ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Pathogenesis of hepatic fibrosis

Pathogenesis of hepatic fibrosis
Literature review current through: Jan 2024.
This topic last updated: Jan 02, 2024.

INTRODUCTION — Fibrosis is a wound healing response in which damaged regions are encapsulated by an extracellular matrix or scar. It develops in almost all patients with chronic liver injury at variable rates depending in part upon the cause of liver disease and host factors [1-4]. In contrast, for unclear reasons, patients with self-limited injury (such as fulminant hepatitis) do not develop scarring despite an abundance of fibrogenic stimuli, unless they go on to develop chronic injury.

The composition of the hepatic scar is similar regardless of the cause of injury. Furthermore, hepatic fibrosis represents a paradigm for wound healing in other tissues, including skin, lung, and kidney, since it involves many of the same cell types and mediators.

Fibrosis occurs earliest in regions where injury is most severe, particularly in chronic inflammatory liver disease due to alcohol or viral infection. As an example, pericentral injury is a hallmark of alcoholic hepatitis; the development of pericentral fibrosis (also known as sclerosing hyaline necrosis or perivenular fibrosis) is an early marker of likely progression to panlobular cirrhosis [5].

The development of fibrosis usually requires several months to years of ongoing injury. Three exceptions in adults are veno-occlusive disease, mechanical biliary obstruction, and fibrosing cholestatic hepatitis associated with viral hepatitis in which (for unclear reasons) fibrosis can progress more rapidly.

While fibrosis is reversible in its initial stages, progressive fibrosis can lead to cirrhosis. The exact point when fibrosis becomes irreversible is incompletely understood. However, increasing evidence suggests that even early stages of cirrhosis may be reversible [6-11]. Furthermore, an understanding of the molecular mechanisms involved in fibrogenesis has a number of clinical implications, including the development of interventions designed to impede or reverse hepatic fibrosis, some of which are already available [12,13]. Despite significant advances in understanding hepatic fibrosis and defining targets of therapy, there are no antifibrotic drugs yet approved for clinical use in patients with advanced liver disease.

This topic review will discuss the mechanisms underlying hepatic fibrosis.

EXTRACELLULAR MATRIX COMPOSITION OF THE NORMAL AND FIBROTIC LIVER — Extracellular matrix (ECM) refers to a group of macromolecules that comprise the scaffolding of normal and fibrotic liver. These include collagens, noncollagen glycoproteins, matrix-bound growth factors, glycosaminoglycans, proteoglycans, and matricellular proteins. A great deal of progress has been made in identifying new members of these families and in understanding how these molecules interact [14-16]. In addition to providing the scaffolding of the liver, matrix molecules are now recognized to have a variety of other functions. As an example, some serve as transmembrane transducers of extracellular signals.

Marked heterogeneity exists within different tissue regions in matrix composition with respect to the variety of isoforms within each class of molecules, their stoichiometry, and their intermolecular interactions. Furthermore, hybrid molecules have been identified that contain both collagenous and proteoglycan domains.

In the normal liver, collagen types I, III, V, and XI (sometimes referred to as "fibril-forming" collagens) are principally found in the capsule, around large vessels, and in the portal triad, while only scattered fibrils containing types I and III collagen can be found in the subendothelial space. Smaller amounts of other collagens including types IV, VI, XIV (previously called "undulin"), and XVIII can also be found.

Also present are glycoproteins and matricellular proteins including subendothelial deposits of fibronectin, laminin, tenascin, SPARC, and von Willebrand factor. The proteoglycans consist primarily of heparan sulfate proteoglycans such as perlecan, as well as small amounts of decorin, biglycan, fibromodulin, aggrecan, glypican, syndecan, and lumican [14].

The appearance of elastin may mark the conversion of reversible to irreversible fibrosis, and its degradation by macrophages may be an important determinant of elastic accumulation [17-21].

As the liver becomes fibrotic, significant changes occur in the ECM quantitatively and qualitatively. The total collagen content increases 3- to 10-fold [22]. There is also a marked increase in the ECM to levels and composition typically seen in wound healing. These include an increase in fibril-forming collagens (ie, types I, III, and IV), some non-fibril forming collagens (types IV and VI), a number of glycoproteins (cellular fibronectin, laminin, SPARC, osteonectin, tenascin, and von Willebrand factor), proteoglycans, and glycosaminoglycans (perlecan, decorin, aggrecan, lumican, and fibromodulin) [23]. Particularly notable is a shift from heparan sulfate-containing proteoglycans to those containing chondroitin and dermatan sulfates. These processes represent a change in the type of ECM in subendothelial space from the normal low-density basement membrane-like matrix to the interstitial type [23].

The replacement of the low-density matrix with the interstitial type has consequences on the function of hepatocytes, hepatic stellate cells, and endothelial cells, in part explaining the synthetic and metabolic dysfunction observed in patients with advanced fibrosis. The high-density matrix also activates hepatic stellate cells leading to the loss of hepatocyte microvilli and disappearance of endothelial fenestrations, which impairs transport of solutes from the sinusoid to the hepatocytes, thereby further contributing to the hepatocyte dysfunction [24]. At the same time, altered function of liver sinusoidal endothelial cells can promote injury and impede regeneration [25]

The liver responds to injury with angiogenic stimulation, with evidence of new blood vessel formation, sinusoidal remodeling, and pericyte (ie, stellate cell) amplification [26,27]. Thus, angiogenic mediators are involved, including platelet-derived growth factor, vascular endothelial growth factor (VEGF) and their cognate receptors, as well as vasoactive mediators that include nitric oxide and carbon monoxide. Increased VEGF concentrations may be particularly important in progression of fibrosis in smokers who have hepatitis C [28].

Progressive accumulation of ECM composition provokes positive feedback pathways that further amplify fibrosis.

First, changes in membrane receptors, in particular integrins, sense altered matrix signals that provoke stellate cell activation and migration through focal adhesion disassembly [29-31]. Matrix-provoked signals also engage membrane-bound GTP-binding proteins, in particular Rho [32] and Rac [33].

Second, activation of cellular matrix metalloproteases leads to release of fibrogenic and proliferative growth factors from matrix-bound reservoirs in the extracellular space [14,34,35].

Third, the enhanced density of ECM leads to increasing matrix stiffness, which may stimulate stellate cell activation through integrin signaling [36]. These experimental findings explain the increasing use of FibroScan [37] and magnetic resonance elastography [38], two clinical techniques which noninvasively assess hepatic stiffness as a reflection of ECM content. (See "Noninvasive assessment of hepatic fibrosis: Overview of serologic tests and imaging examinations".)

BIOLOGIC ACTIVITY OF EXTRACELLULAR MATRIX — Matrix alterations observed during fibrogenesis alter cellular behavior by processes involving cell membrane receptors. One of the best characterized are integrins, which are a large family of homologous membrane linker proteins that control several cellular functions including gene expression, growth, and differentiation. A variety of other cytokines and adhesion proteins involved in hepatic fibrogenesis have also been described.

Integrins — Integrins are composed of alpha and beta subunits whose ligands are matrix molecules rather than cytokines [39]. In particular, integrin ligands contain an Arg-Gly-Asp (RGD) tripeptide sequence. Several integrins and their downstream effectors have been identified in stellate cells, including alpha-1-beta-1, alpha-2-beta-1, alpha-5-beta-1, and alpha-6-beta-4. The common presence of RGD with many integrin ligands has raised the possibility of using competitive RGD antagonists to block integrin-mediated pathways in fibrogenesis.

Additional studies have identified the integrin subunits alpha-1, alpha-V, beta-1 and beta-6 as potentially important therapeutic targets, not only in liver diseases [40] but also in fibrosis of other organs [41-43].

Integrin signaling across the plasma membrane permits communication between the extracellular matrix (ECM) and cytoskeleton. Signaling occurs in conjunction with the phosphorylation of several of intracellular substrates, which is typical of most transmembrane receptors. However, in addition to the "outside-in" signaling pathway, integrins can also signal in the opposite direction (ie, from the inside to the outside of the cell), thereby mediating cytoskeletal changes that can lead to altered conformation of ECM molecules such as fibronectin.

Several integrin and non-integrin receptors have been described in situ on hepatocytes and non-parenchymal cells [29,30,44-47]. Upregulation of alpha-6-beta-1 and alpha-2-beta-1 receptors [5], both of which bind laminin, has been reported in experimental fibrosis. Studies have also defined the integrin phenotypes of isolated cell types from liver. In particular, stellate cells express integrin receptors for collagen and laminin [29,44,45,48], which may contribute to their activation and proliferation in response to deposition of these matrix components during injury.

Integrins are an attractive target for antifibrotic therapies because several integrin heterodimers can activate the fibrogenic cytokine TGF-beta at the cell surface, and therefore, integrin antagonists may interrupt this effect [41] (see 'Soluble growth factors' below). Clinical trials exploring the role of integrins are underway in patients with fibrosis of the liver and other tissues [49].

Other adhesion proteins and cell matrix receptors — A growing number of adhesion proteins and cell matrix receptors other than integrins have been characterized, including cadherins and selectins, which mediate interactions between inflammatory cells and the endothelial wall [50-52]. As an example, upregulation of a tyrosine kinase receptor, discoidin domain receptor 2 (DDR2), has been observed during stellate cell activation leading to enhanced matrix metalloproteinase expression and cell growth [53]. DDR2 is the only tyrosine kinase receptor whose ligand is an ECM molecule rather than a peptide ligand. Its upregulation may be a critical step toward perpetuating liver fibrosis. Cell culture studies suggest this might be a potential therapeutic target [54].

Soluble growth factors — The ECM can also affect cell function indirectly by the release of soluble growth factors (cytokines). The soluble growth factors are controlled by local metalloproteinases (a family of zinc-dependent enzymes) [55]. The cytokines include platelet-derived growth factor, HGF, connective tissue growth factor, tumor necrosis factor alpha, basic fibroblast growth factor, and vascular endothelial cell growth factor [14]. Controlled release of these cytokines from the ECM is a key mechanism for regulating cytokine activity since it provides a local, accessible source of cytokines that can be regulated tightly through the actions of proteases and their inhibitors. In addition, ECM can regulate the activity of proteases through specific binding to collagens or fibronectins [14].

Transforming growth factor (TGF)-beta-1, derived from both paracrine and autocrine sources, remains the classic fibrogenic cytokine [56,57]. Signals downstream of TGF-beta converge upon Smad proteins, which fine-tune and enhance the effects of TGF-beta during stellate cell activation; Smads 2 and 3 are stimulatory, whereas Smad7 is inhibitory [56,58,59] and is antagonized by Id1 [60]. TGF-beta-1 also stimulates collagen transcription in stellate cells through a hydrogen peroxide- and C/EBP-beta-dependent mechanism [61]. The response of Smads in stellate cells evolves as injury becomes chronic, further enhancing fibrogenesis [58,62]. Studies have identified an upstream signaling intermediate, IQGAP1, that attenuates TGF-beta signaling [63].

CELLULAR SOURCES OF ECM IN NORMAL AND FIBROTIC LIVER — A major advance in the understanding of hepatic fibrosis has been the identification of the cellular sources of extracellular matrix (ECM). The hepatic stellate cell (previously called the lipocyte, Ito, fat-storing, or perisinusoidal cell) is the primary source of ECM in normal and fibrotic liver. In addition, related mesenchymal cell types from a variety of sources may have measurable contributions to total matrix accumulation, including classical portal fibroblasts [64-66] (especially in biliary fibrosis), bone marrow-derived cells [67], as well as fibroblasts derived from epithelial-mesenchymal transition (EMT) [68]. EMT is a well-characterized response of the kidney [68] to injury, but its role in liver injury has been less convincing.

Hepatic stellate cells, located in subendothelial space of Disse between hepatocytes and sinusoidal endothelial cells, represent one-third of the non-parenchymal population or approximately 15 percent of the total number of resident cells in normal liver (figure 1) [69]. In normal liver, they are the principal storage site for retinoids (vitamin A metabolites), which accounts for 40 to 70 percent of retinoids in the body. Most of the retinoids are in the form of retinyl esters and are confined to cytoplasmic droplets. Stellate cells actually comprise a somewhat heterogeneous group of cells that are functionally and anatomically similar but differ in their expression of cytoskeletal filaments, their retinoid content, and in their potential for activation [70-72]. A series of studies have used single-cell sequencing to fully characterize the heterogeneity of hepatic stellate cells in mouse models and human liver [73-78]. These studies underscore the heterogeneity of stellate cells, but also reinforce the conclusion that activated stellate cells/myofibroblasts are the primary fibrogenic cell in injured liver. Stellate cells with fibrogenic potential have been identified in locations other than the liver, such as the pancreas [79]. Furthermore, remarkable plasticity of stellate cell phenotype has been documented in vivo and in culture, precluding a strict definition based only on cytoskeletal phenotype [47,80]. Indeed, single cell analyses indicate there are several subtypes, whose distinct functional roles are not yet fully clarified. Stellate cells with fibrogenic potential are not confined to the liver and have been identified in other organs such as the pancreas, where they contribute to primary or metastatic cancers as cancer-associated fibroblasts [81-83] or to desmoplasia in chronic pancreatitis [84].

