J.A.; Cifuentes, A.; Rodríguez-Valera, F.; Pascual, S.; Sola-Vera, J.; Carnicer, F.; et al. Detection and identification of bacterial DNA in patients with cirrhosis and culture-negative, nonneutrocytic ascites. Hepatology 2002, 36, 135–141. [Google Scholar] [CrossRef] [PubMed]
  135. Mortensen, C.; Jensen, J.S.; Hobolth, L.; Dam-Larsen, S.; Madsen, B.S.; Andersen, O.; Møller, S.; Bendtsen, F. Association of markers of bacterial translocation with immune activation in decompensated cirrhosis. Eur. J. Gastroenterol. Hepatol. 2014, 26, 1360–1366. [Google Scholar] [CrossRef]
  136. Liu, L.; Zhang, C.; Hu, Y.; Zhou, L.; Tan, Q. Changes in gut toll-like receptor-4 and nod-like receptor family pyrin domain containing-3 innate pathways in liver cirrhosis rats with bacterial translocation. Clin. Res. Hepatol. Gastroenterol. 2016, 40, 575–583. [Google Scholar] [CrossRef] [PubMed]
  137. Muñoz, L.; José Borrero, M.; Ubeda, M.; Lario, M.; Díaz, D.; Francés, R.; Monserrat, J.; Pastor, O.; Aguado-Fraile, E.; Such, J.; et al. Interaction between intestinal dendritic cells and bacteria translocated from the gut in rats with cirrhosis. Hepatology 2012, 56, 1861–1869. [Google Scholar] [CrossRef]
  138. Bellot, P.; Francés, R.; Such, J. Pathological bacterial translocation in cirrhosis: Pathophysiology, diagnosis and clinical implications. Liver Int. 2013, 33, 31–39. [Google Scholar] [CrossRef] [PubMed]
  139. Männistö, V.; Färkkilä, M.; Pussinen, P.; Jula, A.; Männistö, S.; Lundqvist, A.; Valsta, L.; Salomaa, V.; Perola, M.; Åberg, F. Serum lipopolysaccharides predict advanced liver disease in the general population. JHEP Rep. Innov. Hepatol. 2019, 1, 345–352. [Google Scholar] [CrossRef]
  140. Meng, J.; Wang, Q.; Liu, K.; Yang, S.; Fan, X.; Liu, B.; He, C.; Wu, X. Systemic and Splanchnic Lipopolysaccharide and Endothelin-1 Plasma Levels in Liver Cirrhosis before and after Transjugular Intrahepatic Portosystemic Shunt. Gastroenterol. Res. Pract. 2016, 2016, 8341030. [Google Scholar] [CrossRef]
  141. Li, X.; Jiang, S.; Tapping, R.I. Toll-like receptor signaling in cell proliferation and survival. Cytokine 2010, 49, 1–9. [Google Scholar] [CrossRef]
  142. Gay, N.J.; Symmons, M.F.; Gangloff, M.; Bryant, C.E. Assembly and localization of Toll-like receptor signalling complexes. Nat. Rev. Immunol. 2014, 14, 546–558. [Google Scholar] [CrossRef]
  143. Vidya, M.K.; Kumar, V.G.; Sejian, V.; Bagath, M.; Krishnan, G.; Bhatta, R. Toll-like receptors: Significance, ligands, signaling pathways, and functions in mammals. Int. Rev. Immunol. 2018, 37, 20–36. [Google Scholar] [CrossRef]
  144. Ozato, K.; Tsujimura, H.; Tamura, T. Toll-like receptor signaling and regulation of cytokine gene expression in the immune system. BioTechniques 2002, 33, S66–S75. [Google Scholar] [CrossRef]
  145. Grassin-Delyle, S.; Abrial, C.; Salvator, H.; Brollo, M.; Naline, E.; Devillier, P. The Role of Toll-Like Receptors in the Production of Cytokines by Human Lung Macrophages. J. Innate Immun. 2020, 12, 63–73. [Google Scholar] [CrossRef] [PubMed]
  146. Su, G.L. Lipopolysaccharides in liver injury: Molecular mechanisms of Kupffer cell activation. Am. J. Physiol.-Gastrointest. Liver Physiol. 2002, 283, G256–G265. [Google Scholar] [CrossRef] [PubMed]
  147. Jiang, Z.; Meng, Y.; Bo, L.; Wang, C.; Bian, J.; Deng, X. Sophocarpine Attenuates LPS-Induced Liver Injury and Improves Survival of Mice through Suppressing Oxidative Stress, Inflammation, and Apoptosis. Mediat. Inflamm. 2018, 2018, 5871431. [Google Scholar] [CrossRef]
  148. Wang, X.; Yu, J.-y.; Sun, Y.; Wang, H.; Shan, H.; Wang, S. Baicalin protects LPS-induced blood–brain barrier damage and activates Nrf2-mediated antioxidant stress pathway. Int. Immunopharmacol. 2021, 96, 107725. [Google Scholar] [CrossRef]
  149. Sewal, R.K.; Modi, M.; Saikia, U.N.; Chakrabarti, A.; Medhi, B. Increase in seizure susceptibility in sepsis like condition explained by spiking cytokines and altered adhesion molecules level with impaired blood brain barrier integrity in experimental model of rats treated with lipopolysaccharides. Epilepsy Res. 2017, 135, 176–186. [Google Scholar] [CrossRef] [PubMed]
  150. Jangula, A.; Murphy, E.J. Lipopolysaccharide-induced blood brain barrier permeability is enhanced by alpha-synuclein expression. Neurosci. Lett. 2013, 551, 23–27. [Google Scholar] [CrossRef] [PubMed]
