You are on page 1of 16

Advanced Drug Delivery Reviews 121 (2017) 27–42

Contents lists available at ScienceDirect

Advanced Drug Delivery Reviews

journal homepage: www.elsevier.com/locate/addr

Hepatic stellate cells as key target in liver fibrosis☆


Takaaki Higashi a,b, Scott L. Friedman a, Yujin Hoshida a,⁎
a
Division of Liver Diseases, Department of Medicine, Liver Cancer Program, Tisch Cancer Institute, Division of Liver Diseases, Department of Medicine, Graduate School of Biomedical Sciences, Icahn
School of Medicine at Mount Sinai, New York, USA
b
Department of Gastroenterological Surgery, Graduate School of Medical Science, Kumamoto University, Kumamoto, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Progressive liver fibrosis, induced by chronic viral and metabolic disorders, leads to more than one million deaths
Received 23 January 2017 annually via development of cirrhosis, although no antifibrotic therapy has been approved to date.
Received in revised form 21 March 2017 Transdifferentiation (or “activation”) of hepatic stellate cells is the major cellular source of matrix protein-secret-
Accepted 9 May 2017
ing myofibroblasts, the major driver of liver fibrogenesis. Paracrine signals from injured epithelial cells, fibrotic
Available online 12 May 2017
tissue microenvironment, immune and systemic metabolic dysregulation, enteric dysbiosis, and hepatitis viral
Keywords:
products can directly or indirectly induce stellate cell activation. Dysregulated intracellular signaling, epigenetic
Myofibroblast changes, and cellular stress response represent candidate targets to deactivate stellate cells by inducing reversion
Cirrhosis to inactivated state, cellular senescence, apoptosis, and/or clearance by immune cells. Cell type- and target-
Hepatitis specific pharmacological intervention to therapeutically induce the deactivation will enable more effective and
Alcoholic liver disease less toxic precision antifibrotic therapies.
Non-alcoholic fatty liver disease © 2017 Elsevier B.V. All rights reserved.
Non-alcoholic steatohepatitis

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2. Cellular sources of matrix-producing myofibroblasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1. Myofibroblasts: driver of tissue fibrogenesis in multiple organs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2. Hepatic stellate cells (HSCs): precursor of myofibroblasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3. Resident mesenchymal cells as major source of myofibroblasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4. Other potential cellular sources of myofibroblasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3. Mechanisms of HSC activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1. Extracellular events that promote HSC activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.1. Epithelial cell injury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.2. Altered extracellular matrix (ECM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.3. Immune regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.4. Other cell types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.5. Metabolic dysregulation and enteric dysbiosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.6. Chronic infection of hepatitis virus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

Abbreviations: 5-HT(2B), 5-hydroxytryptamine 2B receptor; ACE, angiotensin-converting enzyme; AMPK, AMP-activated protein kinase; ARB, AT1R blocker; ASK1, apoptosis signal-
regulating kinase 1; AT1R, angiotensin II type 1 receptor; ATX, autotaxin; BDL, bile duct ligation; CB, cannabinoid; CCL, C-C motif chemokine ligand; CCl4, carbon tetrachloride; CCR, C-C
motif chemokine receptor; CpG, cytosine-phosphoguanine dinucleotide; CVCR, C-X-C motif chemokine receptor; CXCL, C-X-C motif chemokine ligand; DAMP, damage-associated
molecular pattern; DKK1, Dickkopf-1; DNMT, DNA methyltransferase; ECM, extracellular matrix; EGF, epidermal growth factor; ER, endoplasmic reticulum; ERK, extracellular signal-
regulated kinase; EZH2, enhancer of zeste 2 polycomb repressive complex 2 subunit; FGF, fibroblast growth factor; FXR, farnesoid-X-receptor; HBV, hepatitis B virus; HCV, hepatitis C
virus; HGF, hepatocyte growth factor; Hh, hedgehog; HMGB1, high mobility group protein B1; HSC, hepatic stellate cell; IL, interleukin; iNOS, inducible nitric oxide synthase; JNK, c-
Jun N-terminal kinase; LPA, lysophosphatidic acid; LRAT, lecithin-retinol acyltransferase; LSEC, liver sinusoidal endothelial cell; LY6C, lymphocyte antigen 6 complex; MeCP2, methyl-
CpG binding protein 2; MMP, matrix metallopeptidase; NAFLD, non-alcoholic fatty liver disease; NASH, non-alcoholic steatohepatitis; NF-kB, nuclear factor kBNK, natural killer; NO,
nitric oxide; OCA, obeticholic acid; PAMP, pathogen-associated molecular pattern; PDGF, platelet-derived growth factor; PPAR, peroxisome proliferator-activated receptor; PRR,
pattern recognition receptor; RAR, retinoic acid receptor; ROS, reactive oxygen species; RXR, retinoid X receptor; SREBP, sterol regulatory element-binding protein; STAT, signal
transducer and activator of transcription; TGF, transforming growth factor; Th, helper T; TLR, toll-like receptor; TNF, tumor necrosis factor; TWEAK, tumor necrosis factor-like weak
inducer of apoptosis; TZD, thiazolidinedione; UPR, unfolded protein response; UTR, untranslated region; VAP-1, vascular adhesion protein 1; VDR, vitamin D receptor; VEGF, vascular
endothelial growth factor.
☆ This review is part of the Advanced Drug Delivery Reviews theme issue on "Fibroblasts and extracellular matrix: Targeting and therapeutic tools in fibrosis and cancer".
⁎ Corresponding author at: Division of Liver Diseases, Department of Medicine, Liver Cancer Program, Tisch Cancer Institute, Graduate School of Biomedical Sciences, Icahn School of
Medicine at Mount Sinai, 1470 Madison Ave, Box 1123, New York, NY 10029, USA.
E-mail address: yujin.hoshida@mssm.edu (Y. Hoshida).