Animal and human studies have defined a process of changes within stellate cells that collectively are termed "activation." During activation, stellate cells undergo a transition from a quiescent vitamin A-rich cell into proliferative, fibrogenic, and contractile myofibroblasts (figure 2). This change is characterized morphologically by enlargement of rough endoplasmic reticulum, diminution of vitamin A droplets, a ruffled nuclear membrane, appearance of contractile filaments, and proliferation. Cells with features of both quiescent and activated cells are sometimes referred to as "transitional" cells. Proliferation of stellate cells occurs predominantly in regions of greatest injury. Their function has been characterized in a variety of human diseases including alcohol-associated liver disease, viral hepatitis, hepatocellular carcinoma, vascular diseases, hematologic malignancy, biliary disease, mucopolysaccharidosis, acetaminophen overdose, leishmaniasis, allograft rejection, and in substance use disorder [85-89].

Sinusoidal endothelial cells are also an important contributor to early fibrosis. Similar to stellate cells, there is considerable heterogeneity of this cell type in normal and fibrotic liver [72]. Endothelial cells from normal liver produce types III and IV collagen, laminin, syndecan, and fibronectin [90-93]. Increased expression of cellular isoforms of fibronectin by these cells is a key early event following acute liver injury because their appearance creates a microenvironment that activates stellate cells.

Other cell types may also contribute to fibrosis. CD8+ lymphocytes are potentially profibrogenic cells based upon their ability to induce fibrogenesis after adoptive transfer from animals with liver injury [94]. However, in hepatitis C virus infection, direct acting antiviral treatment may not resolve the increase in this cell type [95]. Cross-talk between sinusoidal endothelial cells and stellate cells or hepatocytes regulate both fibrosis and regeneration [96,97]. Cross talk between macrophages and stellate cells is a key driver of fibrogenesis as well [73].

Remarkably, very few studies have defined the cellular or matrix composition of congenital hepatic fibrosis, an entity whose pathogenesis is unclear. Current theories suggest that as in adults, congenital fibrosis represents a final common pathway of fetal hepatic injury, whether from biliary malformations, viral infections (especially cytomegalovirus) or other insult, with the stellate cell playing a significant role [98].

Fibrogenic cells in the liver derive not only from resident stellate cells, but also from portal fibroblasts [64,99,100], and potentially from circulating fibrocytes [101], bone marrow [67], and epithelial-mesenchymal cell transition [102,103]. Portal fibroblasts may be particularly important in cholestatic liver diseases and ischemia [100], in which paracrine interactions between cholangiocytes and fibroblasts involve both chemokines [104] and extracellular nucleotides [105]. Although controversial, progressive recruitment of bone marrow derived cells may occur over time, but do not represent a major fraction of the total fibrogenic population in chronic injury. Studies using genetic tracing and single-cell sequencing techniques emphasize that, while other sources including bone marrow are possible, the vast bulk of fibrotic gene expression occurs in myofibroblasts derived from activated stellate cells [73-76,106].

The development of technologies to sequence the full transcriptome of individual cells has transformed our understanding of hepatic fibrogenesis by defining an unprecedented level of heterogeneity among stellate cells and other fibrogenic cells in human and rodent livers [74,75,107,108]. These studies may allow us to target specific subsets of cells that are especially fibrogenic and pro-inflammatory. For example, one study used chimeric-antigen receptor (CAR) T cells engineered to deplete only senescent stellate cells in a fibrosis model in mice, which markedly reduced fibrosis and improved liver function [109].

DEGRADATION OF EXTRACELLULAR MATRIX — Fibrosis reflects a balance between matrix production and degradation. The degradation of extracellular matrix is a key event in hepatic fibrosis. Early disruption of the normal hepatic matrix by matrix proteases hastens its replacement by scar matrix, which has deleterious effects on cell function. As a result, such degradation has been referred to as being "pathologic." On the other hand, resorption of excess matrix in patients with chronic liver disease provides the opportunity to reverse hepatic dysfunction and portal hypertension [110].

An understanding of mechanisms involved in matrix remodeling has evolved significantly in the past several years. A critical element in matrix remodeling is a family of matrix metalloproteinases (MMPs; also known as matrixins). These are calcium-dependent enzymes that specifically degrade collagens and noncollagenous substrates [111,112]. As a general rule, the MMPs fall into five categories based upon their substrate specificity:

Interstitial collagenases (MMP-1, -8, -13), which degrade interstitial collagens

Gelatinases (MMP-2, -9 and fibroblast activation protein), which generally degrade basement membrane collagens and denatured interstitial collagen

Stromelysins (MMP-3, -7, -10, 11), which have a broad substrate range

Membrane type (MMP-14, 15, -16, -17, -24, -25), which are primarily interstitial collagenases

Metalloelastase (MMP-12), which degrades elastin [113]

Metalloproteinases are regulated at many levels, permitting their activity to be restricted to discrete regions within the pericellular environment. Inactive metalloproteinases can be activated through proteolytic cleavage by either membrane-type MMP-1 (MT1-MMP) or plasmin, and inhibited by binding to specific inhibitors known as tissue inhibitors of metalloproteinases (TIMPs). As an example, MT1-MMP and TIMP-2 form a ternary complex with MMP-2, possibly including avb3 integrin, which is essential for optimal MMP-2 activity [111]. Plasmin activity is controlled by its activating enzyme uroplasminogen activator (uPA) and a specific inhibitor, plasminogen activator inhibitor 1 (PAI-1), and can be stimulated by active transforming growth factor beta-1 (TGF-beta-1).

Thus, collagenase activity reflects the relative amounts of activated metalloproteinases and their inhibitors, especially TIMPs. Other protease inhibitors (such as a2-macroglobulin) may also affect net degradative activity.

Fibrosis regression is increasingly observed as treatments for chronic liver disease have improved, especially for hepatitis B virus and hepatitis C virus infection. Thus, interest is focused on mediators of fibrosis regression. In animal models, macrophage subsets have emerged as an important determinant, especially those that have low-cell surface expression of Ly6C [114], but their corresponding human counterparts are not yet clarified. In mouse studies, triggering receptors expressed on myeloid cells 2 (Trem2) is a cell surface molecule that marks macrophages that are more pro-inflammatory and pro-fibrogenic [73].

Pathologic matrix degradation — "Pathologic" matrix degradation refers to the early disruption of the matrix of the normal space between hepatocytes and endothelial cells. Degradation occurs through the actions of at least four enzymes:

Matrix metalloproteinase 2 (MMP-2) (also called "gelatinase A" or "72 kDA type IV collagenase") and MMP-9 ("gelatinase B" or "92 kDa type IV collagenase"), which degrade type IV collagen

Membrane-type metalloproteinase-1 or -2, which activate latent MMP-2

Stromelysin-1, which degrades proteoglycans and glycoproteins and also activates latent collagenases

Stellate cells are the principal source of MMP-2 [115,116] and stromelysin [117]. Activation of latent MMP-2 may require interaction with hepatocytes [118,119]. Markedly increased expression of MMP-2 is characteristic of cirrhosis [120]. MMP-9 is secreted locally by Kupffer cells and macrophages [111]. Disruption of the normal liver matrix is also a requirement for tumor invasion and desmoplasia [121].

As noted above, a paradigm is emerging in which inflammatory cell subsets are the primary source of degradative enzymes that can degrade scar. In particular, macrophage subsets, including one called "Ly6C-lo," are increased when liver fibrosis regression is maximal [122]. Additionally, dendritic cells in liver also harbor proteolytic activity [123]. An increasing focus on inflammatory and immunoregulatory cells in the liver is likely to yield important advances in understanding matrix degradation during fibrosis regression, with clear therapeutic implications [124-126].

Failure to degrade the increased interstitial or scar matrix is a major determinant of progressive fibrosis. MMP-1 is the main protease that can degrade type I collagen, the principal collagen in fibrotic liver. However, sources of this enzyme are not as clearly established as for the type IV collagenases. Stellate cells express MMP-1 mRNA, but little enzyme can be detected [127]. More importantly, progressive fibrosis is associated with marked increases in TIMP-1 [128,129] and TIMP-2 [130], leading to a net decrease in protease activity and therefore more unopposed matrix accumulation. Stellate cells are the major source of these inhibitors [131]. Sustained TIMP-1 expression is emerging as a key reason for progressive fibrosis, and its diminution is an important prerequisite to allow for reversal of fibrosis (see below). Unique mechanisms of TIMP-1 regulation in stellate cells [132] offer the potential for selective inhibition of TIMP-1 expression in order to accelerate resorption of scar matrix in patients with liver disease.

The cross-linking of collagen by lysyl oxidase and tissue transglutaminase, and the "maturation" of hepatic scar through the action of ADAMTS2 (A Disintegrin and Metalloproteinase with ThromboSpondin type repeats metalloproteinase with thrombospondin type I motif) may regulate hepatic fibrosis reversibility. In animal models, even advanced fibrosis is reversible, limited by the extent of collagen cross-linking due to tissue transglutaminase [133] and lysyl oxidase-2 [134]. As advanced fibrosis resolves, the micronodules typical of active cirrhosis dissolve, coalescing into macronodules [133]. This finding correlates with clinical data demonstrating that increased septal thickness and smaller nodule size are significant predictors of poorer clinical outcomes [135]. Clinical trials targeting inhibition of lysyl oxidase 2 with a monoclonal antibody have been negative [136], but the molecule still remains a potential target for antifibrotic therapy using small molecule inhibitors that may more readily access the target molecule in fibrotic liver.

STELLATE CELL ACTIVATION, THE CENTRAL EVENT IN HEPATIC FIBROSIS — Hepatic stellate cell activation is the common pathway leading to hepatic fibrosis. Once they are activated, they release chemokines and other leukocyte chemoattractants while upregulating the expression of important inflammatory receptors such as ICAM-1, chemokine receptors, and mediators of lipopolysaccharide signaling [137].

Activation consists of two major phases:

Initiation (also called a "preinflammatory stage"), which refers to early changes in gene expression and phenotype that render the cells responsive to other cytokines and stimuli [70]. Initiation results mostly from paracrine stimulation.

Perpetuation results from the effects of these stimuli on maintaining the activated phenotype and generating fibrosis. Perpetuation involves autocrine as well as paracrine loops.

Initiation — The earliest changes observed during stellate activation result from paracrine stimulation by all neighboring cell types, including sinusoidal endothelium, Kupffer cells and infiltrating inflammatory monocytes, hepatocytes, and platelets. As noted above, early injury to endothelial cells stimulates production of cellular fibronectin, which has an activating effect on stellate cells [93]. Endothelial cells are also likely to participate in conversion of transforming growth factor (TGF)-beta from the latent to active, profibrogenic form. Platelets are another important source of paracrine stimuli, including platelet-derived growth factor (PDGF), TGF-beta-1, and epidermal growth factor (EGF) [138]. TGF-beta-1, derived from both paracrine and autocrine sources, is the best characterized and most potent fibrogenic cytokine.

Monocyte infiltration and activation also contribute to stellate cell activation. Kupffer cells stimulate matrix synthesis, cell proliferation, and release of retinoids by stellate cells through the actions of cytokines (especially TGF-beta-1) and reactive oxygen intermediates/lipid peroxides [139]. On the other hand, activated macrophages can also lead to stellate cell apoptosis by different mechanisms [140]. Growing appreciation for macrophage heterogeneity indicates that specific subsets serve important and sometimes divergent roles in driving or regressing fibrosis [114,141-144].

Hepatocytes are also a potent source of fibrogenic lipid peroxides. Hepatocyte apoptosis following injury also promotes stellate cell initiation through a process mediated by Fas (a protein involved in causing apoptosis) [145,146]. Apoptosis of parenchymal cells is also an important inflammatory stimulus [147]. The response of stellate cells to apoptotic hepatocytes in part reflects the interaction of hepatocyte DNA with Toll-like receptor 9 (TLR9) expressed on stellate cells [148]. A profibrogenic response can also be elicited by hepatocyte apoptosis following disruption of the antiapoptotic mediator Bcl-xL and by Fas [145,149].

Thus, efforts to block hepatocyte apoptosis therapeutically are being developed as a potential antifibrotic strategy [150]. In contrast, selective stimulation of apoptosis in stellate cells by either TRAIL [151], gliotoxin [152], or proteasome inhibitors [153] is antifibrotic. While hepatocyte necrosis associated with lipid peroxidation is considered a classical inflammatory and fibrogenic stimulus, apoptosis, or programmed cell death, has also been implicated in the fibrogenic response. Apoptotic fragments released from hepatocytes are fibrogenic towards cultured stellate cells [154], and activate Kupffer cells [155] Also, Fas-mediated hepatocyte apoptosis in vivo in experimental animals is fibrogenic [145].