  151. Zaki, A.E.; Wardle, E.N.; Canalese, J.; Ede, R.J.; Williams, R. Potential toxins of acute liver failure and their effects on blood-brain barrier permeability. Experientia 1983, 39, 988–991. [Google Scholar] [CrossRef] [PubMed]
  152. Roberts, R.A.; Ganey, P.E.; Ju, C.; Kamendulis, L.M.; Rusyn, I.; Klaunig, J.E. Role of the Kupffer Cell in Mediating Hepatic Toxicity and Carcinogenesis. Toxicol. Sci. 2007, 96, 2–15. [Google Scholar] [CrossRef] [PubMed]
  153. Thurman, R.G.; Bradford, B.U.; Iimuro, Y.; Knecht, K.T.; Connor, H.D.; Adachi, Y.; Wall, C.; Arteel, G.E.; Raleigh, J.A.; Forman, D.T.; et al. Role of Kupffer Cells, Endotoxin and Free Radicals in Hepatotoxicity Due to Prolonged Alcohol Consumption: Studies in Female and Male Rats. J. Nutr. 1997, 127, 903S–906S. [Google Scholar] [CrossRef]
  154. McCuskey, P.A.; McCuskey, R.S. Electron microscopic study of the effects of endotoxin on the cells of the hepatic sinusoid in normal and BCG sensitized mice. Histol. Histopathol. 1991, 6, 353–362. [Google Scholar] [PubMed]
  155. McMillin, M.; Galindo, C.; Pae, H.Y.; Frampton, G.; Di Patre, P.L.; Quinn, M.; Whittington, E.; DeMorrow, S. Gli1 activation and protection against hepatic encephalopathy is suppressed by circulating transforming growth factor β1 in mice. J. Hepatol. 2014, 61, 1260–1266. [Google Scholar] [CrossRef]
  156. Dhanda, S.; Sandhir, R. Blood-Brain Barrier Permeability Is Exacerbated in Experimental Model of Hepatic Encephalopathy via MMP-9 Activation and Downregulation of Tight Junction Proteins. Mol. Neurobiol. 2018, 55, 3642–3659. [Google Scholar] [CrossRef] [PubMed]
  157. Jefferson, B.; Ali, M.; Grant, S.; Frampton, G.; Ploof, M.; Andry, S.; DeMorrow, S.; McMillin, M. Thrombospondin-1 Exacerbates Acute Liver Failure and Hepatic Encephalopathy Pathology in Mice by Activating Transforming Growth Factor β1. Am. J. Pathol. 2020, 190, 347–357. [Google Scholar] [CrossRef] [PubMed]
  158. Azhari, H.; Swain, M.G. Role of Peripheral Inflammation in Hepatic Encephalopathy. J. Clin. Exp. Hepatol. 2018, 8, 281–285. [Google Scholar] [CrossRef]
  159. Balzano, T.; Leone, P.; Ivaylova, G.; Castro, M.C.; Reyes, L.; Ramón, C.; Malaguarnera, M.; Llansola, M.; Felipo, V. Rifaximin Prevents T-Lymphocytes and Macrophages Infiltration in Cerebellum and Restores Motor Incoordination in Rats with Mild Liver Damage. Biomedicines 2021, 9, 1002. [Google Scholar] [CrossRef]
  160. Furness, J.B.; Callaghan, B.P.; Rivera, L.R.; Cho, H.J. The enteric nervous system and gastrointestinal innervation: Integrated local and central control. Adv. Exp. Med. Biol. 2014, 817, 39–71. [Google Scholar]
  161. Needham, B.D.; Kaddurah-Daouk, R.; Mazmanian, S.K. Gut microbial molecules in behavioural and neurodegenerative conditions. Nat. Rev. Neurosci. 2020, 21, 717–731. [Google Scholar] [CrossRef]
  162. Barajon, I.; Serrao, G.; Arnaboldi, F.; Opizzi, E.; Ripamonti, G.; Balsari, A.; Rumio, C. Toll-like receptors 3, 4, and 7 are expressed in the enteric nervous system and dorsal root ganglia. J. Histochem. Cytochem. 2009, 57, 1013–1023. [Google Scholar] [CrossRef]
  163. Burgueño, J.F.; Barba, A.; Eyre, E.; Romero, C.; Neunlist, M.; Fernández, E. TLR2 and TLR9 modulate enteric nervous system inflammatory responses to lipopolysaccharide. J. Neuroinflamm. 2016, 13, 187. [Google Scholar] [CrossRef] [PubMed]
  164. Geng, Z.H.; Zhu, Y.; Li, Q.L.; Zhao, C.; Zhou, P.H. Enteric Nervous System: The Bridge Between the Gut Microbiota and Neurological Disorders. Front. Aging Neurosci. 2022, 14, 810483. [Google Scholar] [CrossRef] [PubMed]
  165. Nøhr, M.K.; Pedersen, M.H.; Gille, A.; Egerod, K.L.; Engelstoft, M.S.; Husted, A.S.; Sichlau, R.M.; Grunddal, K.V.; Seier Poulsen, S.; Han, S.; et al. GPR41/FFAR3 and GPR43/FFAR2 as Cosensors for Short-Chain Fatty Acids in Enteroendocrine Cells vs FFAR3 in Enteric Neurons and FFAR2 in Enteric Leukocytes. Endocrinology 2013, 154, 3552–3564. [Google Scholar] [CrossRef] [PubMed]
  166. Soret, R.; Chevalier, J.; de Coppet, P.; Poupeau, G.; Derkinderen, P.; Segain, J.P.; Neunlist, M. Short-chain fatty acids regulate the enteric neurons and control gastrointestinal motility in rats. Gastroenterology 2010, 138, 1772–1782. [Google Scholar] [CrossRef]
  167. Linan-Rico, A.; Ochoa-Cortes, F.; Beyder, A.; Soghomonyan, S.; Zuleta-Alarcon, A.; Coppola, V.; Christofi, F.L. Mechanosensory Signaling in Enterochromaffin Cells and 5-HT Release: Potential Implications for Gut Inflammation. Front. Neurosci. 2016, 10, 564. [Google Scholar] [CrossRef]