http://dx.doi.org/10.1016/j.addr.2017.05.007
0169-409X/© 2017 Elsevier B.V. All rights reserved.
28 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

3.2. Molecular dysregulation in activated HSCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31


3.2.1. Membrane receptor signaling pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2.2. Nuclear receptor signaling pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.3. Transcription factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.4. Epigenetic transcriptional dysregulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.5. Dysregulation of cellular homeostasis and stress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4. Deactivation and elimination of fibrogenic HSCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1. Induction of non-fibrogenic state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.1. Reversion, transdifferentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.2. Senescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.2. Induction of HSC death or killing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5. Summary and future direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

1. Introduction clinical studies have indicated that liver fibrosis and even cirrhosis can
be regressed by therapeutic intervention aimed at the primary disease eti-
Chronic tissue injury leads to a sustained scarring response that grad- ology [5]. In this review, we summarize recent findings relevant to direct
ually disrupts normal cellular functional units and eventually causes fail- therapeutic targeting of the major fibrogenic cell type in the liver, the ac-
ure in multiple epithelial organs such as liver, lung, and kidney, which is tivated hepatic stellate cell (HSC), as an antifibrotic strategy.
estimated to account for one-third of deaths worldwide [1]. Progressive
liver fibrosis can be caused by chronic infection of hepatitis B virus 2. Cellular sources of matrix-producing myofibroblasts
(HBV) or hepatitis C virus (HCV), alcohol abuse, non-alcoholic fatty liver
disease (NAFLD)/non-alcoholic steatohepatitis (NASH), and other rela- 2.1. Myofibroblasts: driver of tissue fibrogenesis in multiple organs
tively rare conditions such as autoimmune hepatitis, hemochromatosis,
Wilson's disease, and primary/secondary biliary cholangitis. Cirrhosis is In chronic fibro-proliferative diseases, affecting multiple organs such
the terminal stage of progressive liver fibrosis, which is estimated to affect as lung, kidney, and liver, presence of myofibroblasts is a key common
1% to 2% of global population and results in over 1 million deaths annually feature [6]. The myofibroblast is a fibroblast-like cell with contractile
worldwide [2,3]. properties, which is derived typically from cells of mesenchymal lineage
Lethal complications of cirrhosis include functional liver failure, portal via transdifferentiation, often referred to as “activation”. Proliferating
hypertension-induced variceal bleeding, ascites, and hepatic encephalop- myofibroblasts are the key source of excess extracellular matrix (ECM)
athy, systemic bacterial infection, and liver cancer, especially hepatocellu- molecules such as collagen type I and III as well as other proteins that con-
lar carcinoma (HCC) [3]. Annual direct and indirect costs for the care of stitute pathologic fibrous tissues [7]. Paracrine factors such as platelet-de-
cirrhosis exceed $12 billion in the U.S. alone [4]. Although approved ther- rived growth factor (PDGF), transforming growth factor β (TGFβ), and
apies directly targeting and reversing advanced fibrosis are still lacking, connective tissue growth factor (CTGF), together with other growth