The cytochrome CYP2E1 may have an important role in the generation of reactive oxygen species that stimulates hepatic stellate cells [156]. Cultured hepatic stellate cells grown in the presence of a cell that expresses CYP2E1 (E47 cells) increase the production of collagen, an effect prevented in the presence of antioxidants or a CYP2E1 inhibitor [156]. These data suggest that the CYP2E1-derived reactive oxygen species are responsible for the increased collagen production. In similar experiments using co-cultured hepatic stellate and E47 cells, the addition of arachidonic acid plus ferric nitrilotriacetate (agents that potentiate oxidative stress) further induced collagen protein synthesis [157]. These findings may help to explain the pathogenesis of liver injury in alcoholic liver disease since CYP2E1 is alcohol-inducible.

As noted above, reactive oxygen species (ROS) are generated through lipid peroxidation from hepatocytes, macrophages, stellate cells, and other inflammatory cells [158,159]. In alcohol-associated or nonalcohol-associated steatohepatitis, ROS generation in hepatocytes results from induction of cytochrome P450 2E1 [160,161], leading to pericentral (zone 3) injury. Also, NADPH oxidase (NOX) mediates fibrogenic activation in hepatic stellate cells, as well as in Kupffer cells or resident liver macrophages through generation of oxidant stress [162]. Increasing knowledge about NOX isoforms and their cell-specific activities is leading to their emergence as a therapeutic target [163]. Hepatitis C virus (HCV) as well as HCV/HIV coinfection also produces oxidant stress molecules that amplify inflammation and fibrosis [164,165].

Nitrosative stress from hepatocyte mitochondrial injury and induction of nitric oxide synthase 2 (NOS2) may also be an overlooked pathway of liver injury [166,167]. Hypoxia is typically also a component of the injury milieu and elicits fibrogenic and angiogenic signaling [27].

The identification of toll-like receptor 4 (TLR4), the receptor for bacterial lipopolysaccharide, on Kupffer cells and stellate cells introduces a role for the innate immune system in hepatic fibrosis [168]. Furthermore, signaling by stellate cells in response to LPS and possibly endogenous ligands of TLR4 (eg, high-mobility group box 1 [169,170], biglycan and heparan sulfate) may be more important than in Kupffer cells in eliciting a fibrogenic response by downregulating BAMBI, a transmembrane suppressor of TGF-beta-1, which is the major fibrogenic cytokine in the liver [171]. This finding correlates with evidence that specific single nucleotide polymorphisms of TLR4 contribute to the rate of fibrosis progression in HCV infection [172], thereby linking a genetic risk marker to disease pathogenesis. TLR2 has also been identified as a regulator of inflammation and fibrosis in fatty liver diseases [173,174].

Direct effects of hepatotrophic viruses may also contribute to stellate cell activation. In particular, HCV proteins can activate stellate cells through several mechanisms [175-177], whereas accelerated fibrosis in chronic hepatitis B virus may engage specific immune cell types, especially natural killer cells [178]. Some of the signals driving fibrosis may derive from the gut microbiome.

Gene regulation during stellate cell activation — Attention has focused on regulatory pathways that can respond quickly to injurious stimuli, either by activating or repressing gene transcription, by epigenetic regulation, or by posttranscriptional control. A number of potential antifibrotic targets are present both on the cell surface and within hepatic stellate cells (figure 3) [137]. As fibrosis advances, the collection of potential therapeutic targets evolves from those engaged in paracrine stimulation towards hepatic stellate cells, to largely autocrine interactions by this cell type [179]. These findings suggest that drugs that may be effective in earlier fibrosis may be less useful as the disease advances because the pathways and signals driving disease evolve.

Among the many target genes of transcription factors described in stellate cells, those most comprehensively characterized include type I collagen (alpha 1 and alpha 2 chains), alpha-SMA, TGF-beta-1 and TGF-beta receptors, matrix metalloproteinase (MMP)-2, and tissue inhibitors of metalloproteinases 1 and 2 [180-183]. Nuclear hormone receptors, in particular the vitamin D receptor, peroxisome proliferator-activated receptor gamma, and LXR, have been implicated in fibrogenic gene regulation in hepatic stellate cells [184-186]. Similarly, autophagy, an intracellular pathway that provides metabolic energy, is required for stellate cells to activate and may explain the hydrolysis of retinyl esters that occurs during stellate cell activation [187-189]. Its effects may be mediated in part through enhanced endoplasmic reticulum stress involving the inositol-requiring enzyme 1 (IRE1) branch of the unfolded protein response [189].

Epigenetic regulation is a tightly controlled pathway that modulates stellate cell activation in part through induction of the molecules CBF1 and MeCP2 [181,190-192].

mRNA stabilization also contributes to increased gene expression during stellate cell activation. Specifically, there is a 16-fold increase in collagen alpha 1(I) mRNA stabilization during stellate cell activation due to interaction of a specific protein, alpha CP, to a specific sequence in the 3' untranslated region of the mRNA [193], and also involving the interaction of a 120 kDa protein with the 5' stem-loop structure [194].

Perpetuation — Perpetuation of stellate cell activation involves at least seven discrete changes in cell behavior: proliferation, chemotaxis, fibrogenesis, contractility, matrix degradation, retinoid loss, and white blood cell chemoattractant and cytokine release. The net effect of these changes is to increase accumulation of extracellular matrix. As an example, proliferation and chemotaxis lead to increased numbers of collagen-producing cells as well as more matrix production per cell. Cytokine release by stellate cells can amplify the inflammatory and fibrogenic tissue responses, and matrix proteases may hasten the replacement of normal matrix with one typical of the wound "scar."

Proliferation — Platelet-derived growth factor (PDGF) is the most potent stellate cell mitogen identified [195]. Induction of PDGF receptors early in stellate cell activation increases responsiveness to this potent mitogen [196]. A co-receptor, neuropilin-1, may enhance PDGF signaling at the cell membrane of stellate cells [197]. Downstream pathways of PDGF signaling have been carefully characterized in stellate cells [198]. In addition to proliferation, PDGF stimulates Na+/H+ exchange, providing a potential site for therapeutic intervention by blocking ion transport [199]. Other compounds with mitogenic activity in stellate cells and a potential role in fibrogenesis include vascular endothelial cell growth factor [200], thrombin and its receptor [201,202], EGF, TGF-alpha, keratinocyte growth factor [203] and basic fibroblast growth factor [204]. Signaling pathways for these and other mitogens have been greatly clarified in stellate cells, offering many potential sites for therapeutic intervention [198,205]. In addition, cell cycle engagement has been characterized during stellate cell activation in response to PDGF and other mitogens [206].

PDGF-C and -D isoforms have also been described [207]. PDGF-D may be the most potent and physiologically relevant PDGF subunit in stellate cell activation [208]. Furthermore, whereas both mice with transgenic expression of either PDGF-B [207] or PDGF-C have hepatic fibrosis [200], the PDGF-C transgenic animals also develop hepatocellular carcinoma (HCC) [200], mimicking the progression from fibrosis to cancer that occurs in humans.

Chemotaxis — Stellate cells can migrate towards cytokine chemoattractants [209,210], explaining in part why stellate cells align within inflammatory septae in vivo. These chemoattractants include PDGF [211,212], MCP-1 [213], and CXCR3 [214].

In contrast to PDGF, adenosine blunts chemotaxis, thereby providing a counter-regulatory pathway that fixes cells at sites of injury [215]. Paradoxically, enhanced adenosine signaling may also contribute to alcoholic fibrosis by stimulating stellate cell fibrogenesis [216], which not only represents a potential fibrogenic mechanism, but also may explain the protective effect of caffeine (which inhibits adenosine generation) reported in epidemiologic studies [217].

Fibrogenesis — The most direct way that stellate cells influence fibrosis is by increasing matrix production and scar formation. The best-studied component of hepatic scar is collagen type I, the expression of which is regulated posttranscriptionally in hepatic stellate cells. The most potent stimulus for collagen I production is TGF-beta, which is derived from both paracrine and autocrine sources; TGF-beta also stimulates the production of other matrix components including cellular fibronectin and proteoglycans [23,218]. Other factors that stimulate collagen I by activated stellate cells in culture include retinoids, angiotensin II [219], interleukin-1 beta, tumor necrosis factor, and acetaldehyde. However, none of these is as potent as TGF-beta-1. The expression of type I collagen can be inhibited when protein factors responsible for increasing expression are blocked from binding to a conserved stem-loop in the 5' untranslated region of type I collagen mRNA, suggesting that this region is important in the regulation of collagen synthesis by stellate cells [220].

TGF-beta-1 stimulates collagen in stellate cells through a hydrogen peroxide and C/EBPb-dependent mechanism [61], which also involves Rho kinase [221]. Signals downstream of TGF-beta include a family of bifunctional molecules known as SMADs, upon which many extracellular and intracellular signals converge to fine-tune and enhance the effects of TGF-beta during fibrogenesis [198]. The response of SMADs in stellate cells differs between acute and chronic injury to further favor matrix production [222].

As mentioned above, lipid peroxidation products are emerging as important stimuli to extracellular matrix production [223] (see 'Cellular sources of ECM in normal and fibrotic liver' above). Their effects may be amplified by loss of antioxidant capacity of stellate cells as they activate [224]. These important insights have provided the rationale for the study of antioxidants in the treatment of a variety of liver diseases.

Hepatic iron concentration may also influence fibrogenesis, at least in patients with hepatitis C. Hepatic iron stores in such patients correlated with histologic disease severity and with stellate cell numbers [225].

Connective tissue growth factor (CTGF/CCN2) is also a potent fibrogenic signal towards stellate cells [226-229] that is upregulated by hyperglycemia and hyperinsulinemia [230]. Interestingly, TGF-beta stimulates CTGF primarily in hepatocytes, not stellate cells [231,232], a notable exception to the general rule that cytokine signaling in stellate cell activation is typically autocrine.

Neurohumoral signaling contributes to stellate cell responses [233]. Specifically cannabinoids are potent mediators of hepatic steatosis, stellate cell activation and fibrosis (reviewed in [234]), as well as provoking the hemodynamic alterations associated with advanced liver disease [235]. Two receptors, CB1 and CB2, exert opposing effects, with CB1 a fibrogenic pathway and CB2 antifibrotic. Thus, antagonism of CB1 signaling in stellate cells represents a promising antifibrotic strategy [236]. In addition, neurotrophin receptor signaling has emerged as a potential target based on its contributions to stellate cell activation and its regulation by thyroid hormone [237-239].

In contrast, agonism of CB2 receptors, which are also expressed by stellate cells, reverses fibrosis in experimental animals [240]. The fundamental challenge of developing cannabinoid therapeutics for liver disease is to minimize central nervous system effects, since CB1 and CB2 receptors are abundantly expressed in brain. Similarly, opioids signal in stellate cells and promote fibrogenesis [241,242], which is antagonized by naltrexone. Finally, sympathetic neurotransmitters also contribute to activation pathways [243].

Leptin, a ubiquitous adipokine, is profibrogenic, and its levels are elevated in patients who are obese [244-246]. In contrast, adiponectin is a natural counter-regulator to leptin whose activity is diminished in advanced liver diseases and obesity, raising the prospect that it may have therapeutic value as an antifibrotic [246-248]. In addition, ghrelin, an orexigenic hormone, may attenuate hepatic fibrosis based on animal studies [249].

Components of Hedgehog signaling, a potent developmental pathway, have been identified in hepatic stellate cells and contribute to their activation and fibrogenesis [250-252].

Contractility — Contractility of stellate cells may be a major determinant of early and late increases in portal resistance during liver fibrosis. The collagenous bands typical of end-stage cirrhosis contain large numbers of activated stellate cells [253]. These impede portal blood flow by constricting individual sinusoids and by contracting the cirrhotic liver. The acquisition of a contractile phenotype during stellate cell activation has been documented in culture and in vivo, and is mediated in part by receptors which interact with the extracellular matrix [198].

In the process of becoming contractile, stellate cells develop increased expression of the cytoskeletal protein alpha smooth muscle actin. If smooth muscle actin is required for contraction, then inactivating it could represent a therapeutic target for treating portal hypertension.

The major stimulus for stellate cell contraction is endothelin-1, whose receptors are expressed on quiescent and activated stellate cells [254]. Endothelin activity is antagonized by the nuclear receptor FXR [255]. With activation, receptor expression does not increase (unlike PDGF receptors), but there is a shift in the predominant type of endothelin receptor and increased sensitivity to autocrine endothelin-1 [253,256]. Locally produced vasodilator substances (particularly nitric oxide) may counteract the constrictive effects of endothelin-1 [253]. In vivo studies suggest that carbon monoxide also mediates sinusoidal relaxation through its effects on stellate cells [257].