  168. Viola, M.F.; Boeckxstaens, G. Muscularis macrophages: Trained guardians of enteric neurons. Cell Res. 2022, 32, 229–230. [Google Scholar] [CrossRef]
  169. Nijhuis, L.E.; Olivier, B.J.; de Jonge, W.J. Neurogenic regulation of dendritic cells in the intestine. Biochem. Pharmacol. 2010, 80, 2002–2008. [Google Scholar] [CrossRef]
  170. Progatzky, F.; Pachnis, V. The role of enteric glia in intestinal immunity. Curr. Opin. Immunol. 2022, 77, 102183. [Google Scholar] [CrossRef]
  171. Xu, X.; Chen, R.; Zhan, G.; Wang, D.; Tan, X.; Xu, H. Enterochromaffin Cells: Sentinels to Gut Microbiota in Hyperalgesia? Front. Cell. Infect. Microbiol. 2021, 11, 760076. [Google Scholar] [CrossRef]
  172. Arora, T.; Vanslette, A.M.; Hjorth, S.A.; Bäckhed, F. Microbial regulation of enteroendocrine cells. Med 2021, 2, 553–570. [Google Scholar] [CrossRef] [PubMed]
  173. Li, Z.; Chalazonitis, A.; Huang, Y.Y.; Mann, J.J.; Margolis, K.G.; Yang, Q.M.; Kim, D.O.; Côté, F.; Mallet, J.; Gershon, M.D. Essential roles of enteric neuronal serotonin in gastrointestinal motility and the development/survival of enteric dopaminergic neurons. J. Neurosci. 2011, 31, 8998–9009. [Google Scholar] [CrossRef] [PubMed]
  174. De Vadder, F.; Grasset, E.; Mannerås Holm, L.; Karsenty, G.; Macpherson, A.J.; Olofsson, L.E.; Bäckhed, F. Gut microbiota regulates maturation of the adult enteric nervous system via enteric serotonin networks. Proc. Natl. Acad. Sci. USA 2018, 115, 6458–6463. [Google Scholar] [CrossRef]
  175. Holzer, P.; Farzi, A. Neuropeptides and the microbiota-gut-brain axis. Adv. Exp. Med. Biol. 2014, 817, 195–219. [Google Scholar] [PubMed]
  176. Bauché, D.; Marie, J.C. Transforming growth factor β: A master regulator of the gut microbiota and immune cell interactions. Clin. Transl. Immunol. 2017, 6, e136. [Google Scholar] [CrossRef] [PubMed]
  177. Eckburg, P.B.; Bik, E.M.; Bernstein, C.N.; Purdom, E.; Dethlefsen, L.; Sargent, M.; Gill, S.R.; Nelson, K.E.; Relman, D.A. Diversity of the human intestinal microbial flora. Science 2005, 308, 1635–1638. [Google Scholar] [CrossRef]
  178. Zoetendal, E.G.; von Wright, A.; Vilpponen-Salmela, T.; Ben-Amor, K.; Akkermans, A.D.; de Vos, W. Mucosa-associated bacteria in the human gastrointestinal tract are uniformly distributed along the colon and differ from the community recovered from feces. Appl. Environ. Microbiol. 2002, 68, 3401–3407. [Google Scholar] [CrossRef]
  179. Belzer, C.; de Vos, W.M. Microbes inside—from diversity to function: The case of Akkermansia. ISME J. 2012, 6, 1449–1458. [Google Scholar] [CrossRef]
  180. O’Hara, A.M.; Shanahan, F. The gut flora as a forgotten organ. EMBO Rep. 2006, 7, 688–693. [Google Scholar] [CrossRef]
  181. Ahn, J.-S.; Lkhagva, E.; Jung, S.; Kim, H.-J.; Chung, H.-J.; Hong, S.-T. Fecal Microbiome Does Not Represent Whole Gut Microbiome. Cell. Microbiol. 2023, 2023, 6868417. [Google Scholar] [CrossRef]
  182. Patidar, K.R.; Bajaj, J.S. Antibiotics for the treatment of hepatic encephalopathy. Metab. Brain Dis. 2013, 28, 307–312. [Google Scholar] [CrossRef]
  183. Tapper, E.B.; Essien, U.R.; Zhao, Z.; Ufere, N.N.; Parikh, N.D. Racial and ethnic disparities in rifaximin use and subspecialty referrals for patients with hepatic encephalopathy in the United States. J. Hepatol. 2022, 77, 377–382. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, X.; Du, Z.R.; Wang, X.; Sun, X.R.; Zhao, Q.; Zhao, F.; Wong, W.T.; Wong, K.H.; Dong, X.-L. Polymannuronic acid prebiotic plus Lacticaseibacillus rhamnosus GG probiotic as a novel synbiotic promoted their separate neuroprotection against Parkinson’s disease. Food Res. Int. 2022, 155, 111067. [Google Scholar] [CrossRef]
  185. Zhang, Y.; Ding, N.; Hao, X.; Zhao, J.; Zhao, Y.; Li, Y.; Li, Z. Manual acupuncture benignly regulates blood-brain barrier disruption and reduces lipopolysaccharide loading and systemic inflammation, possibly by adjusting the gut microbiota. Front. Aging Neurosci. 2022, 14, 1018371. [Google Scholar] [CrossRef]
  186. Birmann, P.T.; Casaril, A.M.; Pesarico, A.P.; Caballero, P.S.; Smaniotto, T.Â.; Rodrigues, R.R.; Moreira, Â.N.; Conceição, F.R.; Sousa, F.S.S.; Collares, T.; et al. Komagataella pastoris KM71H modulates neuroimmune and oxidative stress parameters in animal models of depression: A proposal for a new probiotic with antidepressant-like effect. Pharmacol. Res. 2021, 171, 105740. [Google Scholar] [CrossRef]
