Next Article in Journal
Intercellular Molecular Crosstalk Networks within Invasive and Immunosuppressive Tumor Microenvironment Subtypes Associated with Clinical Outcomes in Four Cancer Types
Previous Article in Journal
Brain Endothelial Cells Activate Neuroinflammatory Pathways in Response to Early Cerebral Small Vessel Disease (CSVD) Patients’ Plasma
Previous Article in Special Issue
Mixed-Culture Propagation of Uterine-Tissue-Resident Macrophages and Their Expression Properties of Steroidogenic Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hepatic Macrophages as Targets for the MSC-Based Cell Therapy in Non-Alcoholic Steatohepatitis

by
Irina V. Kholodenko
* and
Konstantin N. Yarygin
Laboratory of Cell Biology, Orekhovich Institute of Biomedical Chemistry, 119121 Moscow, Russia
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(11), 3056; https://doi.org/10.3390/biomedicines11113056
Submission received: 15 October 2023 / Revised: 2 November 2023 / Accepted: 3 November 2023 / Published: 14 November 2023
(This article belongs to the Special Issue Macrophages in Health and Non-infectious Disease 4.0)

Abstract

:
Non-alcoholic steatohepatitis (NASH) is a serious public health issue associated with the obesity pandemic. Obesity is the main risk factor for the non-alcoholic fatty liver disease (NAFLD), which progresses to NASH and then to end-stage liver disease. Currently, there are no specific pharmacotherapies of NAFLD/NASH approved by the FDA or other national regulatory bodies and the treatment includes lifestyle adjustment and medicines for improving lipid metabolism, enhancing sensitivity to insulin, balancing oxidation, and counteracting fibrosis. Accordingly, further basic research and development of new therapeutic approaches are greatly needed. Mesenchymal stem cells (MSCs) and MSC-derived extracellular vesicles prevent induced hepatocyte death in vitro and attenuate NASH symptoms in animal models of the disease. They interact with hepatocytes directly, but also target other liver cells, including Kupffer cells and macrophages recruited from the blood flow. This review provides an update on the pathogenesis of NAFLD/NASH and the key role of macrophages in the development of the disease. We examine in detail the mechanisms of the cross-talk between the MSCs and the macrophages, which are likely to be among the key targets of MSCs and their derivatives in the course of NAFLD/NASH cell therapy.

1. Introduction

Obesity is a serious public health problem. It is associated with multiple comorbidities, including type 2 diabetes, cardiovascular diseases, and cancers [1]. Obesity is also one of the main and most common risk factors for the non-alcoholic fatty liver disease (NAFLD). Progressive NAFLD, in 60% of cases, leads to non-alcoholic steatohepatitis (NASH); 41% of patients with NASH develop liver fibrosis, which progresses further to cirrhosis in 22% of patients, of which 2% end up with hepatocellular carcinoma within 3 years [2]. In NASH, hepatocytes undergo steatosis, while ongoing inflammation results in tissue fibrosis and end-stage liver disease [3,4]. Although the rate and course of the NAFLD/NASH progression varies among individuals, it typically proceeds through four stages: (1) lipid accumulation in the liver (NAFLD), (2) early NASH, which, in addition to fat accumulation, is also characterized by liver tissue inflammation, (3) fibrosis, manifested by chronic liver inflammation, tissue damage, and excessive accumulation of extracellular matrix (ECM) proteins, and (4) cirrhosis, the most dangerous stage of NASH, which evolves into generally fatal (without organ transplantation) end-stage liver disease [5].
More than 20 years ago, Day and Saksena [6] proposed the ‘two-hit’ theory of NASH pathogenesis. According to this theory, the first hit is steatosis, which sensitizes the liver to the second hit, rendered by some of a plethora of factors including oxidative stress, endotoxins, lipotoxicity, and necroinflammation. The accumulation of lipids such as diacylglycerol increases insulin resistance, mitochondrial and endoplasmic reticulum stress, and autophagic defects, collectively known as lipotoxicity. This event triggers immune responses in Kupffer cells and hepatic stellate cells, promotes tissue infiltration with neutrophils and lymphocytes, and ultimately leads to fibrosis and death of hepatocytes [7]. Somewhat later, Tilg and Moschen proposed the “multiple parallel hits” hypothesis for the pathogenesis of NASH [8]. They stated that NASH is a pathogenetically complex condition that involves cross-talk between several metabolically active tissues (adipose tissue and the liver/gut axis). According to the “multiple hits” theory, multiple disorders, including insulin resistance, hepatic lipid accumulation, oxidative stress, mitochondrial dysfunction, abnormal gut microbiota characteristics, nutritional factors, and genetic and epigenetic factors, may occur in parallel [8].
Although several drugs are at the late stage of clinical development, NASH therapy is still based on lifestyle adjustment (diet, exercise, etc.) and the prescription of medicines improving lipid metabolism, enhancing sensitivity to insulin, balancing oxidation, and counteracting fibrosis. There are currently no FDA-approved pharmacotherapies for the treatment of NAFLD/NASH [9]. Therefore, new therapeutic options for NASH are greatly needed. Promising results have been obtained in a variety of preclinical studies conducted in various animal models of NAFLD/NASH using mesenchymal stem/stromal cells (MSCs) and their derivatives, including conditioned media, extracellular vesicles, and apoptotic bodies [10,11,12]. The beneficial effects of the cell therapy of many diseases with MSCs are at least partly associated with immunomodulation via direct and indirect interactions with different immune cells [13], including macrophages [14]. The reactions of Kupffer cells, the liver macrophages, to MSC transplantation have also been reported [15].
In this review, we provide an update on the pathogenesis of NAFLD/NASH and the key role of Kupffer cells and recruited macrophages in the development of inflammation and disease progression. We examine in detail the mechanisms of the cross-talk between MSCs and macrophages, which likely mediate the therapeutic effects of MSCs and their derivatives in the treatment of NAFLD/NASH.

2. NAFLD/NASH Pathogenesis

Inflammatory liver diseases, in the absence of contagious pathogens, are denoted as “sterile”. These conditions include xenobiotic intoxication, e.g., as a result of para-acetaminophenol (APAP) administration, cholestatic liver injury, liver ischemia–reperfusion injury (IRI), non-alcoholic steatohepatitis (NASH), and alcoholic liver disease (ALD). Triggers of pathological events in sterile liver diseases include various xenobiotics or endogenous bioactive molecules that initiate the release of intracellular compounds from living or dying parenchymal cells. It has been shown that both acute and chronic liver diseases involve the massive death of hepatocytes [16]. Extensive hepatocyte death can be and commonly is diagnosed by observing an increase in the serum levels of alanine aminotransferase (ALT) and aspartate aminotransferase (AST). These parameters have significant prognostic value for patients with NASH [17,18] and other liver diseases since they are highly correlated with overall and liver-specific mortality in the general population [19,20], highlighting the role of cell death as a major driver of liver disease progression and the development of liver fibrosis, cirrhosis, and hepatocellular carcinoma.
In liver pathologies cells die in different ways. For example, a wide range of cell death types, including apoptotic [21], necrotic [22], autophagic [23], pyroptotic [24], and ferroptotic [25] cell death were found in alcoholic liver disease (ALD). Autophagy and apoptosis are considered to be the predominant modes of hepatocyte death in ALD, while the contribution of necroptotic cell death is still a controversial issue [26]. In APAP hepatotoxicity, mitochondrial permeability transition-mediated regulated necrosis is the main type of cell death, playing the key role in liver damage [27], while the contribution of autophagy and apoptosis seems unlikely [26].
NASH is characterized by steatosis, ballooned hepatocytes containing hyaline inclusions (Mallory bodies), and severe inflammation leading to fibrosis, cirrhosis, and hepatocellular carcinoma. There is strong evidence that hepatocyte cell death is a driver of inflammation and fibrosis in NASH [28]. However, the predominant mode of cell death in this metabolic disease is still debated. Liver biopsies of patients with NASH showed increased TUNEL assay positivity and caspase-3/7 expression, indicating the prevalence of apoptotic cell death of hepatocytes in NASH [29]. Also, in a high-fat diet (HFD) mouse model of NASH, it was shown that use of the pan-caspase inhibitor Emricasan significantly attenuated liver injury, inflammation, fibrosis, and apoptosis, suggesting a critical role for apoptosis in the pathogenesis of NASH [30]. However, despite the clearly beneficial effects of the pan-caspase inhibitor in animal models and early clinical trials in a small group of NASH patients [31,32], randomized clinical trials of the pan-caspase inhibitor Emricasan conducted in a larger group of NASH patients did not show significant clinical improvement [33]. In the short term, treatment with Emricasan reduced serum ALT levels, but it not only did not improve the liver histology in patients with NASH, but even worsened fibrosis and hepatocyte ballooning. The authors of the study suggested that treatment with the pan-caspase inhibitor could provoke the onset of the alternative mechanisms of cell death inciting more pronounced fibrosis and hepatocyte ballooning [33]. Nevertheless, many studies suggest that in NASH apoptosis appears to be the predominant mode of cell death and can occur through both intrinsic (via lipotoxicity and organelle stress) and extrinsic (via cell surface receptors) pathways. Many authors do not exclude the contribution of pyroptosis and ferroptosis in the development of NAFLD/NASH. While the participation of pyroptosis in the pathogenesis of NAFLD/NASH was mainly demonstrated in animal models [34,35] and less reliably in patients with NASH [36], the hallmarks of ferroptosis, such as lipid peroxidation, ROS accumulation, and increased liver iron stores (both in the parenchymal and non-parenchymal compartment) were identified in most NASH patients [37]. In addition to apoptotic cell death of hepatocytes in NASH livers, there is also evidence of necrotic cell death. Free cholesterol accumulation in NASH patients [38,39] was demonstrated to cause apoptotic and necrotic hepatocyte death through activation of the c-Jun N-terminal kinase 1 (JNK1) [40].
Dying hepatocytes release a large number of different molecules capable of causing infection-independent activation of immune responses and their repertoire depends on the type of cell death [41]. According to Matzinger’s “danger theory” of innate immunity [41,42,43], these cell death type-specific libraries of endogenous molecules, released in all tissues in response to the non-contagious harmful stimuli and named damage-associated molecular patterns (DAMPs), are scanned by the immune cells and recognized as “self-safety” or “self-danger” signals. Because DAMPs induce inflammatory responses in the absence of pathogenic infection, the inflammation they induce is called sterile inflammation [44]. DAMPs can activate not only innate immune cells such as macrophages, neutrophils, and dendritic cells, but also non-immune cells including fibroblasts [45], hepatic stellate cells [46], and endothelial cells [47]. Activation of both immune and non-immune cells by DAMPs results in the production of various cytokines and chemokines that trigger adaptive immune responses. Although sterile inflammation plays a critical role in tissue repair, regeneration, and homeostasis, unresolved chronic inflammation leads to the development of sterile inflammatory diseases, including the liver diseases mentioned above. DAMPs comprise nuclear proteins (e.g., high mobility group box-1, HMGB-1), cytosolic proteins (e.g., keratin-18), and mitochondrial components or mitochondrial DAMPs (mtDNA), including N-formyl peptides, cardiolipin, adenosine triphosphate (ATP), mitochondrial transcription factor A (TFAM), and nucleic acids in various conformations (e.g., single-/double-stranded RNA or DNA) [48]. DAMPs can initiate immune responses through the activation of classical pattern recognition receptors (PRRs), which include membrane-bound Toll-like receptors (TLRs) and C-type lectin receptors (CLRs), cytoplasmic NOD-like receptors (NLRs), retinoic acid inducible gene I (RIG-I)-like receptors (RLRs), and multiple intracellular DNA sensors, as well as via non-PRR transmembrane proteins, including receptors for advanced glycation end products (RAGE), triggering receptors expressed on myeloid cells (TREMs), G-protein-coupled receptors (GPCRs), and ion channels [49].
NASH is a chronic state of sterile inflammation initiated by the presence of DAMPs released from damaged hepatocytes. DAMPs triggering NASH-associated inflammation have been reliably identified and involve extracellular ATP, keratin 18, HMGB1, various mtDAMPs, etc. Free fatty acid-induced lipoapoptosis in human hepatocytes stimulates ATP release, promoting monocyte recruitment [50,51]. In addition, free ATP attracts circulating neutrophils, ensuring their adhesion within the hepatic sinusoids, while mitochondrial DAMP N-formyl peptides direct neutrophils to the site of injury [52]. Lipotoxic damage also includes mitochondrial permeability transition with cytochrome c release, mitochondrial impairment, and oxidative stress. Liver biopsies from patients with NAFLD or NASH have been found to exhibit both structural and functional mitochondrial abnormalities [53]. Structural abnormalities include organelle enlargement, while functional abnormalities result in increased ROS production or accumulation of lipid peroxides [53]. Furthermore, plasma from NASH patients has recently been shown to contain intact mitochondria and high levels of oxidized mitochondrial DNA enclosed in hepatocyte-derived microparticles [54,55]. Keratin 18 is the major epithelial-specific liver intermediate filament protein expressed by hepatocytes and cholangiocytes [56,57]. In apoptosis, keratin 18 is one of the most important substrates of activated caspases generating a neo-epitope after cleavage. Caspase-cleaved fragments of keratin 18 are released into the blood as part of inclusion bodies [58,59]. During necrotic cell death, intact keratin 18 is released into the blood, where it circulates as part of extracellular vesicles [60]. Thus, the levels of fragmented keratin 18 and intact keratin 18 reflect apoptotic and total cell death, respectively, while the ratio serves as an apoptotic index [60]. Patients with NASH show an increased number of caspase-cleaved keratin 18 fragments compared to patients with simple steatosis [29]. However, the clinical assessment of the diagnostic value of the measurement of the keratin 18 fragment levels is still controversial (reviewed in [60]). Keratin 18 has so far attracted attention due to its usefulness in the establishment of cell death modes and as a biomarker of liver damage rather than as a DAMP molecule.
The high-mobility group protein B1 (HMGB1), which is released from damaged hepatocytes in chronic liver diseases, is probably the most widely assessed DAMP molecule [61]. Many studies have shown that HMGB1 is involved in the pathogenesis of NAFLD. HMGB1 plays a critical role in the initiation and maintenance of a chronic inflammatory state in liver tissue. Since hepatic inflammation is one of the histological features distinguishing simple steatosis and NASH, HMGB1 likely mediates the transition from steatosis to NASH. In NAFLD, lipotoxicity leads to the release of HMGB1 and it is a driver of sterile inflammation [62]. This observation was substantiated using a large cohort study of pediatric patients with biopsy-confirmed NAFLD, which showed higher levels of circulating HMGB1 in children with NAFLD than in obese-only controls. The level of circulating HMGB1 not only correlated with the degree of fibrosis, but also with the levels of inflammatory mediators such as TGF-β and MCP-1 [63]. These results suggest a critical role for HMGB1 in the progression of NAFLD [63]. Therefore, serum HMGB1 may serve as a potential biomarker for the early diagnosis of NAFLD and as a target for therapies aimed at the prevention and treatment of NAFLD-associated inflammation. HMGB1 protein is both a nuclear factor and a secreted protein. In the cell nucleus, it acts as an architectural chromatin-binding factor that bends DNA and promotes protein assembly at specific DNA targets. HMGB1 is secreted by activated monocytes and macrophages and passively released by necrotic or damaged cells [64]. HMGB1 interacts with several receptors such as TLR2, TLR4, and RAGE. TLR4 is the primary receptor for HMGB1 in mediating macrophage activation and cytokine release during tissue injury. Activation of the HMGB1 binding receptors results in a wide range of inflammatory responses, including the production of pro-inflammatory cytokines and the recruitment of immune cells to the site of injury. HMGB1 may also act in conjunction with other pro-inflammatory mediators, such as single-stranded DNA, LPS, IL-1β, and nucleosomes to induce inflammation. Structurally, among the two DNA-binding domains of HMGB1, the B-box recapitulates the inflammatory activity of the full-length protein, whereas the A-box counteracts it [65]. In multiple experimental models of murine liver fibrosis associated with cholestasis, alcoholic steatohepatitis or NASH increased expression and release of HMGB1 are induced [66]. Notably, neutralization of HMGB1 protected against liver fibrosis, whereas injection of recombinant HMGB1 promoted liver fibrosis [66]. Extracellular HMGB1 is capable of activating the RAGE-PI3K-AKT1/2/3 pathway or the RAGE-ERK pathway to upregulate type I collagen synthesis by hepatic stellate cells [66,67]. Importantly, HMGB1 does not stimulate profibrotic signaling through TLR2, TLR4, and TLR9. The receptor specificity in the profibrotic response appears to be distinct from that in liver inflammation, where HMGB1 interacts with RAGE and TLRs. Via the TLR4, HMGB1 secreted from lipotoxic hepatocytes facilitates a paracrine cytolytic effect on neighboring hepatocytes loaded with cholesterol [40,61]. NASH pathogenesis is presented in Figure 1.
Thus, the pathogenesis of NAFLD/NASH starts with the death of hepatocytes and the release of DAMPs from dying cells. DAMPs primarily activate resident hepatic macrophages and, subsequently, together with the cytokines and chemokines released by macrophages, recruit various cells of innate and adaptive immunity to the liver, ultimately leading to massive inflammation.