Matrix degradation — As discussed above, quantitative and qualitative changes in matrix protease activity have an important role in extracellular matrix remodeling accompanying fibrosing liver injury. Because stellate cells express virtually all the components required for pathologic matrix degradation, they have a key role not only in matrix production, but also matrix degradation.

Retinoid loss — Activation of stellate cells is accompanied by the loss of the characteristic perinuclear retinoid (vitamin A) droplets. In culture, retinoid is stored as retinyl esters, whereas the form of retinoid released outside the cell during activation is retinol, suggesting that there is intracellular hydrolysis of esters prior to export [85]. Whether retinoid loss is required for stellate cells to activate, and which retinoids might accelerate or prevent activation, is incompletely understood. The enzyme PNPLA3 has been proposed as a "gatekeeper" of stellate cell retinoid content, and interestingly, polymorphisms in this gene have also been linked to risk of nonalcoholic fatty liver disease [258]. Free cholesterol may also have an activating role towards stellate cells [259].

Several nuclear retinoid receptors have been identified in stellate cells [260]. These molecules bind intracellular retinoid ligands and regulate gene expression, but it is uncertain whether they have a regulatory role in fibrogenesis. The question has important clinical implications since efforts are being made to use retinoids therapeutically.

Peroxisome proliferator-activated receptors (PPAR), in particular PPARg, have been identified in stellate cells [261,262]. Their expression decreases with activation [261,262]. Ligands for this newly identified nuclear receptor family downregulate stellate cell activation [262].

The intracellular lipid storage protein, adipose differentiation-related protein (ADRP), has been uncovered in stellate cells; its expression is reduced during cellular activation, and it is induced by retinoid exposure, suggesting that ADRP may have a regulatory role linking lipid content to cellular activation through an unknown mechanism [263].

Elegant mechanisms of gene regulation mediated by the vitamin D receptor point towards a potential therapeutic target since vitamin D appears to be antifibrotic [184,264].

Inflammatory signaling and WBC chemoattraction — Stellate cells are assuming an increasingly central role in our understanding of hepatic inflammation. They can amplify the inflammatory response by inducing infiltration of mono- and polymorphonuclear leukocytes. Activated stellate cells produce chemokines that include MCP-1 [265], CCL21 [266], RANTES, and CCR5 [267], among several others [268-271]. These molecules are particularly attractive therapeutic targets, in part because chemokine receptors are G-coupled protein receptors, which are especially targetable by antagonists [272]. Stellate cells also express toll-like receptors (see above) [168], indicating a capacity to interact with bacterial lipopolysaccharide, which in turn stimulates stellate cells [273,274]. Stellate cells can also function as antigen-presenting cells [275] that can stimulate lymphocyte proliferation or apoptosis [276]. Stellate cells produce neutrophil chemoattractants, which could contribute to the neutrophil accumulation characteristic of alcoholic liver disease.

In addition to regulating leukocyte behavior, stellate cells may in turn be affected by specific lymphocyte populations [277]. As an example, CD8 cells harbor more fibrogenic activity towards stellate cells than CD4 cells [94], which may explain in part the increased hepatic fibrosis seen in patients with HCV/HIV coinfection, where CD4/CD8 ratios are reduced, than in patients mono-infected with HCV alone.

Links between stellate and progenitor cells, fibrosis, and cancer — The observation that stellate cells express the stem cell marker CD133 raises the possibility that they are true progenitor cells [278-280]. Importantly, the requirement for fibrosis to occur before HCC develops in patients with chronic HCV remains unexplained. Potential explanations have included the presence of secreted survival factors that prevent apoptosis of DNA-damaged hepatocytes and activated stellate cells (eg, Gas6 [281]), reduced tumor surveillance due to decreasing NK cell number and function, and/or the accelerated shortening of telomeres that accompanies progressive fibrosis. How fibrosis promotes HCC is an important, unanswered question. An intimate, yet unexplained, physical link between activated stellate cells, the ductal reaction (clusters of bile ductular expansion), and fibrosis raises important questions about the interactions of these cell types and their roles in fibrosis and regeneration [282].

Reversion of stellate cell activation and the role of senescence — The fate of activated stellate cells as liver injury resolves has been an ongoing focus of investigation. Earlier studies implicated apoptosis, or programmed cell death, as an important pathway by which the liver clears activated stellate cells [129]. An additional pathway has emerged in which activated cells "revert" back to a quiescent phenotype [283-285]. Interestingly, reverted cells harbor the capacity to reactivate more quickly and extensively than cells that have never been activated. Thus, efforts to "turn off" stellate cell activation represent a therapeutic focus in efforts to regress fibrosis. A growing list of "de-activating" factors and epigenetic modifications underlying de-activation/reversion offers clues to potential therapeutic targets, including GATA4, Lhx2, Tcf21, NF1 and others [179,286].

Senescence of hepatic stellate cells has also emerged as an important property of this cell type. The senescence-associated secretory phenotype is a distinct phenotype of activated stellate cells that can both limit the extent of fibrosis and alter the microenvironment, thereby regulating the polarity of macrophages and the ability to resist the development of HCC [287-289]. As noted above, selective clearance of senescent stellate cells is antifibrotic in mouse models of fibrosis and may be a potential strategy for human disease [109]. Interleukin-22, BMP4, and Gremlin 1 have been identified as potential regulators of stellate cell senescence [290,291].

SUMMARY

Background – Emerging insights into the pathobiology of hepatic fibrosis point to new targets for antifibrotic therapy and heightened prospects for success. (See 'Introduction' above.)

Extracellular matrix (ECM) refers to a group of macromolecules that comprise the scaffolding of normal and fibrotic liver. These include collagens, noncollagen glycoproteins, matrix-bound growth factors, glycosaminoglycans, proteoglycans, and matricellular proteins. A great deal of progress has been made in identifying new members of these families and in understanding how these molecules interact. (See 'Extracellular matrix composition of the normal and fibrotic liver' above.)

Matrix alterations observed during fibrogenesis alter cellular behavior by processes involving cell membrane receptors. (See 'Biologic activity of extracellular matrix' above.)

Cellular sources of extracellular matrix – A major advance in the understanding of hepatic fibrosis has been the identification of the cellular sources of ECM. The hepatic stellate cell (previously called the lipocyte, Ito, fat-storing, or perisinusoidal cell) is the primary source of ECM in normal and fibrotic liver. (See 'Cellular sources of ECM in normal and fibrotic liver' above.)

Degradation of extracellular matrix – Fibrosis reflects a balance between matrix production and degradation. The degradation of extracellular matrix is a key event in hepatic fibrosis. Early disruption of the normal hepatic matrix by matrix proteases hastens its replacement by scar matrix, which has deleterious effects on cell function. Macrophages are emerging as important determinants of fibrosis regression. (See 'Degradation of extracellular matrix' above.)

Hepatic stellate cell activation – Hepatic stellate cell activation is the common pathway leading to hepatic fibrosis. Once activated, hepatic stellate cells release chemokines and other leukocyte chemoattractants, while upregulating the expression of important inflammatory receptors such as ICAM-1, chemokine receptors, and mediators of lipopolysaccharide signaling. (See 'Stellate cell activation, the central event in hepatic fibrosis' above.)