  187. Ranuh, R.G.; Athiyyah, A.F.; Darma, A.; Riawan, W.; Gunawan, P.I.; Surono, I.S.; Sudarmo, S.M. Probiotic Lactobacillus plantarum IS-10506, Expression of Glial Fibrillary Acidic Protein and Platelet Endothelial Cell Adhesion Molecule-1 by Astrocytes and Endothelial Integrity: The Importance of Intestinal Microbiota as Blood Brain-Barrier Stabilizer. Int. J. Probiotics Prebiotics 2022, 17, 1–6. [Google Scholar] [CrossRef] [PubMed]
  188. Dhaliwal, J.; Singh, D.P.; Singh, S.; Pinnaka, A.K.; Boparai, R.K.; Bishnoi, M.; Kondepudi, K.K.; Chopra, K. Lactobacillus plantarum MTCC 9510 supplementation protects from chronic unpredictable and sleep deprivation-induced behaviour, biochemical and selected gut microbial aberrations in mice. J. Appl. Microbiol. 2018, 125, 257–269. [Google Scholar] [CrossRef]
  189. Zhang, Z.; Li, J.; Jiang, S.; Xu, M.; Ma, T.; Sun, Z.; Zhang, J. Lactobacillus fermentum HNU312 alleviated oxidative damage and behavioural abnormalities during brain development in early life induced by chronic lead exposure. Ecotoxicol. Environ. Saf. 2023, 251, 114543. [Google Scholar] [CrossRef]
  190. Du, J.; Yang, X.; Yu, D.; Xue, L. Probiotics Improve Flora Alternation and Memory Deficits in Aged Mice. FASEB J. 2019, 33, 516–517. [Google Scholar] [CrossRef]
  191. Saha, P.; Skidmore, P.T.; Holland, L.A.; Mondal, A.; Bose, D.; Seth, R.K.; Sullivan, K.; Janulewicz, P.A.; Horner, R.; Klimas, N. Andrographolide attenuates gut-brain-axis associated pathology in gulf war illness by modulating bacteriome-virome associated inflammation and microglia-neuron proinflammatory crosstalk. Brain Sci. 2021, 11, 905. [Google Scholar] [CrossRef] [PubMed]
  192. Hong, C.T.; Chan, L.; Chen, K.Y.; Lee, H.H.; Huang, L.K.; Yang, Y.S.H.; Liu, Y.R.; Hu, C.J. Rifaximin Modifies Gut Microbiota and Attenuates Inflammation in Parkinson’s Disease: Preclinical and Clinical Studies. Cells 2022, 11, 3468. [Google Scholar] [CrossRef] [PubMed]
  193. Kanamaru, H.; Kawakita, F.; Nishikawa, H.; Nakano, F.; Asada, R.; Suzuki, H. Clarithromycin Ameliorates Early Brain Injury after Subarachnoid Hemorrhage via Suppressing Periostin-Related Pathways in Mice. Neurotherapeutics 2021, 18, 1880–1890. [Google Scholar] [CrossRef]
  194. Luo, A.; Li, S.; Wang, X.; Xie, Z.; Li, S.; Hua, D. Cefazolin Improves Anesthesia and Surgery-Induced Cognitive Impairments by Modulating Blood-Brain Barrier Function, Gut Bacteria and Short Chain Fatty Acids. Front. Aging Neurosci. 2021, 13, 748637. [Google Scholar] [CrossRef] [PubMed]
  195. Zhao, Z.; Ning, J.; Bao, X.Q.; Shang, M.; Ma, J.; Li, G.; Zhang, D. Fecal microbiota transplantation protects rotenone-induced Parkinson’s disease mice via suppressing inflammation mediated by the lipopolysaccharide-TLR4 signaling pathway through the microbiota-gut-brain axis. Microbiome 2021, 9, 226. [Google Scholar] [CrossRef]
  196. Li, K.; Wei, S.; Hu, L.; Yin, X.; Mai, Y.; Jiang, C.; Peng, X.; Cao, X.; Huang, Z.; Zhou, H. Protection of Fecal Microbiota Transplantation in a Mouse Model of Multiple Sclerosis. Mediat. Inflamm 2020, 2020, 2058272. [Google Scholar] [CrossRef]
  197. Jing, Y.; Bai, F.; Wang, L.; Yang, D.; Yan, Y.; Wang, Q.; Zhu, Y.; Yu, Y.; Chen, Z. Fecal Microbiota Transplantation Exerts Neuroprotective Effects in a Mouse Spinal Cord Injury Model by Modulating the Microenvironment at the Lesion Site. Microbiol. Spectr. 2022, 10, e0017722. [Google Scholar] [CrossRef] [PubMed]
  198. Sun, N.; Hu, H.; Wang, F.; Li, L.; Zhu, W.; Shen, Y.; Xiu, J.; Xu, Q. Antibiotic-induced microbiome depletion in adult mice disrupts blood-brain barrier and facilitates brain infiltration of monocytes after bone-marrow transplantation. Brain Behav. Immun. 2021, 92, 102–114. [Google Scholar] [CrossRef]
  199. Roberfroid, M.; Gibson, G.R.; Hoyles, L.; McCartney, A.L.; Rastall, R.; Rowland, I.; Wolvers, D.; Watzl, B.; Szajewska, H.; Stahl, B. Prebiotic effects: Metabolic and health benefits. Br. J. Nutr. 2010, 104, S1–S63. [Google Scholar] [CrossRef]
  200. Zendeboodi, F.; Khorshidian, N.; Mortazavian, A.M.; da Cruz, A.G. Probiotic: Conceptualization from a new approach. Curr. Opin. Food Sci. 2020, 32, 103–123. [Google Scholar] [CrossRef]
  201. Cui, Y.; Xu, L.; Wang, F.; Wang, Z.; Tong, X.; Yan, H. Orally Administered Brain Protein Combined with Probiotics Increases Treg Differentiation to Reduce Secondary Inflammatory Damage Following Craniocerebral Trauma. Front. Immunol. 2022, 13, 928343. [Google Scholar] [CrossRef] [PubMed]