3. Role of Kupffer Cells in the Pathogenesis of NASH

Despite the abundance of data, the pathophysiology of NASH remains incompletely understood at both molecular and cellular level. It is clear, though, that the innate and adaptive immune systems play critical roles in the pathogenesis of NASH. Key players are the pro-inflammatory cytokines such as TNFα, IL-1β, and IL-6, as well as adipokines derived from liver and adipose tissue [68]. These cytokines are the main mediators of hepatic acute phase reaction triggering host intra-liver defenses, but if the problem turns chronic, they become implicated in the development of fibrosis and cirrhosis and participate not only in the deterioration of tissue homeostasis, but also affect liver-specific metabolic functions by altering the gene expression of key regulatory enzymes of the carbohydrate and lipid metabolism [69]. The pro-inflammatory environment in the liver favors the deterioration of very low-density lipoprotein (VLDL) secretion and transcriptional upregulation of de novo lipogenesis, as well as the weakening of mitochondrial fatty acid oxidation, generally increasing lipid accumulation in the liver [70]. It may be surprising that NASH develops in immunodeficient mice that lack B- and T-cells, as well as NK cells. It has become clear that there is a connection between the immune system and the metabolic disorders associated with obesity. The innate immune system senses nutritional stress and mediates the progression of obesity, NASH, and type 2 diabetes. Winkler et al. [71] used an immunodeficient mouse model that allowed the separation of adaptive and innate immune responses and showed that NASH develops even in the absence of adaptive immunity. Since the adaptive immune system cells did not play a significant role in this model, it is likely that the resident liver macrophages and dendritic cells of the innate immune system were the main mediators of inflammatory hepatic steatosis at the cellular level. It is the pro-inflammatory cytokine TNFα, secreted during chronic nutritional overload by the resident liver macrophages, Kupffer cells, that is considered the key player in NASH pathogenesis [72]. Downstream targets of Kupffer cells are hepatocytes and hepatic stellate cells, which are activated and transformed into myofibroblasts in response to TNFα.
Kupffer cells playing the central role in liver tissue inflammation are specialized macrophages lining the walls of the hepatic sinusoids from the lumen side and accounting for about 30% of sinusoidal cells [73]. Kupffer cells carry out five major functions critical for maintaining the liver and whole-body homeostasis. These include (1) clearance of cellular debris and metabolic waste [74,75,76], (2) maintenance of iron homeostasis through phagocytosis of red blood cells and subsequent recycling of iron [77,78,79,80], (3) regulation of cholesterol homeostasis through cholesteryl ester production transfer protein, which is important for reducing circulating levels of high-density lipoprotein-cholesterol and increasing levels of very low-density lipoprotein-cholesterol [81], (4) mediating antimicrobial defense [82,83], and (5) promoting immunological tolerance [84,85].
The origin of Kupffer cells in embryogenesis was described in detail by Li et al. [86]. The Kupffer cell pool formation follows a program that is consistent between mice and humans and occurs in three waves at different stages of embryonic development. Embryonic Kupffer cells reside in the liver throughout life and their population is preserved due to proliferation [87,88]. In adults, the pool of Kupffer cells can be replenished by bone marrow monocytes migrating to the liver and differentiating into Kupffer cells [89].
Kupffer cell populations in mice and humans are heterogeneous. Regrettably, human Kupffer cells do not have uniform markers for identification. In three independent scRNA-seq studies, human Kupffer cells were identified as CD163+MARCO+CD5L+TIMD4+ [90,91,92]. Functionally, these cells exhibit anti-inflammatory, anti-tumor, and immunomodulation activity [90,91,92]. Compared to MARCO- macrophages, Kupffer cells express fewer inflammatory (TNFα), but more immunosuppressive genes (e.g., PDL1) when stimulated with LPS or IFN-γ [90]. Another study showed that Kupffer cells display the CD32intCD68+CD14+ phenotype and have the potential to regulate the immune response [93]. In a recent study, VSIG4 was identified as the best marker of human Kupffer cells [94]. Blériot et al. [95] discovered two subpopulations of embryonic Kupffer cells in the mouse liver that differ phenotypically and functionally: a major CD206loESAM- population (KC1) and a minor CD206hiESAM+ population (KC2). KC2 were more likely to express genes involved in metabolic processes in both steady state and diet-induced obesity and hepatic steatosis [95]. Human embryonic Kupffer cells express CD49a, whereas bone marrow derived Kupffer cells do not express this marker. Actually, CD49a can be used to distinguish those two pools of Kupffer cells. Human embryonic Kupffer cells express high levels of both the pro-inflammatory TNFα and IL-12 and the anti-inflammatory IL-10 cytokines, suggesting a dual role in inflammation. However, LPS does not affect the expression of these cytokines in the embryonic Kupffer cells. In contrast, these three cytokines are expressed, though at low levels, in Kupffer cells derived from the bone marrow monocytes and their expression is upregulated by LPS [96]. Therefore, the embryonic Kupffer cells appear to be functional in homeostasis, whereas bone marrow derived Kupffer cells become functional in cases of injury.
In mice, macrophages (MoMϕs) recruited from the bone marrow to the liver have the CD11b+F4/80intLy6C+CSF1R+ phenotype, while the resident Kupffer cells display the CD11blowF4/80highClec4F+ phenotype [97,98,99,100,101]. MoMϕs differentiate from circulating monocytes, which are derived from bone marrow CX3CR1+CD117+Lin progenitor cells [102]. In mouse models of liver disease, hepatic MoMϕs are divided into two major subpopulations according to the level of Ly6C expression: the inflammatory Ly6Chigh and the restorative Ly6Clow MoMϕs. Depending on the signals expressed by the liver microenvironment, recruited MoMϕs can differentiate into cells of different phenotypes. For example, the recruitment of inflammatory Ly6Chigh MoMϕs specifically relies on the CCL2/CCR2, CCL1/CCR8, and CCL25/CCR9 signaling pathways, with chemoattractants secreted by activated Kupffer cells, hepatic stellate cells, and liver sinusoidal endothelial cells [103,104,105,106,107]. Inhibition or elimination of these signaling pathways in mice results in reduced MoMϕ recruitment, hepatic inflammation, and overall fibrosis [103,108]. However, it should be noted that MoMϕs are highly plastic, as shown by the potential of Ly6Chigh MoMϕs to switch to the restorative Ly6Clow phenotype [109].
Due to their central position in the hepatic microenvironment, their long cytoplasmic protrusions, and the high density of pattern recognition receptors (PRRs) on their surface, including Toll-like receptors (TLRs) and nucleotide binding oligomerization domain-like receptors (NLRs), Kupffer cells act as the first line of responders for liver damage [110]. Kupffer cells play diverse roles in regulating inflammation in NASH. NASH induces the upregulation of 891 Kupffer cell genes associated with ECM remodeling, lipid metabolism, bacterial clearance, and recruitment of circulating monocytes [111]. Kupffer cells are able to enhance or attenuate hepatic inflammation. Mitochondrial DNA from apoptotic hepatocytes activates the STING/NF-kB signaling pathway in Kupffer cells and leads to the amplification of inflammation [112]. As initial sensors of liver injury, Kupffer cells recruit monocytes, neutrophils, and maybe other leukocytes through the secretion of a variety of chemokines, including CCL2 and CXCL1 [113]. Kupffer cells are the main source of CCL2 [114,115], which attracts CCR2+ monocytes to the damaged liver. Kupffer cells also secrete CXCL1, CXCL2, and CXCL8 to attract neutrophils [114], which contribute to hepatic ischemia–reperfusion (IR) injury and heat-induced liver injury [52,113,116]. It appears that infiltrating Ly6Chigh MoMϕs also release chemokines and promote leukocyte recruitment during liver disease. For example, in mouse models of hepatic fibrosis induced by carbon tetrachloride (CCl4) and an MCD diet, Ly6Chigh MoMϕs express CXCL16 and promote the recruitment of CXCR6+ natural killer T (NKT) cells, which exacerbate inflammation and fibrogenesis [117]. In mice fed a high fat diet, Ly6Chigh MoMϕs produce CCL5 and CXCL9 in a S100 calcium-binding protein A9 (S100A9)-dependent manner. These chemokines lead to liver recruitment of both CD4+ and CD8+ T cells, which contribute to insulin resistance [118,119]. Studies of chronically inflamed livers from patients with alcoholic liver disease (ALD), NASH, primary biliary cholangitis, or primary sclerosing cholangitis have also shown that intermediate CD14highCD16+ monocytes (similar to Ly6Chigh MoMϕs in the murine liver), which are derived from infiltrating classical CD14highCD16 monocytes, secrete pro-inflammatory cytokines and chemokines such as TNFα, IL-1β, CCL1, and CCL2 [120]. Figure 2 schematically shows the inflammatory process mediated by Kupffer cells in the NASH liver.
Inflammosomes are multiprotein complexes that can sense danger signals sent by pathogens and damaged cells via TLRs and NLRs. Inflammosome activation induces caspase-1-mediated cleavage and maturation of the IL-1β and IL-18 cytokines [121]. In the liver, gut-derived PAMPs (pathogen-associated molecular patterns), cell damage-induced DAMPs (e.g., ATP), crystals (e.g., cholesterol), palmitic acid, and ROS present the well-characterized signals that induce inflammosome activation in macrophages [35,122,123,124]. DAMPs such as ATP and uric acid, which are released from damaged hepatocytes, induce inflammosome activation in hepatic Kupffer cells in mouse models of ASH and NASH [125,126]. Macrophage inflammosome activation plays an important role in the pathogenesis of chronic liver diseases such as ALD and NASH [127]. IL-1β, which is released as a result of the inflammosome activation in Kupffer cells, plays a critical role in the promotion of alcohol-induced steatosis, inflammation, and liver injury [128]. Deletion of caspase-1 or caspase-1 adapter in mice results in impaired IL-1β production, thereby attenuating ALD symptoms [128]. With regard to NASH, in vitro studies have shown that the NLRP3 inflammosome can sense lipotoxicity-associated increases in intracellular levels of ceramides, cholesterol crystals, saturated fatty acid content, mtDNA levels, and ROS content, causing caspase-1 induction in macrophages and subsequently promoting IL-1β production [129,130,131,132]. In mouse models of NASH induced by supplying various diets, including an atherogenic diet, methionine- and choline-deficient diet (MCD), high fat diet (HFD), and Western diet, activation of NLRP3 inflammosomes and subsequent production of IL-1β exacerbates inflammatory responses while increasing the IL-6 and CCL2 levels and the numbers of infiltrating MoMφ and neutrophils [132]. Genetic deletion or pharmacological inhibition of NLRP3 in mice significantly suppresses tissue inflammation and improves pathological features of NASH such as fibrosis and insulin resistance [129]. Liver samples from patients with NASH showed significantly increased expression of the inflammosomal genes (NLRP3 and caspase-1) compared with normal liver tissue [133] or liver samples from patients with non-alcoholic fatty liver disease [134].
It has been shown that in NASH the pool of embryonic Kupffer cells is reduced with increasing dietary cholesterol [135,136]. A reduction in the number of Kupffer cells is also observed in MCD diet-induced NASH [135] and in hepatocellular carcinoma [137]. This may be caused by apoptosis or cell death induced by the disease [138,139]. During MCD feeding in mice, the pool of the Ly6CloClec4F+Tim4+ resident Kupffer cells was significantly depleted, while that of the recruited Ly6CloClec4FTim4 monocyte-derived macrophages was enriched. During recovery, the number of resident Kupffer cells normalized, as well as the level of recruited monocyte-derived macrophages returning to the baseline. It is noteworthy that Ly6CloClec4F+Tim4- monocyte-derived Kupffer cells did not self-renew during the recovery period, but were reduced, while recovery from MCD diet feeding was characterized by an increased content of Ki-67+ proliferating resident embryonic Kupffer cells [135]. MoMϕs are considered the main contributors to the replenishment of the macrophage pool. This process depends on the concerted actions of hepatic stellate cells and hepatic sinusoidal endothelial cells, which orchestrate monocyte occupancy and the imprinting of the Kupffer cell phenotype, including the expression of the transcription factors ID3 and liver X receptor-alpha (LXR-α) [140,141].
Because hepatic macrophages are the primary component of the immune response during the development of sterile inflammation in the setting of non-alcoholic fatty liver disease, the precise orchestration of their activation, differentiation, and polarization plays a critical role in resolving or triggering subsequent disease progression. As mentioned above, the transition from simple steatosis (mild form of NAFLD) to NASH depends on the development of liver inflammation. Early in the disease, Kupffer cells rapidly divide and secrete cytokines and chemokines such as IL-1, TNF-α, MCP-1, and C-C motif chemokine ligand (CCL)5 [142], reflecting their involvement in controlling the inflammatory response in NASH and their major role in recruiting inflammatory cells to the liver [143,144]. A balance between the states of macrophage polarization is crucial for the development of steatohepatitis. In the reparative phase of acute inflammation, hepatic macrophages can undergo activation into M2 macrophages displaying an immunosuppressive but profibrogenic phenotype. Activated macrophages can produce high levels of TGF-β1 and IL-13, leading to the progression of fibrosis [145]. It has been shown that Kupffer cells can both secrete and respond to the pro-inflammatory cytokines, such as IL-6, and anti-inflammatory cytokines, such as IL-10. The anti-inflammatory M2 macrophages can induce apoptosis of the M1 Kupffer cells, which have been reported to slow down the progression of fatty liver diseases [146]. Additionally, the macrophage-derived IL-10 promoted lipid catabolism can subsequently prevent hepatic inflammation [147]. Macrophages have been implicated in the progression of various stages of NAFLD. For example, a study conducted in a young Korean population demonstrated a high number of CD68 + Kupffer cells in biopsy specimens from patients with advanced NAFLD [148]. In another study, increased numbers of activated macrophages were found interspersed among injured hepatocytes in children with NAFLD [149]. Kupffer cells can be directly stimulated by the excess of free fatty acids and cholesterol. The chemokines they release (CC-chemokine ligand 1 (CCL1), CCL2, and CCL5) promote the differentiation of monocytes into M1 activated macrophages [150]. In phase II clinical trials, inhibition of CCL2 and CCL5 signaling through selective blockade of CCR2 and CCR5 receptors attenuated the signs of NASH, reduced the levels of circulating biomarkers of systemic inflammation, including high-sensitivity C-reactive protein, IL-6, fibrinogen, and IL-1ß, and counteracted the development of fibrosis [151]. Tacke, in his review article [152], described several effective approaches to targeting hepatic macrophages in the treatment of NASH, including specific inhibitors of inflammatory signaling aimed at suppressing Kupffer cell activation (for example, the ASK-1 inhibitor selonsertib); inhibition of monocyte recruitment through pharmacological strategies to interfere with chemokine signaling, including monoclonal antibodies against chemokines or their receptor(s), receptor antagonists that prevent chemokine binding, inhibition of chemokines by aptamer molecules or small molecule inhibitors blocking chemokine-induced intracellular signaling; and use of novel drug carrier materials that may deliver specific drugs (e.g., gene silencing siRNA, inhibitors of inflammatory signaling, and enhancers of autophagy) to hepatic macrophage subsets in order to inhibit their disease-aggravating phenotype.
Thus, both pools of Kupffer cells and recruited macrophages play a critical role in the pathogenesis of liver diseases. The above-mentioned therapeutic approaches targeting hepatic macrophages, modulating their activity and recruitment, although showing some success, are not completely effective in resolving NAFLD/NASH and other liver diseases. Probably, their less than desired effectiveness is explained by the multifactorial and multidirectional nature of the pathogenetic mechanisms of liver diseases. In this regard, the use of mesenchymal stem cells and their derivatives for therapy might turn out to be a promising approach since the mechanisms of their therapeutic effects are also multifactor and may be influencing multiple pathogenetic links.