  1. Poynard T, Mathurin P, Lai CL, et al. A comparison of fibrosis progression in chronic liver diseases. J Hepatol 2003; 38:257.
  2. Bataller R, North KE, Brenner DA. Genetic polymorphisms and the progression of liver fibrosis: a critical appraisal. Hepatology 2003; 37:493.
  3. Hillebrandt S, Goos C, Matern S, Lammert F. Genome-wide analysis of hepatic fibrosis in inbred mice identifies the susceptibility locus Hfib1 on chromosome 15. Gastroenterology 2002; 123:2041.
  4. Friedman SL. Evolving challenges in hepatic fibrosis. Nat Rev Gastroenterol Hepatol 2010; 7:425.
  5. Ismair MG, Stieger B, Cattori V, et al. Hepatic uptake of cholecystokinin octapeptide by organic anion-transporting polypeptides OATP4 and OATP8 of rat and human liver. Gastroenterology 2001; 121:1185.
  6. Bonis PA, Friedman SL, Kaplan MM. Is liver fibrosis reversible? N Engl J Med 2001; 344:452.
  7. Dienstag JL, Goldin RD, Heathcote EJ, et al. Histological outcome during long-term lamivudine therapy. Gastroenterology 2003; 124:105.
  8. Falize L, Guillygomarc'h A, Perrin M, et al. Reversibility of hepatic fibrosis in treated genetic hemochromatosis: a study of 36 cases. Hepatology 2006; 44:472.
  9. Mallet V, Gilgenkrantz H, Serpaggi J, et al. Brief communication: the relationship of regression of cirrhosis to outcome in chronic hepatitis C. Ann Intern Med 2008; 149:399.
  10. Veldt BJ, Heathcote EJ, Wedemeyer H, et al. Sustained virologic response and clinical outcomes in patients with chronic hepatitis C and advanced fibrosis. Ann Intern Med 2007; 147:677.
  11. Roberts S, Gordon A, McLean C, et al. Effect of sustained viral response on hepatic venous pressure gradient in hepatitis C-related cirrhosis. Clin Gastroenterol Hepatol 2007; 5:932.
  12. Friedman SL. Hepatic Fibrosis: The consequences of liver disease. In: Schiff's Diseases of the Liver, 9th edition, Lippincott, Williams & Wilkins, 2002.
  13. Lee YA, Wallace MC, Friedman SL. Pathobiology of liver fibrosis: a translational success story. Gut 2015; 64:830.
  14. Schuppan D, Ruehl M, Somasundaram R, Hahn EG. Matrix as a modulator of hepatic fibrogenesis. Semin Liver Dis 2001; 21:351.
  15. Villesen IF, Daniels SJ, Leeming DJ, et al. Review article: the signalling and functional role of the extracellular matrix in the development of liver fibrosis. Aliment Pharmacol Ther 2020; 52:85.
  16. Karsdal M. Biochemistry of collagens, laminins and elastin; structure, function and biomarkers, Academic Press, 2016.
  17. Perepelyuk M, Terajima M, Wang AY, et al. Hepatic stellate cells and portal fibroblasts are the major cellular sources of collagens and lysyl oxidases in normal liver and early after injury. Am J Physiol Gastrointest Liver Physiol 2013; 304:G605.
  18. Muiznieks LD, Keeley FW. Molecular assembly and mechanical properties of the extracellular matrix: A fibrous protein perspective. Biochim Biophys Acta 2013; 1832:866.
  19. Miao M, Sitarz E, Bellingham CM, et al. Sequence and domain arrangements influence mechanical properties of elastin-like polymeric elastomers. Biopolymers 2013; 99:392.
  20. Pellicoro A, Ramachandran P, Iredale JP. Reversibility of liver fibrosis. Fibrogenesis Tissue Repair 2012; 5:S26.
  21. Pellicoro A, Aucott RL, Ramachandran P, et al. Elastin accumulation is regulated at the level of degradation by macrophage metalloelastase (MMP-12) during experimental liver fibrosis. Hepatology 2012; 55:1965.
  22. Rojkind M, Giambrone MA, Biempica L. Collagen types in normal and cirrhotic liver. Gastroenterology 1979; 76:710.
  23. Gressner AM. The cell biology of liver fibrogenesis - an imbalance of proliferation, growth arrest and apoptosis of myofibroblasts. Cell Tissue Res 1998; 292:447.
  24. McGuire RF, Bissell DM, Boyles J, Roll FJ. Role of extracellular matrix in regulating fenestrations of sinusoidal endothelial cells isolated from normal rat liver. Hepatology 1992; 15:989.
  25. Hammoutene A, Rautou PE. Role of liver sinusoidal endothelial cells in non-alcoholic fatty liver disease. J Hepatol 2019; 70:1278.
  26. Lee JS, Semela D, Iredale J, Shah VH. Sinusoidal remodeling and angiogenesis: a new function for the liver-specific pericyte? Hepatology 2007; 45:817.
  27. Rosmorduc O, Housset C. Hypoxia: a link between fibrogenesis, angiogenesis, and carcinogenesis in liver disease. Semin Liver Dis 2010; 30:258.
  28. Dev A, Patel K, Conrad A, et al. Relationship of smoking and fibrosis in patients with chronic hepatitis C. Clin Gastroenterol Hepatol 2006; 4:797.
  29. Yang C, Zeisberg M, Mosterman B, et al. Liver fibrosis: insights into migration of hepatic stellate cells in response to extracellular matrix and growth factors. Gastroenterology 2003; 124:147.
  30. Zhou X, Murphy FR, Gehdu N, et al. Engagement of alphavbeta3 integrin regulates proliferation and apoptosis of hepatic stellate cells. J Biol Chem 2004; 279:23996.
  31. Melton AC, Soon RK Jr, Park JG, et al. Focal adhesion disassembly is an essential early event in hepatic stellate cell chemotaxis. Am J Physiol Gastrointest Liver Physiol 2007; 293:G1272.
  32. Yee HF Jr. Rho directs activation-associated changes in rat hepatic stellate cell morphology via regulation of the actin cytoskeleton. Hepatology 1998; 28:843.
  33. Choi SS, Sicklick JK, Ma Q, et al. Sustained activation of Rac1 in hepatic stellate cells promotes liver injury and fibrosis in mice. Hepatology 2006; 44:1267.
  34. Schuppan D, Schmid M, Somasundaram R, et al. Collagens in the liver extracellular matrix bind hepatocyte growth factor. Gastroenterology 1998; 114:139.
  35. Geervliet E, Bansal R. Matrix Metalloproteinases as Potential Biomarkers and Therapeutic Targets in Liver Diseases. Cells 2020; 9.
  36. Wells RG. The role of matrix stiffness in regulating cell behavior. Hepatology 2008; 47:1394.
  37. Kazemi F, Kettaneh A, N'kontchou G, et al. Liver stiffness measurement selects patients with cirrhosis at risk of bearing large oesophageal varices. J Hepatol 2006; 45:230.
  38. Yin M, Talwalkar JA, Glaser KJ, et al. Assessment of hepatic fibrosis with magnetic resonance elastography. Clin Gastroenterol Hepatol 2007; 5:1207.
  39. Ruoslahti E, Engvall E. Integrins and vascular extracellular matrix assembly. J Clin Invest 1997; 99:1149.
  40. Patsenker E, Popov Y, Stickel F, et al. Inhibition of integrin alphavbeta6 on cholangiocytes blocks transforming growth factor-beta activation and retards biliary fibrosis progression. Gastroenterology 2008; 135:660.
  41. Henderson NC, Arnold TD, Katamura Y, et al. Targeting of αv integrin identifies a core molecular pathway that regulates fibrosis in several organs. Nat Med 2013; 19:1617.
  42. Reed NI, Jo H, Chen C, et al. The αvβ1 integrin plays a critical in vivo role in tissue fibrosis. Sci Transl Med 2015; 7:288ra79.
  43. Martin K, Pritchett J, Llewellyn J, et al. PAK proteins and YAP-1 signalling downstream of integrin beta-1 in myofibroblasts promote liver fibrosis. Nat Commun 2016; 7:12502.
  44. Carloni V, Defranco RM, Caligiuri A, et al. Cell adhesion regulates platelet-derived growth factor-induced MAP kinase and PI-3 kinase activation in stellate cells. Hepatology 2002; 36:582.
  45. Carloni V, Pinzani M, Giusti S, et al. Tyrosine phosphorylation of focal adhesion kinase by PDGF is dependent on ras in human hepatic stellate cells. Hepatology 2000; 31:131.
  46. Couvelard A, Scoazec JY, Feldmann G. Expression of cell-cell and cell-matrix adhesion proteins by sinusoidal endothelial cells in the normal and cirrhotic human liver. Am J Pathol 1993; 143:738.
  47. Friedman SL. Stellate cells: a moving target in hepatic fibrogenesis. Hepatology 2004; 40:1041.
  48. Racine-Samson L, Rockey DC, Bissell DM. The role of alpha1beta1 integrin in wound contraction. A quantitative analysis of liver myofibroblasts in vivo and in primary culture. J Biol Chem 1997; 272:30911.
  49. https://clinicaltrials.gov/ct2/show/NCT03183570 (Accessed on December 30, 2021).
  50. Adams DH, Hubscher SG, Fisher NC, et al. Expression of E-selectin and E-selectin ligands in human liver inflammation. Hepatology 1996; 24:533.
  51. Adams DH, Burra P, Hubscher SG, et al. Endothelial activation and circulating vascular adhesion molecules in alcoholic liver disease. Hepatology 1994; 19:588.
  52. Nakanuma Y, Yasoshima M, Tsuneyama K, Harada K. Histopathology of primary biliary cirrhosis with emphasis on expression of adhesion molecules. Semin Liver Dis 1997; 17:35.
  53. Olaso E, Ikeda K, Eng FJ, et al. DDR2 receptor promotes MMP-2-mediated proliferation and invasion by hepatic stellate cells. J Clin Invest 2001; 108:1369.
  54. Luo Z, Liu H, Sun X, et al. RNA interference against discoidin domain receptor 2 ameliorates alcoholic liver disease in rats. PLoS One 2013; 8:e55860.
  55. Vlodavsky I, Miao HQ, Medalion B, et al. Involvement of heparan sulfate and related molecules in sequestration and growth promoting activity of fibroblast growth factor. Cancer Metastasis Rev 1996; 15:177.
  56. Inagaki Y, Okazaki I. Emerging insights into Transforming growth factor beta Smad signal in hepatic fibrogenesis. Gut 2007; 56:284.
  57. Breitkopf K, Godoy P, Ciuclan L, et al. TGF-beta/Smad signaling in the injured liver. Z Gastroenterol 2006; 44:57.
  58. Liu C, Gaça MD, Swenson ES, et al. Smads 2 and 3 are differentially activated by transforming growth factor-beta (TGF-beta ) in quiescent and activated hepatic stellate cells. Constitutive nuclear localization of Smads in activated cells is TGF-beta-independent. J Biol Chem 2003; 278:11721.
  59. Uemura M, Swenson ES, Gaça MD, et al. Smad2 and Smad3 play different roles in rat hepatic stellate cell function and alpha-smooth muscle actin organization. Mol Biol Cell 2005; 16:4214.
  60. Wiercinska E, Wickert L, Denecke B, et al. Id1 is a critical mediator in TGF-beta-induced transdifferentiation of rat hepatic stellate cells. Hepatology 2006; 43:1032.
  61. García-Trevijano ER, Iraburu MJ, Fontana L, et al. Transforming growth factor beta1 induces the expression of alpha1(I) procollagen mRNA by a hydrogen peroxide-C/EBPbeta-dependent mechanism in rat hepatic stellate cells. Hepatology 1999; 29:960.
  62. Dooley S, Delvoux B, Lahme B, et al. Modulation of transforming growth factor beta response and signaling during transdifferentiation of rat hepatic stellate cells to myofibroblasts. Hepatology 2000; 31:1094.
  63. Liu C, Billadeau DD, Abdelhakim H, et al. IQGAP1 suppresses TβRII-mediated myofibroblastic activation and metastatic growth in liver. J Clin Invest 2013; 123:1138.
  64. Wells RG, Kruglov E, Dranoff JA. Autocrine release of TGF-beta by portal fibroblasts regulates cell growth. FEBS Lett 2004; 559:107.
  65. Kinnman N, Housset C. Peribiliary myofibroblasts in biliary type liver fibrosis. Front Biosci 2002; 7:d496.
  66. Kruglov EA, Jain D, Dranoff JA. Isolation of primary rat liver fibroblasts. J Investig Med 2002; 50:179.
  67. Forbes SJ, Russo FP, Rey V, et al. A significant proportion of myofibroblasts are of bone marrow origin in human liver fibrosis. Gastroenterology 2004; 126:955.
  68. Kalluri R, Neilson EG. Epithelial-mesenchymal transition and its implications for fibrosis. J Clin Invest 2003; 112:1776.
  69. Friedman SL. Hepatic stellate cells: protean, multifunctional, and enigmatic cells of the liver. Physiol Rev 2008; 88:125.
  70. Friedman SL. Molecular regulation of hepatic fibrosis, an integrated cellular response to tissue injury. J Biol Chem 2000; 275:2247.
  71. The Hepatic Stellate Cell, Friedman SL (Ed), Thieme, New York 2001. Vol Vol 21.
  72. Geerts A. History, heterogeneity, developmental biology, and functions of quiescent hepatic stellate cells. Semin Liver Dis 2001; 21:311.
  73. Xiong X, Kuang H, Ansari S, et al. Landscape of Intercellular Crosstalk in Healthy and NASH Liver Revealed by Single-Cell Secretome Gene Analysis. Mol Cell 2019; 75:644.
  74. Dobie R, Wilson-Kanamori JR, Henderson BEP, et al. Single-Cell Transcriptomics Uncovers Zonation of Function in the Mesenchyme during Liver Fibrosis. Cell Rep 2019; 29:1832.
  75. Ramachandran P, Dobie R, Wilson-Kanamori JR, et al. Resolving the fibrotic niche of human liver cirrhosis at single-cell level. Nature 2019; 575:512.
  76. Krenkel O, Hundertmark J, Ritz TP, et al. Single Cell RNA Sequencing Identifies Subsets of Hepatic Stellate Cells and Myofibroblasts in Liver Fibrosis. Cells 2019; 8.