  202. Yang, X.; Yu, D.; Xue, L.; Li, H.; Du, J. Probiotics modulate the microbiota-gut-brain axis and improve memory deficits in aged SAMP8 mice. Acta Pharm. Sin. B 2020, 10, 475–487. [Google Scholar] [CrossRef]
  203. Vindigni, S.M.; Surawicz, C.M. Fecal microbiota transplantation. Gastroenterol. Clin. 2017, 46, 171–185. [Google Scholar] [CrossRef] [PubMed]
  204. Vaughn, B.P.; Rank, K.M.; Khoruts, A. Fecal microbiota transplantation: Current status in treatment of GI and liver disease. Clin. Gastroenterol. Hepatol. 2019, 17, 353–361. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The structure of blood–brain barrier. The BBB is formed by endothelial cells, pericytes, basement membrane and perivascular astrocyte end-feet. This barrier strongly restricts paracellular transmission due to the presence of tight junctions between endothelial cells. Tight junctions include two transmembrane proteins claudin and occludin and an intracellular protein zonula occludens-1. AMT: Absorptive-mediated transport; IGF1R: Insulin-like growth factor1 receptor; JAMs: Junctional adhesion molecules; RMT: Receptor-mediated transport; ZO-1: zonula occludens-1. Created with “BioRender.com. (accessed on 2 February 2023)”.
Figure 1. The structure of blood–brain barrier. The BBB is formed by endothelial cells, pericytes, basement membrane and perivascular astrocyte end-feet. This barrier strongly restricts paracellular transmission due to the presence of tight junctions between endothelial cells. Tight junctions include two transmembrane proteins claudin and occludin and an intracellular protein zonula occludens-1. AMT: Absorptive-mediated transport; IGF1R: Insulin-like growth factor1 receptor; JAMs: Junctional adhesion molecules; RMT: Receptor-mediated transport; ZO-1: zonula occludens-1. Created with “BioRender.com. (accessed on 2 February 2023)”.
Biomedicines 11 01272 g001
Figure 2. Proposed relationship between the gut microbiome and the structure of blood-brain barrier in health and advanced liver disease. Almost one to two years after birth gut microbiome community is created and its diversity is altered under the influence of age, diet, lifestyle, and geography during life. In healthy, normal gut microbiota-derived metabolites enter circulation and reach the BBB and maintain the integrity of barrier. Advanced liver disease and cirrhosis altered gut microbiota composition. Altered microbiota produces different metabolites that disrupt the intestinal barrier and result in bacterial translocation. Metabolites and products of altered bacteria (i.e., DNA, LPS, etc.) trigger the innate immune system that initiates the production of pro-inflammatory cytokines. Systemic inflammation activates sinusoidal kupffer cells and causes hepatocyte injury that along with gut-derived products and metabolites impair the integrity of the BBB. Leukocyte infiltration, glial activation, oxidative stress, neuroinflammation, and neurodegeneration are the main consequences of BBB damage. BBB: Blood-brain barrier. Created with “BioRender.com. (accessed on 11 February 2023)”.
Figure 2. Proposed relationship between the gut microbiome and the structure of blood-brain barrier in health and advanced liver disease. Almost one to two years after birth gut microbiome community is created and its diversity is altered under the influence of age, diet, lifestyle, and geography during life. In healthy, normal gut microbiota-derived metabolites enter circulation and reach the BBB and maintain the integrity of barrier. Advanced liver disease and cirrhosis altered gut microbiota composition. Altered microbiota produces different metabolites that disrupt the intestinal barrier and result in bacterial translocation. Metabolites and products of altered bacteria (i.e., DNA, LPS, etc.) trigger the innate immune system that initiates the production of pro-inflammatory cytokines. Systemic inflammation activates sinusoidal kupffer cells and causes hepatocyte injury that along with gut-derived products and metabolites impair the integrity of the BBB. Leukocyte infiltration, glial activation, oxidative stress, neuroinflammation, and neurodegeneration are the main consequences of BBB damage. BBB: Blood-brain barrier. Created with “BioRender.com. (accessed on 11 February 2023)”.