Fig. 1. Extracellular regulation of HSC activation. HSC activation is promoted (sharp arrow) or inhibited (blocked arrow) by liver resident cells, ECM, and circulating cells via paracrine
factors. Red and blue font colors indicate positive and negative regulators of HSC activation, respectively. Pharmacological intervention to each candidate target is shown in parenthesis.
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 29

factors and cytokines/chemokines, activate cellular signaling pathways of that 82% to 96% of myofibroblasts were derived from HSCs in mice
cell proliferation, migration, ECM protein secretion, and contractility that treated with carbon tetrachloride (CCl4), bile duct ligation (BDL), 3,5-
promote transdifferentiation of precursor mesenchymal cells (e.g., diethoxycarbonyl-1,4-dihydro-collidin diet, and Mdr2-knockout mice
pericytes and resident fibroblasts) into fibrogenic myofibroblasts and [28]. Another study, labeling myofibroblasts using Col1a1 promoter in
maintain their activated state [7]. PDGF receptor β (PDGFRβ) and other mice, reported that HSCs were the major source of myofibroblasts
transmembrane proteins such as integrins are induced on the surface of (N87%) in chemical injury (CCl4), whereas portal fibroblasts were
myofibroblasts and precursors to form positive feedback loops for their the major source (N 70%) in an early stage of cholestatic injury (BDL)
perpetuated activation. [29]. These studies collectively suggest that the liver resident mesen-
chymal cells, particularly HSCs, are the major source of fibrogenic
2.2. Hepatic stellate cells (HSCs): precursor of myofibroblasts myofibroblasts, although these rodent-derived, single gene marker-
based findings need to be verified in humans. Furthermore, it is still
HSCs are resident mesenchymal cells that retain features of resident fi- unknown whether cell(s) of origin is/are associated with response to
broblasts (embedded in normal stromal matrix) and pericytes (attached antifibrotic therapies.
to endothelial cells of capillaries), and comprise approximately one-
third of nonparenchymal cells and 15% of total resident cells in normal 2.4. Other potential cellular sources of myofibroblasts
human liver [7]. HSCs reside in a virtual subendothelial space between
the basolateral surface of hepatocytes and the anti-luminal side of fenes- Although relative contribution is plausibly minor or negligible, other
trated sinusoidal endothelial cell layer (space of Disse) filled with thin cell types in addition to HSCs and portal fibroblasts have been proposed
permeable connective tissue, where exchange of biomolecules between as alternative sources of myofibroblasts. Collagen-producing fibrocytes,
portal blood flow from gastrointestinal tract and hepatocytes occurs. In- distinct from HSCs, are recruited from bone marrow in BDL mice [30,31].
tercellular communication via soluble mediators and cytokines/ Mesenchymal stem cells are also suggested as another bone marrow-
chemokines can occur between neighboring cell types such as hepato- derived source of myofibroblasts [32]. Wt1-expressing mesothelial
cytes, biliary epithelial cells, hepatic progenitor cells (oval cells), Kupffer cells (approximately 15% of the cell population) may be a source of
cells (liver resident macrophages), bone marrow-derived macrophages, myofibroblasts near the liver surface via mesothelial–mesenchymal
liver sinusoidal endothelial cells (LSECs), infiltrating immune cells, and transition (MMT) in CCl4 mice [33]. Epithelial-mesenchymal transition
nerve cells [7,8]. In fibrogenic liver, quiescent HSCs transdifferentiate (EMT) of liver parenchymal cells, i.e., hepatocytes and cholangiocytes,
into proliferative, migratory, and contractile myofibroblasts, manifesting is less likely the source of myofibroblasts, although the parenchymal
pro-fibrogenic transcriptional and secretory properties (so called “cell ac- cells could express some of the mesenchymal genes without ECM pro-
tivation”), and secrete ECM molecules that accumulate and form scar tis- tein secretion [32,34–36]. Hedgehog signaling contributes to mesenchy-
sue in the space of Disse that leads to sinusoidal capillarization mal-to-epithelial transition (MET) of a subpopulation of HSC-derived
characterized by loss of endothelial fenestrations [7] (Fig. 1). Vitamin A myofibroblasts into hepatic progenitor cells in the process of liver