  77. Wang S, Li K, Pickholz E, et al. An autocrine signaling circuit in hepatic stellate cells underlies advanced fibrosis in nonalcoholic steatohepatitis. Sci Transl Med 2023; 15:eadd3949.
  78. Fred RG, Steen Pedersen J, Thompson JJ, et al. Single-cell transcriptome and cell type-specific molecular pathways of human non-alcoholic steatohepatitis. Sci Rep 2022; 12:13484.
  79. Bachem MG, Schneider E, Gross H, et al. Identification, culture, and characterization of pancreatic stellate cells in rats and humans. Gastroenterology 1998; 115:421.
  80. Magness ST, Bataller R, Yang L, Brenner DA. A dual reporter gene transgenic mouse demonstrates heterogeneity in hepatic fibrogenic cell populations. Hepatology 2004; 40:1151.
  81. Bachem MG, Schünemann M, Ramadani M, et al. Pancreatic carcinoma cells induce fibrosis by stimulating proliferation and matrix synthesis of stellate cells. Gastroenterology 2005; 128:907.
  82. Cogliati B, Yashaswini CN, Wang S, et al. Friend or foe? The elusive role of hepatic stellate cells in liver cancer. Nat Rev Gastroenterol Hepatol 2023; 20:647.
  83. Cantallops Vilà P, Ravichandra A, Agirre Lizaso A, et al. Heterogeneity, crosstalk, and targeting of cancer-associated fibroblasts in cholangiocarcinoma. Hepatology 2023.
  84. Apte MV, Wilson JS. Mechanisms of pancreatic fibrosis. Dig Dis 2004; 22:273.
  85. Friedman SL. Hepatic stellate cells. Prog Liver Dis 1996; 14:101.
  86. Reeves HL, Burt AD, Wood S, Day CP. Hepatic stellate cell activation occurs in the absence of hepatitis in alcoholic liver disease and correlates with the severity of steatosis. J Hepatol 1996; 25:677.
  87. Enzan H, Himeno H, Iwamura S, et al. Sequential changes in human Ito cells and their relation to postnecrotic liver fibrosis in massive and submassive hepatic necrosis. Virchows Arch 1995; 426:95.
  88. Enzan H, Himeno H, Iwamura S, et al. Alpha-smooth muscle actin-positive perisinusoidal stromal cells in human hepatocellular carcinoma. Hepatology 1994; 19:895.
  89. Schmitt-Gräff A, Krüger S, Bochard F, et al. Modulation of alpha smooth muscle actin and desmin expression in perisinusoidal cells of normal and diseased human livers. Am J Pathol 1991; 138:1233.
  90. Maher JJ, McGuire RF. Extracellular matrix gene expression increases preferentially in rat lipocytes and sinusoidal endothelial cells during hepatic fibrosis in vivo. J Clin Invest 1990; 86:1641.
  91. Herbst H, Frey A, Heinrichs O, et al. Heterogeneity of liver cells expressing procollagen types I and IV in vivo. Histochem Cell Biol 1997; 107:399.
  92. Roskams T, Moshage H, De Vos R, et al. Heparan sulfate proteoglycan expression in normal human liver. Hepatology 1995; 21:950.
  93. Jarnagin WR, Rockey DC, Koteliansky VE, et al. Expression of variant fibronectins in wound healing: cellular source and biological activity of the EIIIA segment in rat hepatic fibrogenesis. J Cell Biol 1994; 127:2037.
  94. Safadi R, Ohta M, Alvarez CE, et al. Immune stimulation of hepatic fibrogenesis by CD8 cells and attenuation by transgenic interleukin-10 from hepatocytes. Gastroenterology 2004; 127:870.
  95. Vranjkovic A, Deonarine F, Kaka S, et al. Direct-Acting Antiviral Treatment of HCV Infection Does Not Resolve the Dysfunction of Circulating CD8+ T-Cells in Advanced Liver Disease. Front Immunol 2019; 10:1926.
  96. DeLeve LD. Liver sinusoidal endothelial cells in hepatic fibrosis. Hepatology 2015; 61:1740.
  97. Ding BS, Cao Z, Lis R, et al. Divergent angiocrine signals from vascular niche balance liver regeneration and fibrosis. Nature 2014; 505:97.
  98. El-Youssef M, Mu Y, Huang L, et al. Increased expression of transforming growth factor-beta1 and thrombospondin-1 in congenital hepatic fibrosis: possible role of the hepatic stellate cell. J Pediatr Gastroenterol Nutr 1999; 28:386.
  99. Jhandier MN, Kruglov EA, Lavoie EG, et al. Portal fibroblasts regulate the proliferation of bile duct epithelia via expression of NTPDase2. J Biol Chem 2005; 280:22986.
  100. Beaussier M, Wendum D, Schiffer E, et al. Prominent contribution of portal mesenchymal cells to liver fibrosis in ischemic and obstructive cholestatic injuries. Lab Invest 2007; 87:292.
  101. Kisseleva T, Uchinami H, Feirt N, et al. Bone marrow-derived fibrocytes participate in pathogenesis of liver fibrosis. J Hepatol 2006; 45:429.
  102. Zeisberg M, Yang C, Martino M, et al. Fibroblasts derive from hepatocytes in liver fibrosis via epithelial to mesenchymal transition. J Biol Chem 2007; 282:23337.
  103. Kim HY, Sakane S, Eguileor A, et al. The Origin and Fate of Liver Myofibroblasts. Cell Mol Gastroenterol Hepatol 2024; 17:93.
  104. Kruglov EA, Nathanson RA, Nguyen T, Dranoff JA. Secretion of MCP-1/CCL2 by bile duct epithelia induces myofibroblastic transdifferentiation of portal fibroblasts. Am J Physiol Gastrointest Liver Physiol 2006; 290:G765.
  105. Dranoff JA, Ogawa M, Kruglov EA, et al. Expression of P2Y nucleotide receptors and ectonucleotidases in quiescent and activated rat hepatic stellate cells. Am J Physiol Gastrointest Liver Physiol 2004; 287:G417.
  106. Mederacke I, Hsu CC, Troeger JS, et al. Fate tracing reveals hepatic stellate cells as dominant contributors to liver fibrosis independent of its aetiology. Nat Commun 2013; 4:2823.
  107. Ramachandran P, Matchett KP, Dobie R, et al. Single-cell technologies in hepatology: new insights into liver biology and disease pathogenesis. Nat Rev Gastroenterol Hepatol 2020; 17:457.
  108. Rosenthal SB, Liu X, Ganguly S, et al. Heterogeneity of HSCs in a Mouse Model of NASH. Hepatology 2021; 74:667.
  109. Amor C, Feucht J, Leibold J, et al. Senolytic CAR T cells reverse senescence-associated pathologies. Nature 2020; 583:127.
  110. Rockey DC, Friedman SL. Fibrosis Regression After Eradication of Hepatitis C Virus: From Bench to Bedside. Gastroenterology 2021; 160:1502.
  111. Benyon RC, Arthur MJ. Extracellular matrix degradation and the role of hepatic stellate cells. Semin Liver Dis 2001; 21:373.
  112. Roderfeld M. Matrix metalloproteinase functions in hepatic injury and fibrosis. Matrix Biol 2018; 68-69:452.
  113. Sato T, Head KZ, Li J, et al. Fibrosis resolution in the mouse liver: Role of Mmp12 and potential role of calpain 1/2. Matrix Biol Plus 2023; 17:100127.
  114. Ramachandran P, Iredale JP, Fallowfield JA. Resolution of liver fibrosis: basic mechanisms and clinical relevance. Semin Liver Dis 2015; 35:119.
  115. Arthur MJ, Stanley A, Iredale JP, et al. Secretion of 72 kDa type IV collagenase/gelatinase by cultured human lipocytes. Analysis of gene expression, protein synthesis and proteinase activity. Biochem J 1992; 287 ( Pt 3):701.
  116. Milani S, Herbst H, Schuppan D, et al. Differential expression of matrix-metalloproteinase-1 and -2 genes in normal and fibrotic human liver. Am J Pathol 1994; 144:528.
  117. Vyas SK, Leyland H, Gentry J, Arthur MJ. Rat hepatic lipocytes synthesize and secrete transin (stromelysin) in early primary culture. Gastroenterology 1995; 109:889.
  118. Théret N, Musso O, L'Helgoualc'h A, Clément B. Activation of matrix metalloproteinase-2 from hepatic stellate cells requires interactions with hepatocytes. Am J Pathol 1997; 150:51.
  119. Théret N, Lehti K, Musso O, Clément B. MMP2 activation by collagen I and concanavalin A in cultured human hepatic stellate cells. Hepatology 1999; 30:462.
  120. Benyon RC, Iredale JP, Goddard S, et al. Expression of tissue inhibitor of metalloproteinases 1 and 2 is increased in fibrotic human liver. Gastroenterology 1996; 110:821.
  121. Arii S, Mise M, Harada T, et al. Overexpression of matrix metalloproteinase 9 gene in hepatocellular carcinoma with invasive potential. Hepatology 1996; 24:316.
  122. Ramachandran P, Iredale JP. Macrophages: central regulators of hepatic fibrogenesis and fibrosis resolution. J Hepatol 2012; 56:1417.
  123. Jiao J, Sastre D, Fiel MI, et al. Dendritic cell regulation of carbon tetrachloride-induced murine liver fibrosis regression. Hepatology 2012; 55:244.
  124. Zimmermann HW, Trautwein C, Tacke F. Functional role of monocytes and macrophages for the inflammatory response in acute liver injury. Front Physiol 2012; 3:56.
  125. Pellicoro A, Ramachandran P, Iredale JP, Fallowfield JA. Liver fibrosis and repair: immune regulation of wound healing in a solid organ. Nat Rev Immunol 2014; 14:181.
  126. Iredale JP, Bataller R. Identifying molecular factors that contribute to resolution of liver fibrosis. Gastroenterology 2014; 146:1160.
  127. Milani S, Herbst H, Schuppan D, et al. Cellular sources of extracellular matrix proteins in normal and fibrotic liver. Studies of gene expression by in situ hybridization. J Hepatol 1995; 22:71.
  128. Murawaki Y, Ikuta Y, Idobe Y, et al. Tissue inhibitor of metalloproteinase-1 in the liver of patients with chronic liver disease. J Hepatol 1997; 26:1213.
  129. Iredale JP, Benyon RC, Pickering J, et al. Mechanisms of spontaneous resolution of rat liver fibrosis. Hepatic stellate cell apoptosis and reduced hepatic expression of metalloproteinase inhibitors. J Clin Invest 1998; 102:538.
  130. Herbst H, Wege T, Milani S, et al. Tissue inhibitor of metalloproteinase-1 and -2 RNA expression in rat and human liver fibrosis. Am J Pathol 1997; 150:1647.
  131. Iredale JP. Hepatic stellate cell behavior during resolution of liver injury. Semin Liver Dis 2001; 21:427.
  132. Trim JE, Samra SK, Arthur MJ, et al. Upstream tissue inhibitor of metalloproteinases-1 (TIMP-1) element-1, a novel and essential regulatory DNA motif in the human TIMP-1 gene promoter, directly interacts with a 30-kDa nuclear protein. J Biol Chem 2000; 275:6657.
  133. Issa R, Zhou X, Constandinou CM, et al. Spontaneous recovery from micronodular cirrhosis: evidence for incomplete resolution associated with matrix cross-linking. Gastroenterology 2004; 126:1795.
  134. Barry-Hamilton V, Spangler R, Marshall D, et al. Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a pathologic microenvironment. Nat Med 2010; 16:1009.
  135. Nagula S, Jain D, Groszmann RJ, Garcia-Tsao G. Histological-hemodynamic correlation in cirrhosis-a histological classification of the severity of cirrhosis. J Hepatol 2006; 44:111.
  136. Sanyal AJ, Harrison SA, Ratziu V, et al. The Natural History of Advanced Fibrosis Due to Nonalcoholic Steatohepatitis: Data From the Simtuzumab Trials. Hepatology 2019; 70:1913.
  137. Tsuchida T, Friedman SL. Mechanisms of hepatic stellate cell activation. Nat Rev Gastroenterol Hepatol 2017; 14:397.
  138. Bachem MG, Melchior R, Gressner AM. The role of thrombocytes in liver fibrogenesis: effects of platelet lysate and thrombocyte-derived growth factors on the mitogenic activity and glycosaminoglycan synthesis of cultured rat liver fat storing cells. J Clin Chem Clin Biochem 1989; 27:555.
  139. Tacke F. Targeting hepatic macrophages to treat liver diseases. J Hepatol 2017; 66:1300.
  140. Fischer R, Cariers A, Reinehr R, Häussinger D. Caspase 9-dependent killing of hepatic stellate cells by activated Kupffer cells. Gastroenterology 2002; 123:845.
  141. Fabre T, Barron AMS, Christensen SM, et al. Identification of a broadly fibrogenic macrophage subset induced by type 3 inflammation. Sci Immunol 2023; 8:eadd8945.
  142. Ramachandran P, Pellicoro A, Vernon MA, et al. Differential Ly-6C expression identifies the recruited macrophage phenotype, which orchestrates the regression of murine liver fibrosis. Proc Natl Acad Sci U S A 2012; 109:E3186.
  143. Moreno-Lanceta A, Medrano-Bosch M, Fundora Y, et al. RNF41 orchestrates macrophage-driven fibrosis resolution and hepatic regeneration. Sci Transl Med 2023; 15:eabq6225.
  144. Guilliams M, Bonnardel J, Haest B, et al. Spatial proteogenomics reveals distinct and evolutionarily conserved hepatic macrophage niches. Cell 2022; 185:379.
  145. Canbay A, Higuchi H, Bronk SF, et al. Fas enhances fibrogenesis in the bile duct ligated mouse: a link between apoptosis and fibrosis. Gastroenterology 2002; 123:1323.
  146. Mehal W, Imaeda A. Cell death and fibrogenesis. Semin Liver Dis 2010; 30:226.
  147. Jaeschke H. Inflammation in response to hepatocellular apoptosis. Hepatology 2002; 35:964.
  148. Watanabe A, Hashmi A, Gomes DA, et al. Apoptotic hepatocyte DNA inhibits hepatic stellate cell chemotaxis via toll-like receptor 9. Hepatology 2007; 46:1509.