Biomedicines 11 01272 g002
Table 1. Alteration of gut microbial communities in HE.
Table 1. Alteration of gut microbial communities in HE.
StudyGroups and Gender (Women/Men)Specimen/SamplesMicrobiota AlternationsLocationReferences
DecreasedIncreased
Clinical studies
Bajaj et al., 2014Cirrhosis with HE vs healthy control (193/64)Multi-tagged pyrosequencing on fecal samplesClostridiales XIV; Ruminococcaceae;
Lachnospiraceae
Enterococcaeae
Staphylococcaceae; Enterobacteriaceae
United States[21]
Bajaj et al., 2015Cirrhosis with HE and without HE vs healthy control (18/84)Fecal specimen analysis using multi-tagged pyrosequencing techniquesLachnospiraceae; Ruminococcaceae; Clostridiales XIVEnterobacteriaceae; EnterococcaceaeUnited States[22]
Bajaj et al., 2012Cirrhosis with HE/cirrhosis without HE (10/50)Sigmoid mucosal specimen using 16S ribosomal RNA (rRNA) sequencingRoseburiaEnterococcus; Veillonella; Megasphaera; BurkholderiaUnited States[23]
Zhang et al., 2013Cirrhosis with MHE/cirrhosis without MHE (40/37)Fecal specimen analysis using 16S rRNA-based pyrosequencing -Streptococcus salivarius (as a gut urease-containing bacteria)China[24]
Sung et al., 2019Acute episode of OHE/compensated cirrhosis (36/129)Profiled fecal microbiome alternations from cohortBacteroidetes phylum Firmicute; Proteobacteria; ActinobacteriaTaiwan[25]
Wang et al., 2019Cirrhosis with MHE/ healthy controls (0/91)16S rRNA sequencing on stool-Pasteurellaceae Haemophilus; Alcaligenaceae
Parasutterella
China[26]
Bajaj et al., 2012Cirrhosis with HE/cirrhosis without HE (4/29)Fecal specimen analysis using 16S rRNA sequencing-VeillonellaceaeUnited States[27]
Bajaj et al., 2012Cirrhosis with HE/healthy controls (4/29)Fecal specimen analysis using 16S rRNA sequencing)Clostridiales_Incertae Sedis XIV; Ruminococcaceae; Lachnospiraceae Leuconostocaceae; EnterobacteriaceaeUnited States[27]
Chen et al., 2012Acute-on-chronic liver failure with HE/healthy controls (42/161)Fecal microbiota analysis (16S rRNA sequencing)Lachnospiraceae-China[28]
Yukawa-Muto et al., 2022Cirrhosis with HE/cirrhosis without HE and healthy controls (34/45)Fecal specimen analysis using16S rRNA and metagenomic sequencing-Streptococcus salivariusJapan[29]
Hua et al., 2022Cirrhosis with HE/cirrhosis without HE and healthy controls (13/37)16S rRNA analysis on fecal samplesLachnospiraceae; Turicibacterales; Turicibacter; TuricibacteraceaePasteurellales; Pasteurellaceae; Haemophilus; SelenomonasChina[30]
Lin et al., 2022Cirrhosis with MHE/ healthy controls (-)16S rRNA high-throughput sequencing on fecal specimenLachnospiraceae; Roseburia; CoprpcpccusVeillonellaChina[31]
Bajaj et al., 2021Cirrhosis with HE/ healthy controls (0/150)Stool metagenomics sequencingFaecalibacterium phage; Myoviridae-United States[32]
In Vivo studies
Kang, D. J. et al., 2016Carbon tetrachloride (CCL4)-induced HE/control C57BL/6 miceFecal samples from large intestine and cecumLachnospiraceae; Ruminococcaceae; Clostridiales XIV; Bifidobacteriaceae; Staphylococcaceae; Enterobacteriaceae; Lactobacillaceae-[33]
Yang et al., 2022Bile duct ligation (BDL)-induced HE/control C57BL/6 miceFecal samples collected from sterile cage bottomBacteroidetes; Bacteroidia; MB-A2-108; Erysipelotrichia; Bacteroidales; Erysipelotrichales; Muribaculaceae; Tannerellaceae; Erysipelotrichaceae; Parabacteroides; GCA-900066225 (a genus of the Lachnospiraceae family) Firmicutes; Bacilli; Clostridiales; Clostridia; Lactobacillales; Lachnospiraceae; Alistipes; Lactobacillus murinus; Lactobacillaceae; Lachnospiraceae; Rikenellaceae; Lactobacillus-[34]
Đurašević et al., 2021Thioacetamide (TAA)-induced liver injury/control ratsFecal sampleMuribaculaceae; Desulfovibrionaceae; Lachnospiraceae Christensenella; Rikenellaceae; Bacteroidaceae; Lactobacillaceae-[35]
Yang et al., 2022CCL4-induced liver fibrosis/control C57BL/6 miceFecal sampleStaphylococcusBacteroides; Acinetobacter-[36]
Wu et al., 2022CCL4-induced liver fibrosis/control C57BL/6 micePyrosequencing analysis on fecal samplesBifidobacterium; Turicibacter; (In addition to these 2 genera, 22 bacterial genera had lower abundance)Lactobacillus; In addition to this genus, 6 bacterial genera had higher abundance-[37]
Cabrera-Rubio et al., 2019BDL-induced liver injury/control C57BL/6 micePyrosequencing on fecal sampleFaecalibacterium prausnitzii; Akkermansia; Prevotella; Bacteroides; unclassified Ruminococcaceae-[38]
De Minicis et al., 2014BDL-induced liver injury/control C57BL/6 miceFecal samples from cecumErysipelotrichaceae Lachnospiraceae; Ruminococcaceae-[39]
Li et al., 2019D-galactosamine (GalN)-induced ALF/control Sprague–Dawley ratsFecal sampleChristensenellaceae; Fastidiosipila; Romboutsia Betaproteobacteria; Burkholderiales -[40]
Abbreviations: ALF: Acute liver failure; BDL: Bile duct ligation; CCL4: Carbone tetrachloride; GalN: Galactosamine; HE: Hepatic encephalopathy; MHE: Minimal hepatic encephalopathy; OHE: Overt hepatic encephalopathy; rRNA: Ribosomal RNA; TAA: Thioacetamide.