4. Cell Therapy of NASH Based on the Cross-Talk between Transplanted MSCs and Liver Macrophages

Transplantation of MSCs and MSC derivatives such as exosomes or conditioned medium (CM) is effective in resolving inflammation, oxidative stress, fibrosis, and fatty acid and triglyceride accumulation in various NAFLD/NASH mouse models [153,154,155,156]. This therapeutic action of MSCs and their derivatives can be at least partly attributed to their immunomodulatory effect on hepatic macrophages. Ezquer et al. [157] showed that intravenously administered syngeneic bone marrow derived MSCs prevented the onset of NASH in obese mice. The observed hepatoprotection was not associated with the resolution of metabolic syndrome, but with the prevention of the inflammatory process. Really, the inflammatory cytokines IL-1β, INF-γ, TNF-α, and TGF-β1 mRNA levels were lower in MSC-treated obese mice than in untreated obese mice [157]. Human UC-MSC exosomes intravenously transplanted into mice with MCD-induced NASH improved MCD-induced weight loss and liver damage, and down-regulated pro-inflammatory cytokines TNF-α, IL-6, and IL-1β in the plasma [158]. Compared with control MCD mice, more F4/80+ macrophages were recruited to the liver in the UC-MSC exosome group, indicating that exosomes potentially alter the distribution and chemotaxis of macrophages in the injured liver. In addition, the levels of the markers of the anti-inflammatory macrophage phenotype, namely CD206, Arginase-1, and IL-10, were increased in MCD mice compared with control diet mice, and their expression was further enhanced in the exosome-treated group, indicating that in the MCD-induced liver injury external exosomes might increase the number of the anti-inflammatory macrophages or promote macrophage polarization into an anti-inflammatory phenotype [158]. In a murine NASH model induced by supplying a combination of Western diet and the repeated administration of low doses of carbon tetrachloride, intravenous administration of conditioned medium from stem cells derived from human exfoliated deciduous teeth reduced liver fibrosis and inflammation, inhibited hepatocyte apoptosis and activation of inflammatory macrophages, and decreased pro-inflammatory and profibrotic mediators including TNF-α, TGF-β, and CCL-2 [159]. In mice with NASH induced using a Western diet combined with lipopolysaccharide treatment, human adipose tissue-derived MSCs and their sEVs significantly decreased serum ALT levels and inflammatory markers but did not affect fat accumulation in the liver. Also, after administering human MSCs and sEVs an improvement in fibrosis and an increase in anti-inflammatory macrophages were observed in the mouse liver [160]. This mechanism may be related to the ability of MSCs to polarize macrophages into the anti-inflammatory M2 phenotype. The M1 and M2 subsets of macrophages play highly dynamic roles in the progression of NASH [161]. In liver fibrosis, it has been found that M1 macrophages are anti-fibrotic due to their ability to actively phagocytose debris and to degrade connective tissue, while M2 macrophages are profibrotic due to the secretion of the tissue-remodeling factors, including fibronectin-1, coagulation factor XIII, tissue-type plasmin activator, matrix-associated protein betaIG-H3, and insulin-like growth factor (IGF) [162]. In vitro, immortalized E1-MYC 16.3 human ESC-derived MSC-EVs are able to significantly polarize M0 macrophages, promoting their differentiation from human peripheral blood monocytes into M2 (CD68+CD206hiCD163hi macrophages) but not into M1 (CD68+PDL1hiCD38hi macrophages). In vivo, these MSC-EVs were shown to induce the enrichment of CD163+ M2 macrophages in the livers of high-fat diet (HFD)-induced NASH animals and a decrease in the serum IL-6 levels. However, despite the increase in the number of profibrotic M2 macrophages in the liver and the reduction in plasma IL-6, MSC-EVs did not exacerbate fibrosis, but reduced it, also showing a reduction in the NAFLD activity score (NAS assessed using three parameters: the presence of micro- and macrovesicular fat deposits (steatosis), hepatocellular ballooning, and inflammatory cell infiltration) [163]. However, another study showed that human amnion-derived MSC-EVs reduced the number of CD68+ Kupffer cells and CD11c+ M1 pro-inflammatory macrophages without affecting the number of CD163+ M2 anti-inflammatory macrophages in the livers of rats with HFD-induced NASH. It is likely that amnion-derived MSC-EVs attenuated the inflammatory response in the NASH rat model by suppressing the activation of Kupffer cells, especially M1 macrophages, and down-regulating the expression of the inflammatory cytokines (TNF-α, IL-1β, and IL-6), rather than by shifting the M1/M2 macrophages ratio [164]. In a rat model of sepsis-induced liver injury, the treatment with human umbilical cord-derived MSCs inhibited the activation of Kupffer cells towards M1 phenotype, attenuated TNF-α and IL-6 expression, and promoted IL-4 and IL-10 expression both in vivo in septic rats and in vitro in the LPS-treated Kupffer cells [165].
The MSC-mediated modulation of immune responses at the macrophage level occurs due to paracrine factors secreted by MSCs, such as the anti-inflammatory cytokines IL-10 and TGF-β and a variety of immunosuppressive factors, such as heme oxygenase (HO-1) and iNOS, indoleamine 2,3-dioxygenase (IDO), or prostaglandin E2 (PGE2), and due to the direct cell-to-cell contacts [14]. Wang et al. [166], for example, showed that MSCs can exert a hepatoprotective effect in D-galactosamine-induced liver failure through PGE2-induced regulation of macrophage polarization. PGE2 secreted by mouse bone marrow-derived MSCs inhibited Kupffer cell apoptosis via TLR4-ERK1/2-caspase 3 pathway regulation in mouse non-heart-beating liver transplantation model [167]. Some of the above-named factors secreted by MSCs, including TGF-β and indoleamine 2,3-dioxygenase, induce M2 polarization of macrophages and reduce macrophage-mediated inflammatory responses [168,169]. Mitochondrial transfer from MSCs to macrophages enhances their swallowing ability [170], and also promotes phagocytosis and suppresses the secretion of the pro-inflammatory cytokines [171]. Another mechanism of immunomodulation mediated by MSCs may be related to their ability to suppress mitochondrial fission in Kupffer cells, thereby promoting their M2 polarization. In a hepatic warm ischemia–reperfusion injury (IRI) model, MSCs significantly limited the phenotypic M1 polarization, but enhanced the M2 polarization of Kupffer cells isolated from the ischemic liver, as evidenced by the decreased levels of iNOS and IL-1β transcripts, but increased levels of Mrc-1 and Arg-1 transcripts. MSCs inhibited mitochondrial fission in Kupffer cells, as evidenced by a decrease in the levels of Drp1 and Dnm2. The overexpression of Drp1 in the Kupffer cells promoted mitochondrial fission during IR injury and abolished MSC-regulated Kupffer cell M1/M2 polarization after IR injury, thereby antagonizing the therapeutic effect of MSCs [15]. Table 1 presents data on the effects exerted by MSCs or MSC derivatives on macrophages/Kupffer cells.
In the murine traumatic hemorrhagic shock model, MSC-derived extracellular vesicles containing IL-10 mainly accumulated in the liver, where they were captured by Kupffer cells and induced the expression of the protein tyrosine phosphatase non receptor 22 (PTPN22) [172], which had been shown to negatively regulate the pro-inflammatory macrophage activation [173]. It subsequently shifted Kupffer cells to the anti-inflammatory phenotype and mitigated liver inflammation and injury [172]. It is known that after transplantation MSCs quickly die within a few hours [174]. A number of studies have shown that those dead MSCs are engulfed by macrophages with the non-classical Ly6Clow phenotype found in the lungs and liver. In vitro experiments showed that following the phagocytosis of UC-MSCs classical human CD14++/CD16 monocytes polarized into the non-classical CD14++CD16+CD206+ monocytes. Such non-classical or M2 macrophages began to express the PDL-1 and IL-10, while the expression of the pro-inflammatory TNF-α was reduced [175]. That is, phagocytosed MSCs induce the M2 polarization of macrophages. Direct cell-to-cell contacts between MSCs and macrophages due to the interaction of CD47 present on the MSCs and SIRPα displayed by the macrophages led to a reduction in the liver tissue inflammation through the inhibition of inflammosomes in Kupffer cells in vivo in the liver IRI model and in vitro in the LPS-treated macrophages. The knockout of CD47 in MSCs increased the expression of NEK7, NLRP3, ASC, and cleaved caspase-1 in Kupffer cells in damaged liver and in macrophages in vitro, and also increased pro-inflammatory cytokines and chemokines, including IL-1β, TNF-α, IL-6, CXCL2, and CXCL10 in macrophages after coculture [176].
Table 1. The effects of MSCs/MSC derivatives on macrophages/Kupffer cells.
Table 1. The effects of MSCs/MSC derivatives on macrophages/Kupffer cells.
MSCs/MSC DerivativesEffect on Kupffer Cells/MacrophagesReferences
PGE2Inhibition of apoptosis via TLR4-ERK1/2-caspase 3 pathway regulation[167]
TGF-βInduction of M2 polarization[168]
Indoleamine 2,3-dioxygenaseInduction of M2 polarization[169]
MitochondriaEnhancement of swallowing ability[170]
Phagocytosis promotion and pro-inflammatory cytokines secretion suppression[171]
MSCsInhibition of Drp-1 dependent mitochondrial fission and promotion of M2 polarization[15]
MSC-derived extracellular vesicles containing IL-10Induction of the expression of the protein tyrosine phosphatase non receptor 22 (PTPN22) and inhibition of the pro-inflammatory macrophage activation[172]
Dead MSCsInduction of M2 polarization[175]
Direct cell-to-cell contacts via CD47 on MSCs and SIRPα on Kupffer cellsInhibition of inflammosomes in Kupffer cells[176]
According to the clinicaltrials.gov website, only one study using mesenchymal stem cells to treat patients with NASH has been enrolled and completed to date (NCT03963921). The complete results of this study have not yet been published. But brief results of this study presented at the EASL Congress 2023 in Vienna showed that the highest dose of human allogeneic liver-derived progenitor cells (HepaStem®) were safe and well tolerated by adult patients with either F3 or F4 NASH (SAF fibrosis score). A preliminary efficacy study demonstrated normalization of ALT and AST serum levels and gradual decreasing of bilirubin and triglyceride levels as well as IL-6 levels and fibrosis over a period of 6 months [177]. In the early- and late-stage NASH mouse models, these human allogeneic liver-derived progenitor cells exhibited pronounced anti-inflammatory and anti-fibrotic effects [178]. In another clinical trial (UMIN000022601), MSCs showed a significant therapeutic effect in patients with NAFLD/NASH. After the autologous adipose tissue-derived regenerative (stem) cells’ (ADRCs) transplantation for the treatment of liver cirrhosis due to non-alcoholic steatohepatitis or fatty liver disease, six out of seven patients showed an improved serum albumin concentration and five out of seven patients showed improved prothrombin activity. In one patient, hepatocyte steatosis decreased by one point, and in another single patient, the lobular inflammation index decreased by one point [179].