  149. Takehara T, Tatsumi T, Suzuki T, et al. Hepatocyte-specific disruption of Bcl-xL leads to continuous hepatocyte apoptosis and liver fibrotic responses. Gastroenterology 2004; 127:1189.
  150. Pockros PJ, Schiff ER, Shiffman ML, et al. Oral IDN-6556, an antiapoptotic caspase inhibitor, may lower aminotransferase activity in patients with chronic hepatitis C. Hepatology 2007; 46:324.
  151. Taimr P, Higuchi H, Kocova E, et al. Activated stellate cells express the TRAIL receptor-2/death receptor-5 and undergo TRAIL-mediated apoptosis. Hepatology 2003; 37:87.
  152. Wright MC, Issa R, Smart DE, et al. Gliotoxin stimulates the apoptosis of human and rat hepatic stellate cells and enhances the resolution of liver fibrosis in rats. Gastroenterology 2001; 121:685.
  153. Anan A, Baskin-Bey ES, Bronk SF, et al. Proteasome inhibition induces hepatic stellate cell apoptosis. Hepatology 2006; 43:335.
  154. Canbay A, Taimr P, Torok N, et al. Apoptotic body engulfment by a human stellate cell line is profibrogenic. Lab Invest 2003; 83:655.
  155. Canbay A, Feldstein AE, Higuchi H, et al. Kupffer cell engulfment of apoptotic bodies stimulates death ligand and cytokine expression. Hepatology 2003; 38:1188.
  156. Nieto N, Friedman SL, Cederbaum AI. Cytochrome P450 2E1-derived reactive oxygen species mediate paracrine stimulation of collagen I protein synthesis by hepatic stellate cells. J Biol Chem 2002; 277:9853.
  157. Nieto N, Friedman SL, Cederbaum AI. Stimulation and proliferation of primary rat hepatic stellate cells by cytochrome P450 2E1-derived reactive oxygen species. Hepatology 2002; 35:62.
  158. Parola M, Robino G. Oxidative stress-related molecules and liver fibrosis. J Hepatol 2001; 35:297.
  159. Jaeschke H. Mechanisms of Liver Injury. II. Mechanisms of neutrophil-induced liver cell injury during hepatic ischemia-reperfusion and other acute inflammatory conditions. Am J Physiol Gastrointest Liver Physiol 2006; 290:G1083.
  160. Castillo T, Koop DR, Kamimura S, et al. Role of cytochrome P-450 2E1 in ethanol-, carbon tetrachloride- and iron-dependent microsomal lipid peroxidation. Hepatology 1992; 16:992.
  161. Chitturi S, Farrell GC. Etiopathogenesis of nonalcoholic steatohepatitis. Semin Liver Dis 2001; 21:27.
  162. De Minicis S, Brenner DA. NOX in liver fibrosis. Arch Biochem Biophys 2007; 462:266.
  163. Sancho P, Mainez J, Crosas-Molist E, et al. NADPH oxidase NOX4 mediates stellate cell activation and hepatocyte cell death during liver fibrosis development. PLoS One 2012; 7:e45285.
  164. Lin W, Wu G, Li S, et al. HIV and HCV cooperatively promote hepatic fibrogenesis via induction of reactive oxygen species and NFkappaB. J Biol Chem 2011; 286:2665.
  165. Lin W, Tsai WL, Shao RX, et al. Hepatitis C virus regulates transforming growth factor beta1 production through the generation of reactive oxygen species in a nuclear factor kappaB-dependent manner. Gastroenterology 2010; 138:2509.
  166. Nussler AK, Di Silvio M, Billiar TR, et al. Stimulation of the nitric oxide synthase pathway in human hepatocytes by cytokines and endotoxin. J Exp Med 1992; 176:261.
  167. Venkatraman A, Shiva S, Wigley A, et al. The role of iNOS in alcohol-dependent hepatotoxicity and mitochondrial dysfunction in mice. Hepatology 2004; 40:565.
  168. Paik YH, Schwabe RF, Bataller R, et al. Toll-like receptor 4 mediates inflammatory signaling by bacterial lipopolysaccharide in human hepatic stellate cells. Hepatology 2003; 37:1043.
  169. Wang FP, Li L, Li J, et al. High mobility group box-1 promotes the proliferation and migration of hepatic stellate cells via TLR4-dependent signal pathways of PI3K/Akt and JNK. PLoS One 2013; 8:e64373.
  170. Seo YS, Kwon JH, Yaqoob U, et al. HMGB1 recruits hepatic stellate cells and liver endothelial cells to sites of ethanol-induced parenchymal cell injury. Am J Physiol Gastrointest Liver Physiol 2013; 305:G838.
  171. Seki E, De Minicis S, Osterreicher CH, et al. TLR4 enhances TGF-beta signaling and hepatic fibrosis. Nat Med 2007; 13:1324.
  172. Huang H, Shiffman ML, Friedman S, et al. A 7 gene signature identifies the risk of developing cirrhosis in patients with chronic hepatitis C. Hepatology 2007; 46:297.
  173. Roh YS, Seki E. Toll-like receptors in alcoholic liver disease, non-alcoholic steatohepatitis and carcinogenesis. J Gastroenterol Hepatol 2013; 28 Suppl 1:38.
  174. Miura K, Yang L, van Rooijen N, et al. Toll-like receptor 2 and palmitic acid cooperatively contribute to the development of nonalcoholic steatohepatitis through inflammasome activation in mice. Hepatology 2013; 57:577.
  175. Clément S, Pascarella S, Conzelmann S, et al. The hepatitis C virus core protein indirectly induces alpha-smooth muscle actin expression in hepatic stellate cells via interleukin-8. J Hepatol 2010; 52:635.
  176. Bataller R, Paik YH, Lindquist JN, et al. Hepatitis C virus core and nonstructural proteins induce fibrogenic effects in hepatic stellate cells. Gastroenterology 2004; 126:529.
  177. Schulze-Krebs A, Preimel D, Popov Y, et al. Hepatitis C virus-replicating hepatocytes induce fibrogenic activation of hepatic stellate cells. Gastroenterology 2005; 129:246.
  178. Jin Z, Sun R, Wei H, et al. Accelerated liver fibrosis in hepatitis B virus transgenic mice: involvement of natural killer T cells. Hepatology 2011; 53:219.
  179. Wang S, Friedman SL. Hepatic fibrosis: A convergent response to liver injury that is reversible. J Hepatol 2020; 73:210.
  180. Rippe RA, Brenner DA. From quiescence to activation: Gene regulation in hepatic stellate cells. Gastroenterology 2004; 127:1260.
  181. Mann J, Oakley F, Akiboye F, et al. Regulation of myofibroblast transdifferentiation by DNA methylation and MeCP2: implications for wound healing and fibrogenesis. Cell Death Differ 2007; 14:275.
  182. Tsukamoto H, She H, Hazra S, et al. Anti-adipogenic regulation underlies hepatic stellate cell transdifferentiation. J Gastroenterol Hepatol 2006; 21 Suppl 3:S102.
  183. Gieling RG, Elsharkawy AM, Caamaño JH, et al. The c-Rel subunit of nuclear factor-kappaB regulates murine liver inflammation, wound-healing, and hepatocyte proliferation. Hepatology 2010; 51:922.
  184. Ding N, Yu RT, Subramaniam N, et al. A vitamin D receptor/SMAD genomic circuit gates hepatic fibrotic response. Cell 2013; 153:601.
  185. Beaven SW, Matveyenko A, Wroblewski K, et al. Reciprocal regulation of hepatic and adipose lipogenesis by liver X receptors in obesity and insulin resistance. Cell Metab 2013; 18:106.
  186. Morán-Salvador E, Titos E, Rius B, et al. Cell-specific PPARγ deficiency establishes anti-inflammatory and anti-fibrogenic properties for this nuclear receptor in non-parenchymal liver cells. J Hepatol 2013; 59:1045.
  187. Hernández-Gea V, Ghiassi-Nejad Z, Rozenfeld R, et al. Autophagy releases lipid that promotes fibrogenesis by activated hepatic stellate cells in mice and in human tissues. Gastroenterology 2012; 142:938.
  188. Hernández-Gea V, Friedman SL. Autophagy fuels tissue fibrogenesis. Autophagy 2012; 8:849.
  189. Hernández-Gea V, Hilscher M, Rozenfeld R, et al. Endoplasmic reticulum stress induces fibrogenic activity in hepatic stellate cells through autophagy. J Hepatol 2013; 59:98.
  190. Oakley F, Meso M, Iredale JP, et al. Inhibition of inhibitor of kappaB kinases stimulates hepatic stellate cell apoptosis and accelerated recovery from rat liver fibrosis. Gastroenterology 2005; 128:108.
  191. Mann J, Chu DC, Maxwell A, et al. MeCP2 controls an epigenetic pathway that promotes myofibroblast transdifferentiation and fibrosis. Gastroenterology 2010; 138:705.
  192. Moran-Salvador E, Garcia-Macia M, Sivaharan A, et al. Fibrogenic Activity of MECP2 Is Regulated by Phosphorylation in Hepatic Stellate Cells. Gastroenterology 2019; 157:1398.
  193. Stefanovic B, Hellerbrand C, Holcik M, et al. Posttranscriptional regulation of collagen alpha1(I) mRNA in hepatic stellate cells. Mol Cell Biol 1997; 17:5201.
  194. Stefanovic B, Hellerbrand C, Brenner DA. Regulatory role of the conserved stem-loop structure at the 5' end of collagen alpha1(I) mRNA. Mol Cell Biol 1999; 19:4334.
  195. Pinzani M. PDGF and signal transduction in hepatic stellate cells. Front Biosci 2002; 7:d1720.
  196. Wong L, Yamasaki G, Johnson RJ, Friedman SL. Induction of beta-platelet-derived growth factor receptor in rat hepatic lipocytes during cellular activation in vivo and in culture. J Clin Invest 1994; 94:1563.
  197. Cao S, Yaqoob U, Das A, et al. Neuropilin-1 promotes cirrhosis of the rodent and human liver by enhancing PDGF/TGF-beta signaling in hepatic stellate cells. J Clin Invest 2010; 120:2379.
  198. Pinzani M, Marra F. Cytokine receptors and signaling in hepatic stellate cells. Semin Liver Dis 2001; 21:397.
  199. Di Sario A, Bendia E, Taffetani S, et al. Selective Na+/H+ exchange inhibition by cariporide reduces liver fibrosis in the rat. Hepatology 2003; 37:256.
  200. Campbell JS, Hughes SD, Gilbertson DG, et al. Platelet-derived growth factor C induces liver fibrosis, steatosis, and hepatocellular carcinoma. Proc Natl Acad Sci U S A 2005; 102:3389.
  201. Yoshiji H, Kuriyama S, Yoshii J, et al. Vascular endothelial growth factor and receptor interaction is a prerequisite for murine hepatic fibrogenesis. Gut 2003; 52:1347.
  202. Marra F, Grandaliano G, Valente AJ, Abboud HE. Thrombin stimulates proliferation of liver fat-storing cells and expression of monocyte chemotactic protein-1: potential role in liver injury. Hepatology 1995; 22:780.
  203. Marra F, DeFranco R, Grappone C, et al. Expression of the thrombin receptor in human liver: up-regulation during acute and chronic injury. Hepatology 1998; 27:462.
  204. Steiling H, Mühlbauer M, Bataille F, et al. Activated hepatic stellate cells express keratinocyte growth factor in chronic liver disease. Am J Pathol 2004; 165:1233.
  205. Kim SY, Cho BH, Kim UH. CD38-mediated Ca2+ signaling contributes to angiotensin II-induced activation of hepatic stellate cells: attenuation of hepatic fibrosis by CD38 ablation. J Biol Chem 2010; 285:576.
  206. Wallace MC, Friedman SL, Mann DA. Emerging and disease-specific mechanisms of hepatic stellate cell activation. Semin Liver Dis 2015; 35:107.
  207. Czochra P, Klopcic B, Meyer E, et al. Liver fibrosis induced by hepatic overexpression of PDGF-B in transgenic mice. J Hepatol 2006; 45:419.
  208. Borkham-Kamphorst E, van Roeyen CR, Ostendorf T, et al. Pro-fibrogenic potential of PDGF-D in liver fibrosis. J Hepatol 2007; 46:1064.
  209. Maher JJ. Interactions between hepatic stellate cells and the immune system. Semin Liver Dis 2001; 21:417.
  210. Marra F, DeFranco R, Grappone C, et al. Expression of monocyte chemotactic protein-1 precedes monocyte recruitment in a rat model of acute liver injury, and is modulated by vitamin E. J Investig Med 1999; 47:66.
  211. Ikeda K, Wakahara T, Wang YQ, et al. In vitro migratory potential of rat quiescent hepatic stellate cells and its augmentation by cell activation. Hepatology 1999; 29:1760.
  212. Kinnman N, Hultcrantz R, Barbu V, et al. PDGF-mediated chemoattraction of hepatic stellate cells by bile duct segments in cholestatic liver injury. Lab Invest 2000; 80:697.
  213. Marra F, Romanelli RG, Giannini C, et al. Monocyte chemotactic protein-1 as a chemoattractant for human hepatic stellate cells. Hepatology 1999; 29:140.
  214. Bonacchi A, Romagnani P, Romanelli RG, et al. Signal transduction by the chemokine receptor CXCR3: activation of Ras/ERK, Src, and phosphatidylinositol 3-kinase/Akt controls cell migration and proliferation in human vascular pericytes. J Biol Chem 2001; 276:9945.
  215. Hashmi AZ, Hakim W, Kruglov EA, et al. Adenosine inhibits cytosolic calcium signals and chemotaxis in hepatic stellate cells. Am J Physiol Gastrointest Liver Physiol 2007; 292:G395.
  216. Friedman SL. Transcriptional regulation of stellate cell activation. J Gastroenterol Hepatol 2006; 21 Suppl 3:S79.
  217. Ruhl CE, Everhart JE. Coffee and tea consumption are associated with a lower incidence of chronic liver disease in the United States. Gastroenterology 2005; 129:1928.
  218. George J, Wang SS, Sevcsik AM, et al. Transforming growth factor-beta initiates wound repair in rat liver through induction of the EIIIA-fibronectin splice isoform. Am J Pathol 2000; 156:115.
  219. Bataller R, Schwabe RF, Choi YH, et al. NADPH oxidase signal transduces angiotensin II in hepatic stellate cells and is critical in hepatic fibrosis. J Clin Invest 2003; 112:1383.