Table 2. Evidence for disrupting the BBB in liver disease and HE.
Table 2. Evidence for disrupting the BBB in liver disease and HE.
StudyCase/ModelMethodFindingsReference
Human studies
Kato et al., 1992Postmortem on 9 patients with ALFElectron microscopic study on cerebral cortex capillariesSwollen and vacuolated endothelial cells, intact tight junctions, enlargement and vacuolated basement membrane, vacuolated pericytes and swollen perivascular astrocyte end-feet[90]
Sa et al., 2010Postmortem analysis of brain samples from 14 patients with ALFElectron microscopyshrunken and vacuolated endothelial cells disrupted tight junctions and mitochondria, as well as swollen perivascular astrocyte end-feet[91]
Animal studies
Livingstone et al., 1977devascularization and total hepatectomy-induced HE/control Wistar ratsTrypan blue, 14C-inulin, 14C-sucrose, 14C-glucose, 14C-phenylalanine, electron microscopyIncreased brain uptake index of 14C-substrates; raised in the cerebral concentration of trypan blue; Swollen and vacuolated perivascular astrocyte end-feet and their mitochondria[92]
Zaki et al., 1984Devascularized and GalN-induced ALF/control albino rats14C-inulin, 14C-sucrose, 14C-glucoseIncreased brain uptake index of 14C substrates[93]
Mossakowski et al., 1985TAA-induced HE ratsElectron microscopy on the cerebral cortexDegenerative mitochondria and organelles in astrocytes,[94]
Traber et al., 1987GalN-induced ALF/control New Zealand White rabbitsHorseradish peroxidase injection and electron microscopyswollen and vacuolated astrocytic foot processes, intact endothelial cells[95]
Nguyen et al., 2006AOM-induced ALF/control C57BL/6 miceEvans blue dye extravasation, electron microscopyIncreased level of Evans blue dye in cerebral tissues; swollen perivascular astrocyte end-feet[96]
Chen et al., 2009AOM-induced ALF/control miceWestern blot analysis of tight junction proteinsDecreased cerebral proteins of occludin, claudin-5 and ZO-1[97]
Kristiansen et al., 2010Hepatic devascularization and portocaval anastomosis-induced ALF/control Norwegian Landrace pigsElectron microscopy examination on frontal lobe, cerebellum, and brain stemPerivascular edema, abnormal processes of astrocytes and pericytes, swollen neuron[98]
Sa et al., 2010GalN+LPS induced ALF/control BALB/c miceElectron microscopy, immunohistochemistry, Evans blue dye extravasationShrunken and vacuolated endothelial cells, disrupted tight junctions, swollen perivascular astrocyte end-feet; lower protein levels of occludin; Increased cerebral level of Evans blue[91]
Wang et al., 2011Acetaminophen-induced ALF/control BALB/c miceEvans blue dye extravasation, electron microscopy, western blot analysisIncreased level of Evans blue dye in brain tissues; shrunken and vacuolated endothelial cells, incomplete tight junctions, swollen perivascular astrocyte end-feet; decreased the protein expression of occludin in cerebral tissues[99]
Quinn et al., 2014BDL-induced liver injury/control Sprague Dawley ratsImmunofluorescence staining of brain microvasculature, Evans blue dye extravasationlosing the microvessel integrity, increased level of Evans blue dye in brain tissues;[100]
Chastre et al., 2014Azoxymethane (AOM)+LPS-induced ALF and coma/control C57BL/6IgG extravasationIncreased protein expression of IgG in brain tissues[101]
Faleiros et al., 2015TAA-induced ALF/control C57BL/6 miceElectron microscopyAbnormal structure of brain capillary activated endothelial cells and disrupted tight junctions, enlargement of perivascular astrocyte end-feet[102]
McMillin et al., 2015AOM-induced ALF/control C57BL/6 miceEvans blue dye extravasationIncreased cerebral level of Evans blue[103]
Thabut et al., 2015BDL+NH3-induced cirrhosis/control ratsFluorochrome extravasationIncreased cerebral fluorescence intensity[104]
Grant et al., 2018TAA and AOM-induced HE/control C57BL/6 miceEvans blue dye extravasationIncreased level of Evans blue dye in brain tissues[105]
Abbreviations: ALF: Acute liver failure; AOM: Azoxymethane; BDL: Bile duct ligation; GalN: Galactosamine; IgG: Immunoglobulin G; LPS: Lipopolysaccharide; NH3: Ammonia; TAA: Thioacetamide; ZO-1: Zonula occludens-1.
Table 3. Experimental evidence of therapeutic options that target the BBB by modulating gut microbiota.
Table 3. Experimental evidence of therapeutic options that target the BBB by modulating gut microbiota.