5. Conclusions

Data reviewed in this article can be summarized in the following way. It has been unequivocally demonstrated that NASH entails massive hepatocyte death, which is the prime cause and driver of the subsequent pathological inflammation and fibrosis of liver tissue. Decaying hepatocytes release a variety of biologically active substances, including those constituting the so-called damage-associated molecular patterns (DAMPs) capable of activating different types of cells, including immune cells, most importantly resident liver macrophages (Kupffer cells), as well as macrophages recruited from the blood flow. Activated macrophages produce and release a wide spectrum of cytokines and chemokines, some of which are the attractants of neutrophils and cells of the adaptive immune system, while the others intensify inflammation and promote fibrosis. These events drastically exacerbate disease progression. Therefore, Kupffer cells and recruited macrophages present in the liver tissue constitute the first line of the immune response in NASH and trigger all the following pathological developments. If so, they can be regarded as a prospective target for the development of new medicines and therapeutic approaches.
The progression of NAFLD/NASH is strongly dependent on the numerical ratio of the pro-inflammatory (M1) and the anti-inflammatory (M2) macrophages present in the liver tissue and, accordingly, on the cytokines and chemokines they produce. MSCs or MSC derivatives shift the balance towards the anti-inflammatory M2 phenotype both in vitro and in vivo, substantially reducing liver inflammation in the latter case. MSCs/MSC derivatives affect the liver tissue in multiple ways and even if some of the effects are potentially harmful, they can be neutralized. Thus, the M2 macrophages are profibrinogenic, but with MSC therapy fibrosis is offset due to the parallel inhibition of the stellate cell and pro-inflammatory immune cell activation and the hepatocyte protection against apoptosis. The multifactorial and multidirectional nature of the MSCs/MSC derivatives-based NASH therapy is its main advantage over the other Kupffer cell-targeted approaches under development.
Accumulating preclinical and clinical data regarding the pathogenesis of NASH and the treatments currently under development indicates that the pathogenesis of this disease is complicated and, therefore, effective therapeutic approaches probably need to be designed to simultaneously achieve several goals. Targeting a single link within the NASH pathogenetic complex, as was done in the pan-caspase inhibitor trials, not only does not produce the desired therapeutic effect, but may lead to enhanced disease progression. Therefore, multi-target cell therapy using MSCs and MSC derivatives may prove to be effective.

Author Contributions

Writing—original draft preparation, I.V.K.; writing—review and editing, K.N.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The study was performed using “Avogadro” large-scale research facilities, and was financially supported by the Ministry of Education and Science of Russian Federation, Agreement No. 075-15-2021-933, unique project ID: RF00121X0004.