(retinoid) storage in cytoplasmic droplets is a unique characteristic fea- regeneration in mice [37].
ture of quiescent HSCs, which is gradually lost during the process of
transdifferentiation, although its causal relationship with HSC activation 3. Mechanisms of HSC activation
is still uncertain [7,9]. Endothelin-1 (ET-1), a potent vasoconstrictor, is se-
creted by activated HSCs, and promotes cell proliferation, fibrogenesis, 3.1. Extracellular events that promote HSC activation
and contraction, which is plausibly linked to portal hypertension in cir-
rhosis [10]. Several cellular events in injured/inflamed hepatic tissue microenvi-
A number HSC-specific marker genes and proteins, including glial ronment as well as extrahepatic factors are known to directly induce
fibrillary acidic protein (GFAP), nerve growth factor receptor (p75), HSC activation in etiology-specific or independent manner [38] (Fig. 1).
desmin, lecithin-retinol acyltransferase (LRAT), integrin αvβ3, mannose
6-phosphate/insulin-like growth factor II receptor (M6P/IGF-IIR), colla- 3.1.1. Epithelial cell injury
gen type I, collagen type VI receptor (CVIR), α-smooth muscle actin Progressive fibrotic liver diseases are accompanied with chronic dam-
(αSMA), PDGFRβ, vimentin, and cytoglobin, have been exploited for age to parenchymal epithelial cells, i.e., hepatocytes and cholangiocytes
their histological detection, genetic targeting, cell fate tracing, therapeutic (biliary epithelial cells). According to the liver disease etiology (viral in-
targeting, and imaging [7,11–14]. In addition, by using genome-wide fection, hepatic toxins, metabolic disorders, and autoimmune diseases),
transcriptome profiling, HSC-specific 122-gene and 194-gene signatures, various modes of cell death (e.g., apoptosis, necrosis, necroptosis, and
associated with poorer patient prognoses, have been identified [15,16]. autophagic cell death), trigger cell death responses [39]. Immune cell-
A widely used experimental culture model is primary human and ro- mediated killing of hepatocytes is a common theme in HBV- and HCV-
dent HSCs, isolated by discontinuous gradient ultracentrifugation with or related liver fibrosis, which involves pro-fibrogenic inflammatory
without cell surface marker-based sorting, that undergo spontaneous chemokines and their receptors such as C-C motif chemokine ligand 2
fibrogenic activation when grown on a hard plastic dish [17,18]. Although (CCL2), CCL21, interleukin-8 (IL-8), IL-17, IL-22, C-X-C motif chemokine
certain distinct differences in transcriptomic dysregulation have been ligand 9 (CXCL9), CXCL10, CXCL11, and C-X-C motif chemokine receptor
noted between in vitro and in vivo settings, co-culture with other cell 1 (CXCR1) [8]. Hepatic nuclear factor κB (NF-κB)-inducing kinase (NIK)
types such as Kupffer cells could partially restore more physiological mo- activates bone marrow-derived macrophages, increases hepatic injury,
lecular dysregulation [19]. Immortalized cell lines of human (LX-2 and and promotes HSC activation and fibrogenesis [40]. The dead or dying ep-
TWNT-4), mouse (JS-1, GRX, and CoI-GFP), and rat (HSC-T6 and CFSC) ithelial cells as well as leukocytes phagocytosing the cells release inflam-
HSCs, and rat portal myofibroblasts (RGF and RGF-N2) were established, matory mediators, damage-associated molecular patterns (DAMPs) or
and have been used for mechanistic analysis of antifibrotic targets and danger signals, which initiate and perpetuate a non-infectious “sterile”
therapies [20–27]. inflammatory response. Among such mediators, tumor necrosis factor
(TNF), IL-6, IL-1β, reactive oxygen species (ROS), hedgehog ligands, and
2.3. Resident mesenchymal cells as major source of myofibroblasts nucleotides can contribute to initiation of HSC activation [8,41,42]. Such
host biomolecules from injured hepatocytes, “alarmins”, include high
Rodent-based studies have elucidated several potential sources of mobility group protein B1 (HMGB1) and IL-33 [43,44]. IL-33 induces IL-
myofibroblasts. Cell fate tracing, genetically labeling HSCs by using 13 production in liver resident innate lymphoid cells type II, enhances
LRAT, involved in the cytoplasmic lipid droplet formation, showed TGFβ signaling through IL-4Rα and signal transducer and activator of
30 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