Disease (Animal Model)InterventionsFindingsReference
Probiotic and prebiotic
Chronic Parkinson’s disease mouse modelProbiotic (Lacticaseibacillus rhamnosus GG) + prebiotic (polymannuronic acid)Improved integrity of the BBB, prevented dopaminergic neuronal loss and increased glial cell-derived neurotrophic factor and BDNF in striatum, and inhibited striatal apoptosis[184]
APP/PS1 mouse model of Alzheimer’s diseaseProbiotic supplement (several beneficial species)Decreased fluorescence intensity of Evans blue in hippocampus, increased expression of tight junction proteins ZO-1 and occludin in brain, and reduced concentration of LPS and pro-inflammatory cytokines in brain[185]
Mouse model of depressive-like behavior (repeated restraint stress and lipopolysaccharide)New probiotic agent (Komagataella pastoris KM71H)Decreased extravasation of Evans blue dye in prefrontal cortex, and prevented neuroinflammation and cerebral oxidative stress[186]
LPS-induced model of systemic inflammation in ratProbiotic (Lactobacillus plantarum IS 10506)Upregulated glial fibrillary acidic protein (GFAP) and platelet endothelial cell adhesion molecule-1 (PECAM1)in brain[187]
Stress (chronic unpredictable mild stress and sleep deprivation in mice)Probiotic (Lactobacillus plantarum MTCC 9510)Decreased Evans blue concentration in brain, increased hippocampal BDNF, prevented neuroinflammation and oxidative stress[188]
Lead toxicity in miceProbiotic (Lactobacillus fermentum HNU312)Increased the integrity of the BBB, decreased neuroinflammation, and improved anxiety-like and depression-like behaviors[189]
Senescence-accelerated mouse prone 8 mouse model of agingProbiotics (species of Lactobacillus and Bifidobacterium)Improved disruption of the BBB, decreased astrocyte reactivity and microglial activation, reduced plasma and cerebral LPS concentrations, decreased mRNA expression of toll-like receptor 4 and nuclear factor-κB, and reduced neuroinflammation in the brain[190]
Acute mice models of Gulf War illnessPrebiotic (andrographolide)Restored claudin-5 protein level, increased BDNF, and decreased microglial activation in brain[191]
Antibiotic therapy
Transgenic MitoPark mouse model of Parkinson’s diseaseRifaximinDecreased circulatory levels of claudin-5 and occludin (protected the BBB), suppressed systemic inflammation, reduced astrocyte reactivity, and decreased microglial activation[192]
Mouse model of subarachnoid hemorrhageClarithromycinReduced extravasation of immunoglobulin G, and increased ZO-1 protein expression,[193]
Mouse model of postoperative cognitive dysfunctionCefazolinIncreased expression of ZO-1 and occludin, and decreased extravasation of Evans blue in brain tissue[194]
Fecal microbiota transplantation
Chronic rotenone-induced Parkinson’s disease mouse modelFMT (from control mice)Restored tight junction proteins occludin, claudin-5, and ZO-1, decreased astrocyte reactivity, reduced microglial activation, decreased concentration of lipopolysaccharide, and suppressed neuroinflammation (TLR4/MyD88/NF-κB signaling pathway) in substantia nigra[195]
Experimental autoimmune encephalomyelitis mouse model of multiple sclerosisFMT (from control mice)Decreased extravasation of Evans blue, increased protein expression of occludin-5 in the spinal cord, and reduced astrocyte reactivity, as well as decreased microglial activation in brain tissue[196]
Mouse model of spinal cord injuryFMT (from control mice)Reduced the level of Evans blue, restored expression of ZO-1 and occludin, decreased astrocyte reactivity, and reduced microglial activation in the spinal cord[197]
Antibiotics-induced microbiome depletion and BBB permeabilizationFMT (pathogen-free mice)Increased expression of ZO-1 and ZO-2 proteins in cerebral microvessels[198]
Abbreviations: BDNF: Brain-derived neurotrophic factor; BBB: Blood-brain barrier; FMT: Fecal microbiota transplantation; GFAP: Glial fibrillary acidic protein; LPS: Lipopolysaccharide; NF-κB: Nuclear factor kappa-light-chain-enhancer of activated B cells; PECAM1: Endothelial cell adhesion molecule-1; TLR4: Toll-like receptor 4; ZO-1: Zonula occludens-1.
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

Shahbazi, A.; Sepehrinezhad, A.; Vahdani, E.; Jamali, R.; Ghasempour, M.; Massoudian, S.; Sahab Negah, S.; Larsen, F.S. Gut Dysbiosis and Blood-Brain Barrier Alteration in Hepatic Encephalopathy: From Gut to Brain. Biomedicines 2023, 11, 1272. https://doi.org/10.3390/biomedicines11051272

AMA Style

Shahbazi A, Sepehrinezhad A, Vahdani E, Jamali R, Ghasempour M, Massoudian S, Sahab Negah S, Larsen FS. Gut Dysbiosis and Blood-Brain Barrier Alteration in Hepatic Encephalopathy: From Gut to Brain. Biomedicines. 2023; 11(5):1272. https://doi.org/10.3390/biomedicines11051272

Chicago/Turabian Style

Shahbazi, Ali, Ali Sepehrinezhad, Edris Vahdani, Raika Jamali, Monireh Ghasempour, Shirin Massoudian, Sajad Sahab Negah, and Fin Stolze Larsen. 2023. "Gut Dysbiosis and Blood-Brain Barrier Alteration in Hepatic Encephalopathy: From Gut to Brain" Biomedicines 11, no. 5: 1272. https://doi.org/10.3390/biomedicines11051272

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