Acknowledgments

Figure 1 and Figure 2 were designed using icons made by Freepik (https://ru.freepik.com/ (accessed on 14 October 2023)).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Polyzos, S.A.; Kountouras, J.; Mantzoros, C.S. Adipose tissue, obesity and non-alcoholic fatty liver disease. Minerva Endocrinol. 2017, 42, 92–108. [Google Scholar] [CrossRef] [PubMed]
  2. Lonardo, A.; Byrne, C.D.; Caldwell, S.H.; Cortez-Pinto, H.; Targher, G. Global epidemiology of nonalcoholic fatty liver disease: Meta-analytic assessment of prevalence, incidence, and outcomes. Hepatology 2016, 64, 1388–1389. [Google Scholar] [CrossRef] [PubMed]
  3. Younossi, Z.M. Non-alcoholic fatty liver disease—A global public health perspective. J. Hepatol. 2019, 70, 531–544. [Google Scholar] [CrossRef] [PubMed]
  4. Bedossa, P. Pathology of non-alcoholic fatty liver disease. Liver Int. 2017, 37 (Suppl. S1), 85–89. [Google Scholar] [CrossRef]
  5. Xu, X.; Poulsen, K.L.; Wu, L.; Liu, S.; Miyata, T.; Song, Q.; Wei, Q.; Zhao, C.; Lin, C.; Yang, J. Targeted therapeutics and novel signaling pathways in non-alcohol-associated fatty liver/steatohepatitis (NAFL/NASH). Signal Transduct. Target. Ther. 2022, 7, 287. [Google Scholar] [CrossRef]
  6. Day, C.; Saksena, S. Non-alcoholic steatohepatitis: Definitions and pathogenesis. J. Gastroenterol. Hepatol. 2002, 17, S377–S384. [Google Scholar] [CrossRef]
  7. Neuschwander-Tetri, B.A. Hepatic lipotoxicity and the pathogenesis of nonalcoholic steatohepatitis: The central role of nontriglyceride fatty acid metabolites. Hepatology 2010, 52, 774–788. [Google Scholar] [CrossRef]
  8. Tilg, H.; Moschen, A.R. Evolution of inflammation in nonalcoholic fatty liver disease: The multiple parallel hits hypothesis. Hepatology 2010, 52, 1836–1846. [Google Scholar] [CrossRef]
  9. Yang, Z.; Wang, L. Current, emerging, and potential therapies for non-alcoholic steatohepatitis. Front. Pharmacol. 2023, 14, 1152042. [Google Scholar] [CrossRef]
  10. Hu, C.; Wu, Z.; Li, L. Mesenchymal stromal cells promote liver regeneration through regulation of immune cells. Int. J. Biol. Sci. 2020, 16, 893–903. [Google Scholar] [CrossRef]
  11. Eguchi, A.; Iwasa, M.; Nakagawa, H. Extracellular vesicles in fatty liver disease and steatohepatitis: Role as biomarkers and therapeutic targets. Liver Int. 2023, 43, 292–298. [Google Scholar] [CrossRef] [PubMed]
  12. Zheng, C.; Sui, B.; Zhang, X.; Hu, J.; Chen, J.; Liu, J.; Wu, D.; Ye, Q.; Xiang, L.; Qiu, X.; et al. Apoptotic vesicles restore liver macrophage homeostasis to counteract type 2 diabetes. J. Extracell. Vesicles 2021, 10, e12109. [Google Scholar] [CrossRef] [PubMed]
  13. Yarygin, K.N.; Lupatov, A.Y.; Sukhikh, G.T. Modulation of Immune Responses by Mesenchymal Stromal Cells. Bull. Exp. Biol. Med. 2016, 161, 561–565. [Google Scholar] [CrossRef] [PubMed]
  14. Lu, D.; Xu, Y.; Liu, Q.; Zhang, Q. Mesenchymal Stem Cell-Macrophage Crosstalk and Maintenance of Inflammatory Microenvironment Homeostasis. Front. Cell Dev. Biol. 2021, 9, 681171. [Google Scholar] [CrossRef]
  15. Shang, L.C.; Wang, M.; Liu, Y.; Zhu, X.; Wang, S. MSCs Ameliorate Hepatic IR Injury by Modulating Phenotypic Transformation of Kupffer Cells Through Drp-1 Dependent Mitochondrial Dynamics. Stem Cell Rev. Rep. 2023, 19, 1965–1980. [Google Scholar] [CrossRef]
  16. Luedde, T.; Kaplowitz, N.; Schwabe, R.F. Cell death and cell death responses in liver disease: Mechanisms and clinical relevance. Gastroenterology 2014, 147, 765–783.e4. [Google Scholar] [CrossRef]
  17. Amarapurka, D.N.; Amarapurkar, A.D.; Patel, N.D.; Agal, S.; Baigal, R.; Gupte, P.; Pramanik, S. Nonalcoholic steatohepatitis (NASH) with diabetes: Predictors of liver fibrosis. Ann. Hepatol. 2006, 5, 30–33. [Google Scholar] [CrossRef]
  18. Angulo, P.; Keach, J.C.; Batts, K.P.; Lindor, K.D. Independent predictors of liver fibrosis in patients with nonalcoholic steatohepatitis. Hepatology 1999, 30, 1356–1362. [Google Scholar] [CrossRef]
  19. Lee, T.H.; Kim, W.R.; Benson, J.T.; Therneau, T.M.; Melton, L.J., 3rd. Serum aminotransferase activity and mortality risk in a United States community. Hepatology 2008, 47, 880–887. [Google Scholar] [CrossRef]
  20. Kim, H.C.; Nam, C.M.; Jee, S.H.; Han, K.H.; Oh, D.K.; Suh, I. Normal serum aminotransferase concentration and risk of mortality from liver diseases: Prospective cohort study. BMJ 2004, 328, 983. [Google Scholar] [CrossRef]
  21. Petrasek, J.; Iracheta-Vellve, A.; Csak, T.; Satishchandran, A.; Kodys, K.; Kurt-Jones, E.A.; Fitzgerald, K.A.; Szabo, G. STING-IRF3 Pathway Links Endoplasmic Reticulum Stress with Hepatocyte Apoptosis in Early Alcoholic Liver Disease. Proc. Natl. Acad. Sci. USA 2013, 110, 16544–16549. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, S.; Ni, H.M.; Dorko, K.; Kumer, S.C.; Schmitt, T.M.; Nawabi, A.; Komatsu, M.; Huang, H.; Ding, W.X. Increased Hepatic Receptor Interacting Protein Kinase 3 Expression Due to Impaired Proteasomal Functions Contributes to Alcohol-Induced Steatosis and Liver Injury. Oncotarget 2016, 7, 17681–17698. [Google Scholar] [CrossRef]
  23. Khambu, B.; Wang, L.; Zhang, H.; Yin, X.-M. The Activation and Function of Autophagy in Alcoholic Liver Disease. Curr. Mol. Pharmacol. 2017, 10, 165–171. [Google Scholar] [CrossRef]
  24. Heo, M.J.; Kim, T.H.; You, J.S.; Blaya, D.; Sancho-Bru, P.; Kim, S.G. Alcohol Dysregulates MiR-148a in Hepatocytes through FoxO1, Facilitating Pyroptosis via TXNIP Overexpression. Gut 2019, 68, 708–720. [Google Scholar] [CrossRef] [PubMed]
  25. Zhou, Z.; Ye, T.J.; DeCaro, E.; Buehler, B.; Stahl, Z.; Bonavita, G.; Daniels, M.; You, M. Intestinal SIRT1 Deficiency Protects Mice from Ethanol-Induced Liver Injury by Mitigating Ferroptosis. Am. J. Pathol. 2020, 190, 82–92. [Google Scholar] [CrossRef] [PubMed]
  26. Shojaie, L.; Iorga, A.; Dara, L. Cell Death in Liver Diseases: A Review. Int. J. Mol. Sci. 2020, 21, 9682. [Google Scholar] [CrossRef] [PubMed]
  27. Jaeschke, H.; Ramachandran, A.; Chao, X.; Ding, W.X. Emerging and Established Modes of Cell Death during Acetaminophen-Induced Liver Injury. Arch. Toxicol. 2019, 93, 3491–3502. [Google Scholar] [CrossRef] [PubMed]
  28. Cazanave, S.C.; Gores, G.J. Mechanisms and clinical implications of hepatocyte lipoapoptosis. Clin. Lipidol. 2010, 5, 71–85. [Google Scholar] [CrossRef]
  29. Feldstein, A.E.; Canbay, A.; Angulo, P.; Taniai, M.; Burgart, L.J.; Lindor, K.D.; Gores, G.J. Hepatocyte apoptosis and fas expression are prominent features of human nonalcoholic steatohepatitis. Gastroenterology 2003, 125, 437–443. [Google Scholar] [CrossRef]
  30. Barreyro, F.J.; Holod, S.; Finocchietto, P.V.; Camino, A.M.; Aquino, J.B.; Avagnina, A.; Carreras, M.C.; Poderoso, J.J.; Gores, G.J. The pan-caspase inhibitor Emricasan (IDN-6556) decreases liver injury and fibrosis in a murine model of non-alcoholic steatohepatitis. Liver Int. 2015, 35, 953–966. [Google Scholar] [CrossRef]
  31. Ratziu, V.; Sheikh, M.Y.; Sanyal, A.J.; Lim, J.K.; Conjeevaram, H.; Chalasani, N.; Abdelmalek, M.; Bakken, A.; Renou, C.; Palmer, M.; et al. A phase 2, randomized, double-blind, placebo-controlled study of GS-9450 in subjects with nonalcoholic steatohepatitis. Hepatology 2012, 55, 419–428. [Google Scholar] [CrossRef] [PubMed]
  32. Shiffman, M.; Freilich, B.; Vuppalanchi, R.; Watt, K.; Chan, J.L.; Spada, A.; Hagerty, D.T.; Schiff, E. Randomised clinical trial: Emricasan versus placebo significantly decreases ALT and caspase 3/7 activation in subjects with non-alcoholic fatty liver disease. Aliment. Pharmacol. Ther. 2019, 49, 64–73. [Google Scholar] [CrossRef] [PubMed]
  33. Harrison, S.A.; Goodman, Z.; Jabbar, A.; Vemulapalli, R.; Younes, Z.H.; Freilich, B.; Sheikh, M.Y.; Schattenberg, J.M.; Kayali, Z.; Zivony, A.; et al. A randomized, placebo-controlled trial of emricasan in patients with NASH and F1-F3 fibrosis. J. Hepatol. 2020, 72, 816–827. [Google Scholar] [CrossRef] [PubMed]
  34. Henao-Mejia, J.; Elinav, E.; Jin, C.; Hao, L.; Mehal, W.Z.; Strowig, T.; Thaiss, C.A.; Kau, A.L.; Eisenbarth, S.C.; Jurczak, M.J.; et al. Inflammasome-mediated dysbiosis regulates progression of NAFLD and obesity. Nature 2012, 482, 179–185. [Google Scholar] [CrossRef] [PubMed]
  35. Mridha, A.R.; Wree, A.; Robertson, A.A.B.; Yeh, M.M.; Johnson, C.D.; Van Rooyen, D.M.; Haczeyni, F.; Teoh, N.C.; Savard, C.; Ioannou, G.N.; et al. NLRP3 inflammasome blockade reduces liver inflammation and fibrosis in experimental NASH in mice. J. Hepatol. 2017, 66, 1037–1046. [Google Scholar] [CrossRef]
  36. Xu, B.; Jiang, M.; Chu, Y.; Wang, W.; Chen, D.; Li, X.; Zhang, Z.; Zhang, D.; Fan, D.; Nie, Y.; et al. Gasdermin D plays a key role as a pyroptosis executor of non-alcoholic steatohepatitis in humans and mice. J. Hepatol. 2018, 68, 773–782. [Google Scholar] [CrossRef]
  37. Nelson, J.E.; Wilson, L.; Brunt, E.M.; Yeh, M.M.; Kleiner, D.E.; Unalp-Arida, A.; Kowdley, K.V. Nonalcoholic Steatohepatitis Clinical Research Network. Relationship between the pattern of hepatic iron deposition and histological severity in nonalcoholic fatty liver disease. Hepatology 2011, 53, 448–457. [Google Scholar] [CrossRef]
  38. Puri, P.; Baillie, R.A.; Wiest, M.M.; Mirshahi, F.; Choudhury, J.; Cheung, O.; Sargeant, C.; Contos, M.J.; Sanyal, A.J. A lipidomic analysis of nonalcoholic fatty liver disease. Hepatology 2007, 46, 1081–1090. [Google Scholar] [CrossRef]
  39. Caballero, F.; Fernandez, A.; De Lacy, A.M.; Fernandez-Checa, J.C.; Caballeria, J.; Garcia-Ruiz, C. Enhanced free cholesterol, SREBP-2 and StAR expression in human NASH. J. Hepatol. 2009, 50, 789–796. [Google Scholar] [CrossRef]
  40. Gan, L.T.; Van Rooyen, D.M.; Koina, M.E.; McCuskey, R.S.; Teoh, N.C.; Farrell, G.C. Hepatocyte free cholesterol lipotoxicity results from JNK1-mediated mitochondrial injury and is HMGB1 and TLR4-dependent. J. Hepatol. 2014, 61, 1376–1384. [Google Scholar] [CrossRef]
  41. Eguchi, A.; Wree, A.; Feldstein, A.E. Biomarkers of liver cell death. J. Hepatol. 2014, 60, 1063–1074. [Google Scholar] [CrossRef]
  42. Matzinger, P. Tolerance, danger, and the extended family. Annu. Rev. Immunol. 1994, 12, 991–1045. [Google Scholar] [CrossRef] [PubMed]
  43. Matzinger, P. The danger model: A renewed sense of self. Science 2002, 296, 301–305. [Google Scholar] [CrossRef]
  44. Chen, G.Y.; Nuñez, G. Sterile inflammation: Sensing and reacting to damage. Nat. Rev. Immunol. 2010, 10, 826–837. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, W.; Lavine, K.J.; Epelman, S.; Evans, S.A.; Weinheimer, C.J.; Barger, P.M.; Mann, D.L. Necrotic myocardial cells release damage-associated molecular patterns that provoke fibroblast activation in vitro and trigger myocardial inflammation and fibrosis in vivo. J. Am. Heart Assoc. 2015, 4, e001993. [Google Scholar] [CrossRef] [PubMed]
  46. An, P.; Wei, L.L.; Zhao, S.; Sverdlov, D.Y.; Vaid, K.A.; Miyamoto, M.; Kuramitsu, K.; Lai, M.; Popov, Y.V. Hepatocyte mitochondria-derived danger signals directly activate hepatic stellate cells and drive progression of liver fibrosis. Nat. Commun. 2020, 11, 2362. [Google Scholar] [CrossRef] [PubMed]
  47. Sun, S.; Sursal, T.; Adibnia, Y.; Zhao, C.; Zheng, Y.; Li, H.; Otterbein, L.E.; Hauser, C.J.; Itagaki, K. Mitochondrial DAMPs increase endothelial permeability through neutrophil dependent and independent pathways. PLoS ONE 2013, 8, e59989. [Google Scholar] [CrossRef]
  48. Woolbright, B.L.; Jaeschke, H. Mechanisms of Inflammatory Liver Injury and Drug-Induced Hepatotoxicity. Curr. Pharmacol. Rep. 2018, 4, 346–357. [Google Scholar] [CrossRef]
  49. Gong, T.; Liu, L.; Jiang, W.; Zhou, R. DAMP-sensing receptors in sterile inflammation and inflammatory diseases. Nat. Rev. Immunol. 2020, 20, 95–112. [Google Scholar] [CrossRef]
  50. Xiao, F.; Waldrop, S.L.; Khimji, A.K.; Kilic, G. Pannexin1 contributes to pathophysiological ATP release in lipoapoptosis induced by saturated free fatty acids in liver cells. Am. J. Physiol. Cell Physiol. 2012, 303, C1034–C1044. [Google Scholar] [CrossRef]
  51. Xiao, F.; Waldrop, S.L.; Bronk, S.F.; Gores, G.J.; Davis, L.S.; Kilic, G. Lipoapoptosis induced by saturated free fatty acids stimulates monocyte migration: A novel role for Pannexin1 in liver cells. Purinergic Signal. 2015, 11, 347–359. [Google Scholar] [CrossRef] [PubMed]