transcription 6 (STAT6) in HSCs, and promotes fibrosis [45]. Loss of the Th17 cells secret IL-17A and IL-22, and are increased in the liver and
ROS-generating enzymes, NADPH oxidase 1 (NOX1), NOX2, or NOX4, at- circulation in chronic liver diseases [67]. IL-17 receptor is expressed on
tenuates hepatocyte injury and liver fibrosis in CCl4 and BDL mice [46,47]. the surface of macrophages and HSCs, and the signaling induces pro-in-
A NOX1/NOX4 inhibitor, GKT137831, reduces hepatic apoptosis and fi- flammatory cytokines such as TNF, IL-1β, IL-6, and IL-17A, and collagen
brosis in CCl4 and BDL mice [48]. Specific modes of hepatocyte and 1 expression in HSCs by activating STAT3 signaling in mice [68]. IL-22
cholangiocyte death may be monitored by circulating cytokeratins as bio- transgenic mice are protected from CCl4-induced fibrosis through
markers [49]. Apoptosis inhibitors such as a pan-caspase inhibitor, increased HSC senescence by binding to its receptors, IL-10R2 and IL-
emricasan (IDN-6556), aiming to reduce hepatocyte death, are under 22R1, up-regulating STAT3 and increasing p53 expression [69]. Recombi-
evaluation in clinical trials in post-transplant HCV and NASH patients nant IL-22 also attenuates HSC activation and fibrosis [70]. On the other
(ClinicalTrials.gov: NCT02138253, NCT02686762, NCT02960204; EN- hand, IL-22 may have adverse effect in HBV infection [71]. Intrahepatic
CORE trials). Apoptosis signal-regulating kinase 1 (ASK1) is a mitogen- IL-8 producing Foxp3 + CD4+ regulatory T cells (Tregs) activate HSCs
activated protein kinase (MAP3K) that activates c-Jun N-terminal kinase and promote fibrosis in chronic hepatitis C [72]. γδ T cells restrict liver fi-
(JNK) and p38 MAPK in response to various cellular stresses. A phase 2 brosis by inducing HSC apoptosis in CCR6-dependent manner [73].
trial of ASK1 inhibitor, selonsertib (GS-4997), resulted in more than 1 B cells, which account for up to half of lymphocytes in the liver,
stage of liver fibrosis improvement in 43% of enrolled NASH patients contribute to fibrogenesis in CCl4 mice [74]. HSC-mediated MyD88-
(NCT02466516). dependent innate B cell activation promotes hepatic fibrosis [75].
Natural killer (NK) cells kill senescent HSCs and produce interferon γ
3.1.2. Altered extracellular matrix (ECM) (IFNγ) that induces apoptosis and cell cycle arrest in HSCs [76]. IL-15 in-
Progressive deposition of ECM proteins in the space of Disse leads to creases NK cells and support their killing of HSCs [77]. Regulatory CD4+
increased density and stiffness of ECM. In addition, matrix composition T cells can suppress NK cells, and thereby support survival of activated
shifts from collagen type IV, heparan sulfate proteoglycan, and laminin HSCs [78]. Natural killer T (NKT) cells also secrete IFNγ and kill activated
to fibrillar collagen type I and III. These changes serve as mechanical HSCs, whereas a subset of NKT cells produce IL-4, IL-13, osteopontin,
stimuli to activate HSCs at least partially via integrin signaling path- and hedgehog ligands, and promote HSC activation and liver fibrosis
ways, forming positive feedback loops [50,51]. LX-2 cells cultured on ac- via CXCR6-CXCL16 axis in chronic liver diseases including NASH [79–
rylamide gels with 12 kPa pressure, mimicking increased tissue stiffness 81].
in fibrotic liver, show higher collagen I expression via RHOA up-regula-
tion, which is suppressed by c-Src [52]. Primary HSCs cultured in 3.1.4. Other cell types
hydrogels stiffened or softening in situ over time revealed time-course Liver sinusoidal endothelial cells (LSECs) maintain HSC quiescence,
dynamic transcriptional changes during the process of culture activa- thus inhibiting intrahepatic vasoconstriction and fibrosis development
tion and regression, respectively [53,54]. Expanded ECM also serve as [82]. In pathological conditions, LSECs become capillarized with lost fen-
reservoir by binding growth factors such as PDGF, hepatocyte growth estrations, lose the hepatoprotective property, and promote angiogene-
factor (HGF), fibroblast growth factor (FGF), epidermal growth factor sis and vasoconstriction. A soluble guanylate cyclase activator, BAY 60-
(EGF), and vascular endothelial growth factor (VEGF), which promote 2770, leads to reversal of the capillarization, restores HSC quiescence
HSC proliferation [6,55]. Reduced matricellular protein, CCN3/NOV, en- without directly affecting HSCs, and reduces fibrosis in thioacetamide