  220. Stefanovic B, Schnabl B, Brenner DA. Inhibition of collagen alpha 1(I) expression by the 5' stem-loop as a molecular decoy. J Biol Chem 2002; 277:18229.
  221. Shimada H, Staten NR, Rajagopalan LE. TGF-β1 mediated activation of Rho kinase induces TGF-β2 and endothelin-1 expression in human hepatic stellate cells. J Hepatol 2011; 54:521.
  222. Tahashi Y, Matsuzaki K, Date M, et al. Differential regulation of TGF-beta signal in hepatic stellate cells between acute and chronic rat liver injury. Hepatology 2002; 35:49.
  223. Svegliati Baroni G, D'Ambrosio L, Ferretti G, et al. Fibrogenic effect of oxidative stress on rat hepatic stellate cells. Hepatology 1998; 27:720.
  224. Whalen R, Rockey DC, Friedman SL, Boyer TD. Activation of rat hepatic stellate cells leads to loss of glutathione S-transferases and their enzymatic activity against products of oxidative stress. Hepatology 1999; 30:927.
  225. Rigamonti C, Andorno S, Maduli E, et al. Iron, hepatic stellate cells and fibrosis in chronic hepatitis C. Eur J Clin Invest 2002; 32 Suppl 1:28.
  226. Paradis V, Dargere D, Bonvoust F, et al. Effects and regulation of connective tissue growth factor on hepatic stellate cells. Lab Invest 2002; 82:767.
  227. Rachfal AW, Brigstock DR. Connective tissue growth factor (CTGF/CCN2) in hepatic fibrosis. Hepatol Res 2003; 26:1.
  228. Gao R, Brigstock DR. Connective tissue growth factor (CCN2) induces adhesion of rat activated hepatic stellate cells by binding of its C-terminal domain to integrin alpha(v)beta(3) and heparan sulfate proteoglycan. J Biol Chem 2004; 279:8848.
  229. Chen L, Charrier AL, Leask A, et al. Ethanol-stimulated differentiated functions of human or mouse hepatic stellate cells are mediated by connective tissue growth factor. J Hepatol 2011; 55:399.
  230. Paradis V, Perlemuter G, Bonvoust F, et al. High glucose and hyperinsulinemia stimulate connective tissue growth factor expression: a potential mechanism involved in progression to fibrosis in nonalcoholic steatohepatitis. Hepatology 2001; 34:738.
  231. Weng HL, Ciuclan L, Liu Y, et al. Profibrogenic transforming growth factor-beta/activin receptor-like kinase 5 signaling via connective tissue growth factor expression in hepatocytes. Hepatology 2007; 46:1257.
  232. Gressner OA, Lahme B, Demirci I, et al. Differential effects of TGF-beta on connective tissue growth factor (CTGF/CCN2) expression in hepatic stellate cells and hepatocytes. J Hepatol 2007; 47:699.
  233. Roskams T, Cassiman D, De Vos R, Libbrecht L. Neuroregulation of the neuroendocrine compartment of the liver. Anat Rec A Discov Mol Cell Evol Biol 2004; 280:910.
  234. Mallat A, Teixeira-Clerc F, Deveaux V, Lotersztajn S. Cannabinoid receptors as new targets of antifibrosing strategies during chronic liver diseases. Expert Opin Ther Targets 2007; 11:403.
  235. Bátkai S, Járai Z, Wagner JA, et al. Endocannabinoids acting at vascular CB1 receptors mediate the vasodilated state in advanced liver cirrhosis. Nat Med 2001; 7:827.
  236. Teixeira-Clerc F, Julien B, Grenard P, et al. CB1 cannabinoid receptor antagonism: a new strategy for the treatment of liver fibrosis. Nat Med 2006; 12:671.
  237. Zvibel I, Atias D, Phillips A, et al. Thyroid hormones induce activation of rat hepatic stellate cells through increased expression of p75 neurotrophin receptor and direct activation of Rho. Lab Invest 2010; 90:674.
  238. Kendall TJ, Hennedige S, Aucott RL, et al. p75 Neurotrophin receptor signaling regulates hepatic myofibroblast proliferation and apoptosis in recovery from rodent liver fibrosis. Hepatology 2009; 49:901.
  239. Passino MA, Adams RA, Sikorski SL, Akassoglou K. Regulation of hepatic stellate cell differentiation by the neurotrophin receptor p75NTR. Science 2007; 315:1853.
  240. Muñoz-Luque J, Ros J, Fernández-Varo G, et al. Regression of fibrosis after chronic stimulation of cannabinoid CB2 receptor in cirrhotic rats. J Pharmacol Exp Ther 2008; 324:475.
  241. Ebrahimkhani MR, Kiani S, Oakley F, et al. Naltrexone, an opioid receptor antagonist, attenuates liver fibrosis in bile duct ligated rats. Gut 2006; 55:1606.
  242. De Minicis S, Candelaresi C, Marzioni M, et al. Role of endogenous opioids in modulating HSC activity in vitro and liver fibrosis in vivo. Gut 2008; 57:352.
  243. Oben JA, Roskams T, Yang S, et al. Hepatic fibrogenesis requires sympathetic neurotransmitters. Gut 2004; 53:438.
  244. Ikejima K, Okumura K, Kon K, et al. Role of adipocytokines in hepatic fibrogenesis. J Gastroenterol Hepatol 2007; 22 Suppl 1:S87.
  245. Choi SS, Syn WK, Karaca GF, et al. Leptin promotes the myofibroblastic phenotype in hepatic stellate cells by activating the hedgehog pathway. J Biol Chem 2010; 285:36551.
  246. Marra F, Bertolani C. Adipokines in liver diseases. Hepatology 2009; 50:957.
  247. Savvidou S, Hytiroglou P, Orfanou-Koumerkeridou H, et al. Low serum adiponectin levels are predictive of advanced hepatic fibrosis in patients with NAFLD. J Clin Gastroenterol 2009; 43:765.
  248. Musso G, Gambino R, Biroli G, et al. Hypoadiponectinemia predicts the severity of hepatic fibrosis and pancreatic Beta-cell dysfunction in nondiabetic nonobese patients with nonalcoholic steatohepatitis. Am J Gastroenterol 2005; 100:2438.
  249. Moreno M, Chaves JF, Sancho-Bru P, et al. Ghrelin attenuates hepatocellular injury and liver fibrogenesis in rodents and influences fibrosis progression in humans. Hepatology 2010; 51:974.
  250. Xie G, Choi SS, Syn WK, et al. Hedgehog signalling regulates liver sinusoidal endothelial cell capillarisation. Gut 2013; 62:299.
  251. Michelotti GA, Xie G, Swiderska M, et al. Smoothened is a master regulator of adult liver repair. J Clin Invest 2013; 123:2380.
  252. Omenetti A, Choi S, Michelotti G, Diehl AM. Hedgehog signaling in the liver. J Hepatol 2011; 54:366.
  253. Rockey DC. Hepatic blood flow regulation by stellate cells in normal and injured liver. Semin Liver Dis 2001; 21:337.
  254. Housset C, Rockey DC, Bissell DM. Endothelin receptors in rat liver: lipocytes as a contractile target for endothelin 1. Proc Natl Acad Sci U S A 1993; 90:9266.
  255. Li J, Kuruba R, Wilson A, et al. Inhibition of endothelin-1-mediated contraction of hepatic stellate cells by FXR ligand. PLoS One 2010; 5:e13955.
  256. Reinehr RM, Kubitz R, Peters-Regehr T, et al. Activation of rat hepatic stellate cells in culture is associated with increased sensitivity to endothelin 1. Hepatology 1998; 28:1566.
  257. Suematsu M, Goda N, Sano T, et al. Carbon monoxide: an endogenous modulator of sinusoidal tone in the perfused rat liver. J Clin Invest 1995; 96:2431.
  258. Pirazzi C, Valenti L, Motta BM, et al. PNPLA3 has retinyl-palmitate lipase activity in human hepatic stellate cells. Hum Mol Genet 2014; 23:4077.
  259. Tomita K, Teratani T, Suzuki T, et al. Free cholesterol accumulation in hepatic stellate cells: mechanism of liver fibrosis aggravation in nonalcoholic steatohepatitis in mice. Hepatology 2014; 59:154.
  260. Ohata M, Lin M, Satre M, Tsukamoto H. Diminished retinoic acid signaling in hepatic stellate cells in cholestatic liver fibrosis. Am J Physiol 1997; 272:G589.
  261. Miyahara T, Schrum L, Rippe R, et al. Peroxisome proliferator-activated receptors and hepatic stellate cell activation. J Biol Chem 2000; 275:35715.
  262. Marra F, Efsen E, Romanelli RG, et al. Ligands of peroxisome proliferator-activated receptor gamma modulate profibrogenic and proinflammatory actions in hepatic stellate cells. Gastroenterology 2000; 119:466.
  263. Lee TF, Mak KM, Rackovsky O, et al. Downregulation of hepatic stellate cell activation by retinol and palmitate mediated by adipose differentiation-related protein (ADRP). J Cell Physiol 2010; 223:648.
  264. Beilfuss A, Sowa JP, Sydor S, et al. Vitamin D counteracts fibrogenic TGF-β signalling in human hepatic stellate cells both receptor-dependently and independently. Gut 2015; 64:791.
  265. Marra F, Pinzani M. Role of hepatic stellate cells in the pathogenesis of portal hypertension. Nefrologia 2002; 22 Suppl 5:34.
  266. Bonacchi A, Petrai I, Defranco RM, et al. The chemokine CCL21 modulates lymphocyte recruitment and fibrosis in chronic hepatitis C. Gastroenterology 2003; 125:1060.
  267. Schwabe RF, Bataller R, Brenner DA. Human hepatic stellate cells express CCR5 and RANTES to induce proliferation and migration. Am J Physiol Gastrointest Liver Physiol 2003; 285:G949.
  268. Wasmuth HE, Lammert F, Zaldivar MM, et al. Antifibrotic effects of CXCL9 and its receptor CXCR3 in livers of mice and humans. Gastroenterology 2009; 137:309.
  269. Sahin H, Trautwein C, Wasmuth HE. Functional role of chemokines in liver disease models. Nat Rev Gastroenterol Hepatol 2010; 7:682.
  270. Aoyama T, Inokuchi S, Brenner DA, Seki E. CX3CL1-CX3CR1 interaction prevents carbon tetrachloride-induced liver inflammation and fibrosis in mice. Hepatology 2010; 52:1390.
  271. Zaldivar MM, Pauels K, von Hundelshausen P, et al. CXC chemokine ligand 4 (Cxcl4) is a platelet-derived mediator of experimental liver fibrosis. Hepatology 2010; 51:1345.
  272. Berres ML, Koenen RR, Rueland A, et al. Antagonism of the chemokine Ccl5 ameliorates experimental liver fibrosis in mice. J Clin Invest 2010; 120:4129.
  273. Brun P, Castagliuolo I, Pinzani M, et al. Exposure to bacterial cell wall products triggers an inflammatory phenotype in hepatic stellate cells. Am J Physiol Gastrointest Liver Physiol 2005; 289:G571.
  274. Pradere JP, Troeger JS, Dapito DH, et al. Toll-like receptor 4 and hepatic fibrogenesis. Semin Liver Dis 2010; 30:232.
  275. Viñas O, Bataller R, Sancho-Bru P, et al. Human hepatic stellate cells show features of antigen-presenting cells and stimulate lymphocyte proliferation. Hepatology 2003; 38:919.
  276. Kobayashi S, Seki S, Kawada N, et al. Apoptosis of T cells in the hepatic fibrotic tissue of the rat: a possible inducing role of hepatic myofibroblast-like cells. Cell Tissue Res 2003; 311:353.
  277. Zimmermann HW, Seidler S, Nattermann J, et al. Functional contribution of elevated circulating and hepatic non-classical CD14CD16 monocytes to inflammation and human liver fibrosis. PLoS One 2010; 5:e11049.
  278. Kordes C, Sawitza I, Müller-Marbach A, et al. CD133+ hepatic stellate cells are progenitor cells. Biochem Biophys Res Commun 2007; 352:410.
  279. Pintilie DG, Shupe TD, Oh SH, et al. Hepatic stellate cells' involvement in progenitor-mediated liver regeneration. Lab Invest 2010; 90:1199.
  280. Wang Y, Yao HL, Cui CB, et al. Paracrine signals from mesenchymal cell populations govern the expansion and differentiation of human hepatic stem cells to adult liver fates. Hepatology 2010; 52:1443.
  281. Lafdil F, Chobert MN, Couchie D, et al. Induction of Gas6 protein in CCl4-induced rat liver injury and anti-apoptotic effect on hepatic stellate cells. Hepatology 2006; 44:228.
  282. Greenbaum LE, Wells RG. The role of stem cells in liver repair and fibrosis. Int J Biochem Cell Biol 2011; 43:222.
  283. Kisseleva T, Brenner DA. Inactivation of myofibroblasts during regression of liver fibrosis. Cell Cycle 2013; 12:381.
  284. Troeger JS, Mederacke I, Gwak GY, et al. Deactivation of hepatic stellate cells during liver fibrosis resolution in mice. Gastroenterology 2012; 143:1073.
  285. Friedman SL. Fibrogenic cell reversion underlies fibrosis regression in liver. Proc Natl Acad Sci U S A 2012; 109:9230.
  286. Nakano Y, Kamiya A, Sumiyoshi H, et al. A Deactivation Factor of Fibrogenic Hepatic Stellate Cells Induces Regression of Liver Fibrosis in Mice. Hepatology 2020; 71:1437.
  287. Krizhanovsky V, Yon M, Dickins RA, et al. Senescence of activated stellate cells limits liver fibrosis. Cell 2008; 134:657.
  288. Lujambio A, Akkari L, Simon J, et al. Non-cell-autonomous tumor suppression by p53. Cell 2013; 153:449.
  289. Yoshimoto S, Loo TM, Atarashi K, et al. Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome. Nature 2013; 499:97.
  290. Kong X, Feng D, Wang H, et al. Interleukin-22 induces hepatic stellate cell senescence and restricts liver fibrosis in mice. Hepatology 2012; 56:1150.
  291. Baboota RK, Rawshani A, Bonnet L, et al. BMP4 and Gremlin 1 regulate hepatic cell senescence during clinical progression of NAFLD/NASH. Nat Metab 2022; 4:1007.
Topic 1233 Version 21.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