  52. McDonald, B.; Pittman, K.; Menezes, G.B.; Hirota, S.A.; Slaba, I.; Waterhouse, C.C.; Beck, P.L.; Muruve, D.A.; Kubes, P. Intravascular danger signals guide neutrophils to sites of sterile inflammation. Science 2010, 330, 362–366, Erratum in Science 2011, 331, 1517. [Google Scholar] [CrossRef] [PubMed]
  53. Begriche, K.; Massart, J.; Robin, M.A.; Bonnet, F.; Fromenty, B. Mitochondrial adaptations and dysfunctions in nonalcoholic fatty liver disease. Hepatology 2013, 58, 1497–1507. [Google Scholar] [CrossRef]
  54. Garcia-Martinez, I.; Santoro, N.; Chen, Y.; Hoque, R.; Ouyang, X.; Caprio, S.; Shlomchik, M.J.; Coffman, R.L.; Candia, A.; Mehal, W.Z. Hepatocyte mitochondrial DNA drives nonalcoholic steatohepatitis by activation of TLR9. J. Clin. Investig. 2016, 126, 859–864. [Google Scholar] [CrossRef]
  55. Handa, P.; Vemulakonda, A.; Kowdley, K.V.; Uribe, M.; Méndez-Sánchez, N. Mitochondrial DNA from hepatocytes as a ligand for TLR9: Drivers of nonalcoholic steatohepatitis? World J. Gastroenterol. 2016, 22, 6965–6971. [Google Scholar] [CrossRef] [PubMed]
  56. Omary, M.B.; Coulombe, P.A.; McLean, W.H. Intermediate filament proteins and their associated diseases. N. Engl. J. Med. 2004, 351, 2087–2100. [Google Scholar] [CrossRef]
  57. Omary, M.B.; Ku, N.O.; Strnad, P.; Hanada, S. Toward unraveling the complexity of simple epithelial keratins in human disease. J. Clin. Investig. 2009, 119, 1794–1805. [Google Scholar] [CrossRef]
  58. MacFarlane, M.; Merrison, W.; Dinsdale, D.; Cohen, G.M. Active caspases and cleaved cytokeratins are sequestered into cytoplasmic inclusions in TRAIL-induced apoptosis. J. Cell Biol. 2000, 148, 1239–1254. [Google Scholar] [CrossRef]
  59. Weerasinghe, S.V.; Ku, N.O.; Altshuler, P.J.; Kwan, R.; Omary, M.B. Mutation of caspase-digestion sites in keratin 18 interferes with filament reorganization, and predisposes to hepatocyte necrosis and loss of membrane integrity. J. Cell Sci. 2014, 127, 1464–1475. [Google Scholar] [CrossRef]
  60. Ku, N.O.; Strnad, P.; Bantel, H.; Omary, M.B. Keratins: Biomarkers and modulators of apoptotic and necrotic cell death in the liver. Hepatology 2016, 64, 966–976. [Google Scholar] [CrossRef]
  61. Li, L.; Chen, L.; Hu, L.; Liu, Y.; Sun, H.Y.; Tang, J.; Hou, Y.J.; Chang, Y.X.; Tu, Q.Q.; Feng, G.S.; et al. Nuclear factor high-mobility group box1 mediating the activation of Toll-like receptor 4 signaling in hepatocytes in the early stage of nonalcoholic fatty liver disease in mice. Hepatology 2011, 54, 1620–1630. [Google Scholar] [CrossRef] [PubMed]
  62. Ibrahim, S.H.; Hirsova, P.; Gores, G.J. Non-alcoholic steatohepatitis pathogenesis: Sublethal hepatocyte injury as a driver of liver inflammation. Gut 2018, 67, 963–972. [Google Scholar] [CrossRef] [PubMed]
  63. Alisi, A.; Nobili, V.; Ceccarelli, S.; Panera, N.; De Stefanis, C.; De Vito, R.; Vitali, R.; Bedogni, G.; Balsano, C.; Cucchiara, S.; et al. Plasma high mobility group box 1 protein reflects fibrosis in pediatric nonalcoholic fatty liver disease. Expert Rev. Mol. Diagn. 2014, 14, 763–771. [Google Scholar] [CrossRef] [PubMed]
  64. Scaffidi, P.; Misteli, T.; Bianchi, M.E. Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 2002, 418, 191–195, Erratum in Nature 2010, 467, 622. [Google Scholar] [CrossRef] [PubMed]
  65. Yang, H.; Ochani, M.; Li, J.; Qiang, X.; Tanovic, M.; Harris, H.E.; Susarla, S.M.; Ulloa, L.; Wang, H.; DiRaimo, R.; et al. Reversing established sepsis with antagonists of endogenous high-mobility group box 1. Proc. Natl. Acad. Sci. USA 2004, 101, 296–301. [Google Scholar] [CrossRef]
  66. Ge, X.; Arriazu, E.; Magdaleno, F.; Antoine, D.J.; Cruz, R.D.; Theise, N.; Nieto, N. High Mobility Group Box-1 Drives Fibrosis Progression Signaling via the Receptor for Advanced Glycation End Products in Mice. Hepatology 2018, 68, 2380–2404. [Google Scholar] [CrossRef]
  67. Arriazu, E.; Ge, X.; Leung, T.M.; Magdaleno, F.; Lopategi, A.; Lu, Y.; Kitamura, N.; Urtasun, R.; Theise, N.; Antoine, D.J.; et al. Signalling via the osteopontin and high mobility group box-1 axis drives the fibrogenic response to liver injury. Gut 2017, 66, 1123–1137, Erratum in Gut 2020, 69, e2. [Google Scholar] [CrossRef]
  68. Vonghia, L.; Michielsen, P.; Francque, S. Immunological mechanisms in the pathophysiology of non-alcoholic steatohepatitis. Int. J. Mol. Sci. 2013, 14, 19867–19890. [Google Scholar] [CrossRef]
  69. Ramadori, G.; Christ, B. Cytokines and the hepatic acute-phase response. Semin. Liver Dis. 1999, 19, 141–155. [Google Scholar] [CrossRef]
  70. Mollica, M.P.; Lionetti, L.; Putti, R.; Cavaliere, G.; Gaita, M.; Barletta, A. From chronic overfeeding to hepatic injury: Role of endoplasmic reticulum stress and inflammation. Nutr. Metab. Cardiovasc. Dis. 2011, 21, 222–230. [Google Scholar] [CrossRef]
  71. Winkler, S.; Borkham-Kamphorst, E.; Stock, P.; Brückner, S.; Dollinger, M.; Weiskirchen, R.; Christ, B. Human mesenchymal stem cells towards non-alcoholic steatohepatitis in an immunodeficient mouse model. Exp. Cell Res. 2014, 326, 230–239. [Google Scholar] [CrossRef] [PubMed]
  72. Pelz, S.; Stock, P.; Brückner, S.; Christ, B. A methionine-choline-deficient diet elicits NASH in the immunodeficient mouse featuring a model for hepatic cell transplantation. Exp. Cell Res. 2012, 318, 276–287. [Google Scholar] [CrossRef] [PubMed]
  73. Bouwens, L.; Baekeland, M.; De Zanger, R.; Wisse, E. Quantitation, tissue distribution and proliferation kinetics of Kupffer cells in normal rat liver. Hepatology 1986, 6, 718–722. [Google Scholar] [CrossRef] [PubMed]
  74. Deppermann, C.; Kratofil, R.M.; Peiseler, M.; David, B.A.; Zindel, J.; Castanheira, F.V.E.S.; van der Wal, F.; Carestia, A.; Jenne, C.N.; Marth, J.D.; et al. Macrophage galactose lectin is critical for Kupffer cells to clear aged platelets. J. Exp. Med. 2020, 217, e20190723. [Google Scholar] [CrossRef] [PubMed]
  75. Brubaker, W.D.; Crane, A.; Johansson, J.U.; Yen, K.; Garfinkel, K.; Mastroeni, D.; Asok, P.; Bradt, B.; Sabbagh, M.; Wallace, T.L.; et al. Peripheral complement interactions with amyloid β peptide: Erythrocyte clearance mechanisms. Alzheimers Dement. 2017, 13, 1397–1409. [Google Scholar] [CrossRef] [PubMed]
  76. Willekens, F.L.; Were, J.M.; Kruijt, J.K.; Roerdinkholder-Stoelwinder, B.; Groenen-Döpp, Y.A.; van den Bos, A.G.; Bosman, G.J.; van Berkel, T.J. Liver Kupffer cells rapidly remove red blood cell-derived vesicles from the circulation by scavenger receptors. Blood 2005, 105, 2141–2145. [Google Scholar] [CrossRef]
  77. Terpstra, V.; van Berkel, T.J. Scavenger receptors on liver Kupffer cells mediate the in vivo uptake of oxidatively damaged red blood cells in mice. Blood 2000, 95, 2157–2163. [Google Scholar] [CrossRef]
  78. Kristiansen, M.; Graversen, J.H.; Jacobsen, C.; Sonne, O.; Hoffman, H.J.; Law, S.K.; Moestrup, S.K. Identification of the haemoglobin scavenger receptor. Nature 2001, 409, 198–201. [Google Scholar] [CrossRef]
  79. Theurl, I.; Hilgendorf, I.; Nairz, M.; Tymoszuk, P.; Haschka, D.; Asshoff, M.; He, S.; Gerhardt, L.M.; Holderried, T.A.; Seifert, M.; et al. On-demand erythrocyte disposal and iron recycling requires transient macrophages in the liver. Nat. Med. 2016, 22, 945–951. [Google Scholar] [CrossRef]
  80. Scott, C.L.; Guilliams, M. The role of Kupffer cells in hepatic iron and lipid metabolism. J. Hepatol. 2018, 69, 1197–1199. [Google Scholar] [CrossRef]
  81. Wang, Y.; van der Tuin, S.; Tjeerdema, N.; van Dam, A.D.; Rensen, S.S.; Hendrikx, T.; Berbée, J.F.; Atanasovska, B.; Fu, J.; Hoekstra, M.; et al. Plasma cholesteryl ester transfer protein is predominantly derived from Kupffer cells. Hepatology 2015, 62, 1710–1722. [Google Scholar] [CrossRef] [PubMed]
  82. Helmy, K.Y.; Katschke, K.J., Jr.; Gorgani, N.N.; Kljavin, N.M.; Elliott, J.M.; Diehl, L.; Scales, S.J.; Ghilardi, N.; van Lookeren Campagne, M. CRIg: A macrophage complement receptor required for phagocytosis of circulating pathogens. Cell 2006, 124, 915–927. [Google Scholar] [CrossRef] [PubMed]
  83. Zeng, Z.; Surewaard, B.G.; Wong, C.H.; Geoghegan, J.A.; Jenne, C.N.; Kubes, P. CRIg Functions as a Macrophage Pattern Recognition Receptor to Directly Bind and Capture Blood-Borne Gram-Positive Bacteria. Cell Host Microbe 2016, 20, 99–106. [Google Scholar] [CrossRef] [PubMed]
  84. You, Q.; Cheng, L.; Kedl, R.M.; Ju, C. Mechanism of T cell tolerance induction by murine hepatic Kupffer cells. Hepatology 2008, 48, 978–990. [Google Scholar] [CrossRef]
  85. Heymann, F.; Peusquens, J.; Ludwig-Portugall, I.; Kohlhepp, M.; Ergen, C.; Niemietz, P.; Martin, C.; van Rooijen, N.; Ochando, J.C.; Randolph, G.J.; et al. Liver inflammation abrogates immunological tolerance induced by Kupffer cells. Hepatology 2015, 62, 279–291. [Google Scholar] [CrossRef]
  86. Li, W.; Chang, N.; Li, L. Heterogeneity and Function of Kupffer Cells in Liver Injury. Front. Immunol. 2022, 13, 940867. [Google Scholar] [CrossRef]
  87. Gomez Perdiguero, E.; Klapproth, K.; Schulz, C.; Busch, K.; Azzoni, E.; Crozet, L.; Garner, H.; Trouillet, C.; de Bruijn, M.F.; Geissmann, F.; et al. Tissue-resident macrophages originate from yolk-sac-derived erythro-myeloid progenitors. Nature 2015, 518, 547–551. [Google Scholar] [CrossRef]
  88. Bertrand, J.Y.; Jalil, A.; Klaine, M.; Jung, S.; Cumano, A.; Godin, I. Three pathways to mature macrophages in the early mouse yolk sac. Blood 2005, 106, 3004–3011. [Google Scholar] [CrossRef]
  89. Schulz, C.; Gomez Perdiguero, E.; Chorro, L.; Szabo-Rogers, H.; Cagnard, N.; Kierdorf, K.; Prinz, M.; Wu, B.; Jacobsen, S.E.; Pollard, J.W.; et al. A lineage of myeloid cells independent of Myb and hematopoietic stem cells. Science 2012, 336, 86–90. [Google Scholar] [CrossRef]
  90. MacParland, S.A.; Liu, J.C.; Ma, X.Z.; Innes, B.T.; Bartczak, A.M.; Gage, B.K.; Manuel, J.; Khuu, N.; Echeverri, J.; Linares, I.; et al. Single cell RNA sequencing of human liver reveals distinct intrahepatic macrophage populations. Nat. Commun. 2018, 9, 4383. [Google Scholar] [CrossRef]
  91. Aizarani, N.; Saviano, A.; Sagar Mailly, L.; Durand, S.; Herman, J.S.; Pessaux, P.; Baumert, T.F.; Grün, D. A human liver cell atlas reveals heterogeneity and epithelial progenitors. Nature 2019, 572, 199–204. [Google Scholar] [CrossRef] [PubMed]
  92. Ramachandran, P.; Dobie, R.; Wilson-Kanamori, J.R.; Dora, E.F.; Henderson, B.E.P.; Luu, N.T.; Portman, J.R.; Matchett, K.P.; Brice, M.; Marwick, J.A.; et al. Resolving the fibrotic niche of human liver cirrhosis at single-cell level. Nature 2019, 575, 512–518. [Google Scholar] [CrossRef] [PubMed]
  93. Wu, X.; Hollingshead, N.; Roberto, J.; Knupp, A.; Kenerson, H.; Chen, A.; Strickland, I.; Horton, H.; Yeung, R.; Soysa, R.; et al. Human Liver Macrophage Subsets Defined by CD32. Front. Immunol. 2020, 11, 2108. [Google Scholar] [CrossRef] [PubMed]
  94. Guilliams, M.; Bonnardel, J.; Haest, B.; Vanderborght, B.; Wagner, C.; Remmerie, A.; Bujko, A.; Martens, L.; Thoné, T.; Browaeys, R.; et al. Spatial proteogenomics reveals distinct and evolutionarily conserved hepatic macrophage niches. Cell 2022, 185, 379–396.e38. [Google Scholar] [CrossRef] [PubMed]
  95. Blériot, C.; Barreby, E.; Dunsmore, G.; Ballaire, R.; Chakarov, S.; Ficht, X.; De Simone, G.; Andreata, F.; Fumagalli, V.; Guo, W.; et al. A subset of Kupffer cells regulates metabolism through the expression of CD36. Immunity 2021, 54, 2101–2116.e6. [Google Scholar] [CrossRef]
  96. Martrus, G.; Goebels, H.; Langeneckert, A.E.; Kah, J.; Flomm, F.; Ziegler, A.E.; Niehrs, A.; Löbl, S.M.; Russu, K.; Hess, L.U.; et al. CD49a Expression Identifies a Subset of Intrahepatic Macrophages in Humans. Front. Immunol. 2019, 10, 1247. [Google Scholar] [CrossRef]
  97. Krenkel, O.; Tacke, F. Liver macrophages in tissue homeostasis and disease. Nat. Rev. Immunol. 2017, 17, 306–321. [Google Scholar] [CrossRef]
  98. Stutchfield, B.M.; Antoine, D.J.; Mackinnon, A.C.; Gow, D.J.; Bain, C.C.; Hawley, C.A.; Hughes, M.J.; Francis, B.; Wojtacha, D.; Man, T.Y.; et al. CSF1 Restores Innate Immunity After Liver Injury in Mice and Serum Levels Indicate Outcomes of Patients With Acute Liver Failure. Gastroenterology 2015, 149, 1896–1909.e14. [Google Scholar] [CrossRef]
  99. Nascimento, M.; Huang, S.C.; Smith, A.; Everts, B.; Lam, W.; Bassity, E.; Gautier, E.L.; Randolph, G.J.; Pearce, E.J. Ly6Chi monocyte recruitment is responsible for Th2 associated host-protective macrophage accumulation in liver inflammation due to schistosomiasis. PLoS Pathog. 2014, 10, e1004282. [Google Scholar] [CrossRef]
  100. Zigmond, E.; Samia-Grinberg, S.; Pasmanik-Chor, M.; Brazowski, E.; Shibolet, O.; Halpern, Z.; Varol, C. Infiltrating monocyte-derived macrophages and resident kupffer cells display different ontogeny and functions in acute liver injury. J. Immunol. 2014, 193, 344–353. [Google Scholar] [CrossRef]
  101. Scott, C.L.; Zheng, F.; De Baetselier, P.; Martens, L.; Saeys, Y.; De Prijck, S.; Lippens, S.; Abels, C.; Schoonooghe, S.; Raes, G.; et al. Bone marrow-derived monocytes give rise to self-renewing and fully differentiated Kupffer cells. Nat. Commun. 2016, 7, 10321. [Google Scholar] [CrossRef] [PubMed]
  102. Fogg, D.K.; Sibon, C.; Miled, C.; Jung, S.; Aucouturier, P.; Littman, D.R.; Cumano, A.; Geissmann, F. A clonogenic bone marrow progenitor specific for macrophages and dendritic cells. Science 2006, 311, 83–87, Erratum in Science 2006, 311, 1242. [Google Scholar] [CrossRef] [PubMed]
  103. Heymann, F.; Hammerich, L.; Storch, D.; Bartneck, M.; Huss, S.; Rüsseler, V.; Gassler, N.; Lira, S.A.; Luedde, T.; Trautwein, C.; et al. Hepatic macrophage migration and differentiation critical for liver fibrosis is mediated by the chemokine receptor C-C motif chemokine receptor 8 in mice. Hepatology 2012, 55, 898–909. [Google Scholar] [CrossRef] [PubMed]
  104. Miura, K.; Yang, L.; van Rooijen, N.; Ohnishi, H.; Seki, E. Hepatic recruitment of macrophages promotes nonalcoholic steatohepatitis through CCR2. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G1310–G1321. [Google Scholar] [CrossRef] [PubMed]
  105. Karlmark, K.R.; Weiskirchen, R.; Zimmermann, H.W.; Gassler, N.; Ginhoux, F.; Weber, C.; Merad, M.; Luedde, T.; Trautwein, C.; Tacke, F. Hepatic recruitment of the inflammatory Gr1+ monocyte subset upon liver injury promotes hepatic fibrosis. Hepatology 2009, 50, 261–274. [Google Scholar] [CrossRef]
  106. Nakamoto, N.; Ebinuma, H.; Kanai, T.; Chu, P.S.; Ono, Y.; Mikami, Y.; Ojiro, K.; Lipp, M.; Love, P.E.; Saito, H.; et al. CCR9+ macrophages are required for acute liver inflammation in mouse models of hepatitis. Gastroenterology 2012, 142, 366–376. [Google Scholar] [CrossRef]
  107. Chu, P.S.; Nakamoto, N.; Ebinuma, H.; Usui, S.; Saeki, K.; Matsumoto, A.; Mikami, Y.; Sugiyama, K.; Tomita, K.; Kanai, T.; et al. C-C motif chemokine receptor 9 positive macrophages activate hepatic stellate cells and promote liver fibrosis in mice. Hepatology 2013, 58, 337–350. [Google Scholar] [CrossRef]
  108. Baeck, C.; Wehr, A.; Karlmark, K.R.; Heymann, F.; Vucur, M.; Gassler, N.; Huss, S.; Klussmann, S.; Eulberg, D.; Luedde, T.; et al. Pharmacological inhibition of the chemokine CCL2 (MCP-1) diminishes liver macrophage infiltration and steatohepatitis in chronic hepatic injury. Gut 2012, 61, 416–426. [Google Scholar] [CrossRef]
  109. Ramachandran, P.; Pellicoro, A.; Vernon, M.A.; Boulter, L.; Aucott, R.L.; Ali, A.; Hartland, S.N.; Snowdon, V.K.; Cappon, A.; Gordon-Walker, T.T.; et al. Differential Ly-6C expression identifies the recruited macrophage phenotype, which orchestrates the regression of murine liver fibrosis. Proc. Natl. Acad. Sci. USA 2012, 109, E3186–E3195. [Google Scholar] [CrossRef]
  110. Guillot, A.; Tacke, F. Liver Macrophages: Old Dogmas and New Insights. Hepatol. Commun. 2019, 3, 730–743. [Google Scholar] [CrossRef]
  111. McGettigan, B.; McMahan, R.; Orlicky, D.; Burchill, M.; Danhorn, T.; Francis, P.; Cheng, L.L.; Golden-Mason, L.; Jakubzick, C.V.; Rosen, H.R. Dietary Lipids Differentially Shape Nonalcoholic Steatohepatitis Progression and the Transcriptome of Kupffer Cells and Infiltrating Macrophages. Hepatology 2019, 70, 67–83. [Google Scholar] [CrossRef] [PubMed]
  112. Yu, Y.; Liu, Y.; An, W.; Song, J.; Zhang, Y.; Zhao, X. STING-mediated inflammation in Kupffer cells contributes to progression of nonalcoholic steatohepatitis. J. Clin. Investig. 2019, 129, 546–555. [Google Scholar] [CrossRef] [PubMed]
  113. Heymann, F.; Tacke, F. Immunology in the liver—from homeostasis to disease. Nat. Rev. Gastroenterol. Hepatol. 2016, 13, 88–110. [Google Scholar] [CrossRef]
  114. Marra, F.; Tacke, F. Roles for chemokines in liver disease. Gastroenterology 2014, 147, 577–594.e1. [Google Scholar] [CrossRef] [PubMed]
  115. Dambach, D.M.; Watson, L.M.; Gray, K.R.; Durham, S.K.; Laskin, D.L. Role of CCR2 in macrophage migration into the liver during acetaminophen-induced hepatotoxicity in the mouse. Hepatology 2002, 35, 1093–1103. [Google Scholar] [CrossRef] [PubMed]
  116. Huang, H.; Tohme, S.; Al-Khafaji, A.B.; Tai, S.; Loughran, P.; Chen, L.; Wang, S.; Kim, J.; Billiar, T.; Wang, Y.; et al. Damage-associated molecular pattern-activated neutrophil extracellular trap exacerbates sterile inflammatory liver injury. Hepatology 2015, 62, 600–614. [Google Scholar] [CrossRef]
  117. Wehr, A.; Baeck, C.; Heymann, F.; Niemietz, P.M.; Hammerich, L.; Martin, C.; Zimmermann, H.W.; Pack, O.; Gassler, N.; Hittatiya, K.; et al. Chemokine receptor CXCR6-dependent hepatic NK T Cell accumulation promotes inflammation and liver fibrosis. J. Immunol. 2013, 190, 5226–5236. [Google Scholar] [CrossRef]
  118. Mossanen, J.C.; Krenkel, O.; Ergen, C.; Govaere, O.; Liepelt, A.; Puengel, T.; Heymann, F.; Kalthoff, S.; Lefebvre, E.; Eulberg, D.; et al. Chemokine (C-C motif) receptor 2-positive monocytes aggravate the early phase of acetaminophen-induced acute liver injury. Hepatology 2016, 64, 1667–1682. [Google Scholar] [CrossRef]
  119. Xia, C.; Razavi, M.; Rao, X.; Braunstein, Z.; Mao, H.; Toomey, A.C.; Wang, Y.; Simon, D.I.; Zhao, S.; Rajagopalan, S.; et al. MRP14 enhances the ability of macrophage to recruit T cells and promotes obesity-induced insulin resistance. Int. J. Obes. 2019, 43, 2434–2447. [Google Scholar] [CrossRef]
  120. Liaskou, E.; Zimmermann, H.W.; Li, K.K.; Oo, Y.H.; Suresh, S.; Stamataki, Z.; Qureshi, O.; Lalor, P.F.; Shaw, J.; Syn, W.K.; et al. Monocyte subsets in human liver disease show distinct phenotypic and functional characteristics. Hepatology 2013, 57, 385–398. [Google Scholar] [CrossRef]
  121. Ogura, Y.; Sutterwala, F.S.; Flavell, R.A. The inflammasome: First line of the immune response to cell stress. Cell 2006, 126, 659–662. [Google Scholar] [CrossRef] [PubMed]
  122. Szabo, G.; Petrasek, J. Inflammasome activation and function in liver disease. Nat. Rev. Gastroenterol. Hepatol. 2015, 12, 387–400. [Google Scholar] [CrossRef] [PubMed]
  123. Wen, H.; Gris, D.; Lei, Y.; Jha, S.; Zhang, L.; Huang, M.T.; Brickey, W.J.; Ting, J.P. Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat. Immunol. 2011, 12, 408–415. [Google Scholar] [CrossRef] [PubMed]
  124. Yu, X.; Lan, P.; Hou, X.; Han, Q.; Lu, N.; Li, T.; Jiao, C.; Zhang, J.; Zhang, C.; Tian, Z. HBV inhibits LPS-induced NLRP3 inflammasome activation and IL-1β production via suppressing the NF-κB pathway and ROS production. J. Hepatol. 2017, 66, 693–702. [Google Scholar] [CrossRef]
  125. Petrasek, J.; Iracheta-Vellve, A.; Saha, B.; Satishchandran, A.; Kodys, K.; Fitzgerald, K.A.; Kurt-Jones, E.A.; Szabo, G. Metabolic danger signals, uric acid and ATP, mediate inflammatory cross-talk between hepatocytes and immune cells in alcoholic liver disease. J. Leukoc. Biol. 2015, 98, 249–256. [Google Scholar] [CrossRef]
  126. Ganz, M.; Bukong, T.N.; Csak, T.; Saha, B.; Park, J.K.; Ambade, A.; Kodys, K.; Szabo, G. Progression of non-alcoholic steatosis to steatohepatitis and fibrosis parallels cumulative accumulation of danger signals that promote inflammation and liver tumors in a high fat-cholesterol-sugar diet model in mice. J. Transl. Med. 2015, 13, 193. [Google Scholar] [CrossRef]
  127. Knorr, J.; Wree, A.; Tacke, F.; Feldstein, A.E. The NLRP3 Inflammasome in Alcoholic and Nonalcoholic Steatohepatitis. Semin. Liver Dis. 2020, 40, 298–306. [Google Scholar] [CrossRef]
  128. Petrasek, J.; Bala, S.; Csak, T.; Lippai, D.; Kodys, K.; Menashy, V.; Barrieau, M.; Min, S.Y.; Kurt-Jones, E.A.; Szabo, G. IL-1 receptor antagonist ameliorates inflammasome-dependent alcoholic steatohepatitis in mice. J. Clin. Investig. 2012, 122, 3476–3489. [Google Scholar] [CrossRef]
  129. Vandanmagsar, B.; Youm, Y.H.; Ravussin, A.; Galgani, J.E.; Stadler, K.; Mynatt, R.L.; Ravussin, E.; Stephens, J.M.; Dixit, V.D. The NLRP3 inflammasome instigates obesity-induced inflammation and insulin resistance. Nat. Med. 2011, 17, 179–188. [Google Scholar] [CrossRef]
  130. Martínez-Micaelo, N.; González-Abuín, N.; Pinent, M.; Ardévol, A.; Blay, M. Dietary fatty acid composition is sensed by the NLRP3 inflammasome: Omega-3 fatty acid (DHA) prevents NLRP3 activation in human macrophages. Food Funct. 2016, 7, 3480–3487. [Google Scholar] [CrossRef]
  131. Pan, J.; Ou, Z.; Cai, C.; Li, P.; Gong, J.; Ruan, X.Z.; He, K. Fatty acid activates NLRP3 inflammasomes in mouse Kupffer cells through mitochondrial DNA release. Cell. Immunol. 2018, 332, 111–120. [Google Scholar] [CrossRef] [PubMed]
  132. Jin, K.; Liu, Y.; Shi, Y.; Zhang, H.; Sun, Y.; Zhangyuan, G.; Wang, F.; Yu, W.; Wang, J.; Tao, X.; et al. PTPROt aggravates inflammation by enhancing NF-κB activation in liver macrophages during nonalcoholic steatohepatitis. Theranostics 2020, 10, 5290–5304, Erratum in Theranostics 2022, 12, 3604–3606. [Google Scholar] [CrossRef]
  133. Csak, T.; Ganz, M.; Pespisa, J.; Kodys, K.; Dolganiuc, A.; Szabo, G. Fatty acid and endotoxin activate inflammasomes in mouse hepatocytes that release danger signals to stimulate immune cells. Hepatology 2011, 54, 133–144. [Google Scholar] [CrossRef]
  134. Wree, A.; McGeough, M.D.; Peña, C.A.; Schlattjan, M.; Li, H.; Inzaugarat, M.E.; Messer, K.; Canbay, A.; Hoffman, H.M.; Feldstein, A.E. NLRP3 inflammasome activation is required for fibrosis development in NAFLD. J. Mol. Med. 2014, 92, 1069–1082. [Google Scholar] [CrossRef] [PubMed]
  135. Devisscher, L.; Scott, C.L.; Lefere, S.; Raevens, S.; Bogaerts, E.; Paridaens, A.; Verhelst, X.; Geerts, A.; Guilliams, M.; Van Vlierberghe, H. Non-alcoholic steatohepatitis induces transient changes within the liver macrophage pool. Cell. Immunol. 2017, 322, 74–83. [Google Scholar] [CrossRef]
  136. Daemen, S.; Gainullina, A.; Kalugotla, G.; He, L.; Chan, M.M.; Beals, J.W.; Liss, K.H.; Klein, S.; Feldstein, A.E.; Finck, B.N.; et al. Dynamic Shifts in the Composition of Resident and Recruited Macrophages Influence Tissue Remodeling in NASH. Cell Rep. 2021, 34, 108626, Erratum in Cell Rep. 2022, 41, 111660. [Google Scholar] [CrossRef] [PubMed]
  137. Lefere, S.; Degroote, H.; Van Vlierberghe, H.; Devisscher, L. Unveiling the depletion of Kupffer cells in experimental hepatocarcinogenesis through liver macrophage subtype-specific markers. J. Hepatol. 2019, 71, 631–633. [Google Scholar] [CrossRef]
  138. Tran, S.; Baba, I.; Poupel, L.; Dussaud, S.; Moreau, M.; Gélineau, A.; Marcelin, G.; Magréau-Davy, E.; Ouhachi, M.; Lesnik, P.; et al. Impaired Kupffer Cell Self-Renewal Alters the Liver Response to Lipid Overload during Non-alcoholic Steatohepatitis. Immunity 2020, 53, 627–640.e5. [Google Scholar] [CrossRef]
  139. Seidman, J.S.; Troutman, T.D.; Sakai, M.; Gola, A.; Spann, N.J.; Bennett, H.; Bruni, C.M.; Ouyang, Z.; Li, R.Z.; Sun, X.; et al. Niche-Specific Reprogramming of Epigenetic Landscapes Drives Myeloid Cell Diversity in Nonalcoholic Steatohepatitis. Immunity 2020, 52, 1057–1074.e7. [Google Scholar] [CrossRef]
  140. Bonnardel, J.; T’Jonck, W.; Gaublomme, D.; Browaeys, R.; Scott, C.L.; Martens, L.; Vanneste, B.; De Prijck, S.; Nedospasov, S.A.; Kremer, A.; et al. Stellate Cells, Hepatocytes, and Endothelial Cells Imprint the Kupffer Cell Identity on Monocytes Colonizing the Liver Macrophage Niche. Immunity 2019, 51, 638–654.e9. [Google Scholar] [CrossRef]
  141. Sakai, M.; Troutman, T.D.; Seidman, J.S.; Ouyang, Z.; Spann, N.J.; Abe, Y.; Ego, K.M.; Bruni, C.M.; Deng, Z.; Schlachetzki, J.C.M.; et al. Liver-Derived Signals Sequentially Reprogram Myeloid Enhancers to Initiate and Maintain Kupffer Cell Identity. Immunity 2019, 51, 655–670.e8. [Google Scholar] [CrossRef] [PubMed]
  142. Dixon, L.J.; Barnes, M.; Tang, H.; Pritchard, M.T.; Nagy, L.E. Kupffer cells in the liver. Compr. Physiol. 2013, 3, 785–797. [Google Scholar] [CrossRef] [PubMed]
  143. Duarte, N.; Coelho, I.C.; Patarrão, R.S.; Almeida, J.I.; Penha-Gonçalves, C.; Macedo, M.P. How Inflammation Impinges on NAFLD: A Role for Kupffer Cells. Biomed. Res. Int. 2015, 2015, 984578. [Google Scholar] [CrossRef] [PubMed]
  144. Liu, W.; Baker, R.D.; Bhatia, T.; Zhu, L.; Baker, S.S. Pathogenesis of nonalcoholic steatohepatitis. Cell. Mol. Life Sci. 2016, 73, 1969–1987. [Google Scholar] [CrossRef]
  145. Schuppan, D.; Surabattula, R.; Wang, X.Y. Determinants of fibrosis progression and regression in NASH. J. Hepatol. 2018, 68, 238–250. [Google Scholar] [CrossRef]
  146. Wan, J.; Benkdane, M.; Teixeira-Clerc, F.; Bonnafous, S.; Louvet, A.; Lafdil, F.; Pecker, F.; Tran, A.; Gual, P.; Mallat, A.; et al. M2 Kupffer cells promote M1 Kupffer cell apoptosis: A protective mechanism against alcoholic and nonalcoholic fatty liver disease. Hepatology 2014, 59, 130–142. [Google Scholar] [CrossRef]
  147. Han, Y.H.; Kim, H.J.; Na, H.; Nam, M.W.; Kim, J.Y.; Kim, J.S.; Koo, S.H.; Lee, M.O. RORα Induces KLF4-Mediated M2 Polarization in the Liver Macrophages that Protect against Nonalcoholic Steatohepatitis. Cell Rep. 2017, 20, 124–135. [Google Scholar] [CrossRef]
  148. Park, J.W.; Jeong, G.; Kim, S.J.; Kim, M.K.; Park, S.M. Predictors reflecting the pathological severity of non-alcoholic fatty liver disease: Comprehensive study of clinical and immunohistochemical findings in younger Asian patients. J. Gastroenterol. Hepatol. 2007, 22, 491–497. [Google Scholar] [CrossRef]
  149. Lotowska, J.M.; Sobaniec-Lotowska, M.E.; Lebensztejn, D.M. The role of Kupffer cells in the morphogenesis of nonalcoholic steatohepatitis—Ultrastructural findings. The first report in pediatric patients. Scand. J. Gastroenterol. 2013, 48, 352–357. [Google Scholar] [CrossRef]
  150. Krenkel, O.; Tacke, F. Macrophages in Nonalcoholic Fatty Liver Disease: A Role Model of Pathogenic Immunometabolism. Semin. Liver Dis. 2017, 37, 189–197. [Google Scholar] [CrossRef]
  151. Friedman, S.L.; Ratziu, V.; Harrison, S.A.; Abdelmalek, M.F.; Aithal, G.P.; Caballeria, J.; Francque, S.; Farrell, G.; Kowdley, K.V.; Craxi, A.; et al. A randomized, placebo-controlled trial of cenicriviroc for treatment of nonalcoholic steatohepatitis with fibrosis. Hepatology 2018, 67, 1754–1767. [Google Scholar] [CrossRef] [PubMed]
  152. Tacke, F. Targeting hepatic macrophages to treat liver diseases. J. Hepatol. 2017, 66, 1300–1312. [Google Scholar] [CrossRef]
  153. Tawfeek, G.A.; Kasem, H.A. Curcumin preconditioned mesenchymal stem cells derived exosomes transplantation ameliorate and protect against non- alcoholic steatohepatitis by regulation the expression of key genes of inflammation and oxidative stress. Transpl. Immunol. 2023, 78, 101837. [Google Scholar] [CrossRef] [PubMed]
  154. Yang, F.; Wu, Y.; Chen, Y.; Xi, J.; Chu, Y.; Jin, J.; Yan, Y. Human umbilical cord mesenchymal stem cell-derived exosomes ameliorate liver steatosis by promoting fatty acid oxidation and reducing fatty acid synthesis. JHEP Rep. 2023, 5, 100746. [Google Scholar] [CrossRef] [PubMed]
  155. Lee, C.W.; Hsiao, W.T.; Lee, O.K. Mesenchymal stromal cell-based therapies reduce obesity and metabolic syndromes induced by a high-fat diet. Transl. Res. 2017, 182, 61–74.e8. [Google Scholar] [CrossRef]
  156. Kholodenko, I.V.; Kholodenko, R.V.; Yarygin, K.N. The Crosstalk between Mesenchymal Stromal/Stem Cells and Hepatocytes in Homeostasis and under Stress. Int. J. Mol. Sci. 2023, 24, 15212. [Google Scholar] [CrossRef]
  157. Ezquer, M.; Ezquer, F.; Ricca, M.; Allers, C.; Conget, P. Intravenous administration of multipotent stromal cells prevents the onset of non-alcoholic steatohepatitis in obese mice with metabolic syndrome. J. Hepatol. 2011, 55, 1112–1120. [Google Scholar] [CrossRef]
  158. Shi, Y.; Yang, X.; Wang, S.; Wu, Y.; Zheng, L.; Tang, Y.; Gao, Y.; Niu, J. Human umbilical cord mesenchymal stromal cell-derived exosomes protect against MCD-induced NASH in a mouse model. Stem Cell Res. Ther. 2022, 13, 517. [Google Scholar] [CrossRef]
  159. Muto, H.; Ito, T.; Tanaka, T.; Yokoyama, S.; Yamamoto, K.; Imai, N.; Ishizu, Y.; Maeda, K.; Honda, T.; Ishikawa, T.; et al. Conditioned medium from stem cells derived from human exfoliated deciduous teeth ameliorates NASH via the Gut-Liver axis. Sci. Rep. 2021, 11, 18778. [Google Scholar] [CrossRef]
  160. Watanabe, T.; Tsuchiya, A.; Takeuchi, S.; Nojiri, S.; Yoshida, T.; Ogawa, M.; Itoh, M.; Takamura, M.; Suganami, T.; Ogawa, Y.; et al. Development of a non-alcoholic steatohepatitis model with rapid accumulation of fibrosis, and its treatment using mesenchymal stem cells and their small extracellular vesicles. Regen. Ther. 2020, 14, 252–261. [Google Scholar] [CrossRef]
  161. Grunhut, J.; Wang, W.; Aykut, B.; Gakhal, I.; Torres−Hernandez, A.; Miller, G. Macrophages in Nonalcoholic Steatohepatitis: Friend or Foe? Eur. Med. J. Hepatol. 2018, 6, 100–109. [Google Scholar] [CrossRef] [PubMed]
  162. Heymann, F.; Trautwein, C.; Tacke, F. Monocytes and macrophages as cellular targets in liver fibrosis. Inflamm. Allergy Drug Targets 2009, 8, 307–318. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, B.; Zhang, B.; Lai, R.C.; Sim, W.K.; Lam, K.P.; Lim, S.K. MSC-sEV Treatment Polarizes Pro-Fibrotic M2 Macrophages without Exacerbating Liver Fibrosis in NASH. Int. J. Mol. Sci. 2023, 24, 8092. [Google Scholar] [CrossRef] [PubMed]
  164. Ohara, M.; Ohnishi, S.; Hosono, H.; Yamamoto, K.; Yuyama, K.; Nakamura, H.; Fu, Q.; Maehara, O.; Suda, G.; Sakamoto, N. Extracellular Vesicles from Amnion-Derived Mesenchymal Stem Cells Ameliorate Hepatic Inflammation and Fibrosis in Rats. Stem Cells Int. 2018, 2018, 3212643. [Google Scholar] [CrossRef]
  165. Liang, X.; Li, T.; Zhou, Q.; Pi, S.; Li, Y.; Chen, X.; Weng, Z.; Li, H.; Zhao, Y.; Wang, H.; et al. Mesenchymal stem cells attenuate sepsis-induced liver injury via inhibiting M1 polarization of Kupffer cells. Mol. Cell. Biochem. 2019, 452, 187–197. [Google Scholar] [CrossRef]
  166. Wang, J.; Liu, Y.; Ding, H.; Shi, X.; Ren, H. Mesenchymal stem cell-secreted prostaglandin E2 ameliorates acute liver failure via attenuation of cell death and regulation of macrophage polarization. Stem Cell Res. Ther. 2021, 12, 15. [Google Scholar] [CrossRef]
  167. Tian, Y.; Wang, J.; Wang, W.; Ding, Y.; Sun, Z.; Zhang, Q.; Wang, Y.; Xie, H.; Yan, S.; Zheng, S. Mesenchymal stem cells improve mouse non-heart-beating liver graft survival by inhibiting Kupffer cell apoptosis via TLR4-ERK1/2-Fas/FasL-caspase3 pathway regulation. Stem Cell Res. Ther. 2016, 7, 157. [Google Scholar] [CrossRef]
  168. Liu, F.; Qiu, H.; Xue, M.; Zhang, S.; Zhang, X.; Xu, J.; Chen, J.; Yang, Y.; Xie, J. MSC-secreted TGF-β regulates lipopolysaccharide-stimulated macrophage M2-like polarization via the Akt/FoxO1 pathway. Stem Cell Res. Ther. 2019, 10, 345. [Google Scholar] [CrossRef]
  169. François, M.; Romieu-Mourez, R.; Li, M.; Galipeau, J. Human MSC suppression correlates with cytokine induction of indoleamine 2,3-dioxygenase and bystander M2 macrophage differentiation. Mol. Ther. 2012, 20, 187–195. [Google Scholar] [CrossRef]
  170. Jackson, M.V.; Krasnodembskaya, A.D. Analysis of Mitochondrial Transfer in Direct Co-cultures of Human Monocyte-derived Macrophages (MDM) and Mesenchymal Stem Cells (MSC). Bio-Protocol 2017, 7, e2255. [Google Scholar] [CrossRef]
  171. Morrison, T.J.; Jackson, M.V.; Cunningham, E.K.; Kissenpfennig, A.; McAuley, D.F.; O’Kane, C.M.; Krasnodembskaya, A.D. Mesenchymal Stromal Cells Modulate Macrophages in Clinically Relevant Lung Injury Models by Extracellular Vesicle Mitochondrial Transfer. Am. J. Respir. Crit. Care Med. 2017, 196, 1275–1286. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, Y.; Zhang, X.; Zhang, H.; Song, P.; Pan, W.; Xu, P.; Wang, G.; Hu, P.; Wang, Z.; Huang, K.; et al. Mesenchymal Stem Cells Derived Extracellular Vesicles Alleviate Traumatic Hemorrhagic Shock Induced Hepatic Injury via IL-10/PTPN22-Mediated M2 Kupffer Cell Polarization. Front. Immunol. 2022, 12, 811164. [Google Scholar] [CrossRef] [PubMed]
  173. Chang, H.H.; Miaw, S.C.; Tseng, W.; Sun, Y.W.; Liu, C.C.; Tsao, H.W.; Ho, I.C. PTPN22 modulates macrophage polarization and susceptibility to dextran sulfate sodium-induced colitis. J. Immunol. 2013, 191, 2134–2143. [Google Scholar] [CrossRef] [PubMed]
  174. Kholodenko, I.V.; Kholodenko, R.V.; Majouga, A.G.; Yarygin, K.N. Apoptotic MSCs and MSC-Derived Apoptotic Bodies as New Therapeutic Tools. Curr. Issues Mol. Biol. 2022, 44, 5153–5172. [Google Scholar] [CrossRef]
  175. de Witte, S.F.H.; Luk, F.; Sierra Parraga, J.M.; Gargesha, M.; Merino, A.; Korevaar, S.S.; Shankar, A.S.; O’Flynn, L.; Elliman, S.J.; Roy, D.; et al. Immunomodulation By Therapeutic Mesenchymal Stromal Cells (MSC) Is Triggered Through Phagocytosis of MSC By Monocytic Cells. Stem Cells 2018, 36, 602–615. [Google Scholar] [CrossRef] [PubMed]
  176. Sheng, M.; Lin, Y.; Xu, D.; Tian, Y.; Zhan, Y.; Li, C.; Farmer, D.G.; Kupiec-Weglinski, J.W.; Ke, B. CD47-Mediated Hedgehog/SMO/GLI1 Signaling Promotes Mesenchymal Stem Cell Immunomodulation in Mouse Liver Inflammation. Hepatology 2021, 74, 1560–1577. [Google Scholar] [CrossRef]
  177. Francque, S.; Moreno, C.; Saliba, F.; Vargas, V.; Ibañez, L.; Soriano, G.; Genov, J.; Katzarov, K.; Janczewska, E.; Iliescu, E.L.; et al. TOP-077. Safety, tolerability, and preliminary efficacy of ascending doses of Human Allogeneic Liver-derived Progenitor Cells (HepaStem®) in patients with cirrhotic and pre-cirrhotic non-alcoholic steatohepatitis (NASH). J. Hepatol. 2023, 78 (Suppl. S1), S807–S808. [Google Scholar] [CrossRef]
  178. Najimi, M.; Michel, S.; Binda, M.M.; Gellynck, K.; Belmonte, N.; Mazza, G.; Gordillo, N.; Vainilovich, Y.; Sokal, E. Human Allogeneic Liver-Derived Progenitor Cells Significantly Improve NAFLD Activity Score and Fibrosis in Late-Stage NASH Animal Model. Cells 2022, 11, 2854. [Google Scholar] [CrossRef]
  179. Sakai, Y.; Fukunishi, S.; Takamura, M.; Kawaguchi, K.; Inoue, O.; Usui, S.; Takashima, S.; Seki, A.; Asai, A.; Tsuchimoto, Y.; et al. Clinical trial of autologous adipose tissue-derived regenerative (stem) cells therapy for exploration of its safety and efficacy. Regen. Ther. 2021, 18, 97–101. [Google Scholar] [CrossRef]
Figure 1. NASH pathogenesis. Excessive consumption of high-fat foods and an unhealthy lifestyle leads to obesity. Obesity is one of the causes of the non-alcoholic fatty liver disease (NAFLD). Excessive accumulation of fats in the liver leads to lipotoxicity and death of hepatocytes. Dead and/or damaged hepatocytes release large amounts of DAMPs, which cause severe inflammation, thereby exacerbating hepatocyte damage.
Figure 1. NASH pathogenesis. Excessive consumption of high-fat foods and an unhealthy lifestyle leads to obesity. Obesity is one of the causes of the non-alcoholic fatty liver disease (NAFLD). Excessive accumulation of fats in the liver leads to lipotoxicity and death of hepatocytes. Dead and/or damaged hepatocytes release large amounts of DAMPs, which cause severe inflammation, thereby exacerbating hepatocyte damage.
Biomedicines 11 03056 g001
Figure 2. Activation of Kupffer cells in NASH liver. DAMPs released by dead hepatocytes activate Kupffer cells or recruit macrophages, which begin to produce a wide variety of the pro-inflammatory cytokines/chemokines and inflammosomes. Chemokines stimulate the recruitment of the innate and adaptive immune cells to the liver, leading to increased inflammation and further liver damage.
Figure 2. Activation of Kupffer cells in NASH liver. DAMPs released by dead hepatocytes activate Kupffer cells or recruit macrophages, which begin to produce a wide variety of the pro-inflammatory cytokines/chemokines and inflammosomes. Chemokines stimulate the recruitment of the innate and adaptive immune cells to the liver, leading to increased inflammation and further liver damage.
Biomedicines 11 03056 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kholodenko, I.V.; Yarygin, K.N. Hepatic Macrophages as Targets for the MSC-Based Cell Therapy in Non-Alcoholic Steatohepatitis. Biomedicines 2023, 11, 3056. https://doi.org/10.3390/biomedicines11113056

AMA Style

Kholodenko IV, Yarygin KN. Hepatic Macrophages as Targets for the MSC-Based Cell Therapy in Non-Alcoholic Steatohepatitis. Biomedicines. 2023; 11(11):3056. https://doi.org/10.3390/biomedicines11113056

Chicago/Turabian Style

Kholodenko, Irina V., and Konstantin N. Yarygin. 2023. "Hepatic Macrophages as Targets for the MSC-Based Cell Therapy in Non-Alcoholic Steatohepatitis" Biomedicines 11, no. 11: 3056. https://doi.org/10.3390/biomedicines11113056

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