hances fibrogenic gene expression in primary HSCs and CFSC cells [56]. (TAA)-treated cirrhotic rats [83].
Lysyl oxidase-like-2 (LOXL2) is a matrix enzyme expressed by HSC, Close proximity between hepatic progenitor cells and HSCs is often
which catalyzes crosslinking of collagens and elastins, and its inhibition observed at the site of ductular reaction, a pathological feature accom-
by monoclonal antibody (AB0023) reduces liver and lung fibrosis in panying fibrotic liver diseases, which may have functional implication
experimental models [57]. A humanized monoclonal LOXL2 anti- in HSC activation and/or liver regeneration possibly via contribution of
body, simtuzumab (GS-6624), has been tested in phase 2 trials for HSCs to progenitor cell pool [84,85].
patients with HCV (NCT01707472), primary sclerosing cholangitis Osteopontin, a multifunctional secreted phosphoprotein, induces
(NCT01672853), and fibrotic (NCT01672866) or cirrhotic collagen I expression in HSC by affecting biliary epithelial cells and pro-
(NCT01672879) NASH. moting ductular reaction [86,87]. Platelets are a source of pro-fibrogenic
cytokines and growth factors such as PDGF, TGFβ, and EGF, and they
3.1.3. Immune regulation can also suppress collagen expression via HGF-Met signaling pathway
Macrophages represent a heterogeneous population (10–15% of liver [7,88].
cells), which can have both pro- and anti-fibrogenic effects through
paracrine regulation of HSC activation [41,58]. Macrophage depletion 3.1.5. Metabolic dysregulation and enteric dysbiosis
in murine models has revealed their pro-fibrogenic role [59]. Polarized A so-called Western diet can alter the intestinal microbiome (intesti-
and plastic activation of macrophages is traditionally classified into clas- nal dysbiosis), and lead to release and increased exposure to pathogen-
sic M1 and alternative M2 activation associated with helper T (Th)1 associated molecular patterns (PAMPs) that can activate HSCs through
(TNFα, IL-1, IL-12, and inducible nitric oxide synthase [iNOS]) and Th2 toll-like receptors (TLRs) [89]. A high-fat diet increases the proportion
(IL-4, IL-10, and IL-13) responses, respectively [60]. p53 depletion in of intestinal Gram-negative endotoxin-producing bacteria, resulting in
HSCs in mice leads to M2 activation that promotes HCC proliferation higher rates of bacterial translocation, and accelerated fibrogenesis
by affecting the tumor microenvironment [61]. Lymphocyte antigen 6 by enhancing HSC activation in CCl4 and BDL mice [90]. Transplantation
complex (LY6C)hi monocyte-derived macrophages are recruited in C-C of the fibrogenic microbiome to control diet-fed mice exacerbates liver
motif chemokine receptor 2 (CCR2)-dependent manner and secrete damage and fibrosis, suggesting the causal role of microbiome in
pro-fibrogenic mediators such as TGFβ, PDGF, CCL2, CCL3, CCL5, CCL7, fibrogenic HSC activation. In contrast, commensal microbiota is hepato-
and CCL8 [8,62]. Galectin-3 is a macrophage-derived lectin, which pro- protective and prevents liver fibrosis by mitigating HSC activation in
motes HSC activation and can be pharmacologically inhibited by GR- mice [91]. Lipotoxicity caused by free intracellular cholesterol accumula-
MD-02, currently in a phase 2 trial in NASH patients (NCT02462967) tion sensitizes HSCs to TGFβ-induced activation in NASH mouse model
[63,64]. In contrast, LY6Clo macrophages represent a fibrolytic subset [92].
that expands during fibrosis regression [62]. C-X3-C motif chemokine
receptor 1 (CX3CR1) induces differentiation to LY6Clo cells, which pro- 3.1.6. Chronic infection of hepatitis virus
mote apoptosis of activated HSCs and secrete matrix metallopeptidase Chronic infection by hepatotropic viruses, HBV and HCV, has been
12 (MMP-12) and MMP-13 to promote ECM degradation [65,66]. one of the major risk factors for fibrotic liver diseases worldwide [93].
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 31

Fig. 2. Cellular signaling pathways involved in HSC activation and deactivation. A variety of cell surface receptor-, nuclear receptor-, and intracellular signal transmitter-mediated
mechanisms of HSC activation has been identified. Dysregulation of cellular homeostasis and epigenetic modifications also regulate HSC activation. HSC deactivation can be induced
via specific targets. Red and blue font colors indicate positive and negative regulators of HSC activation, respectively. Black font color indicates involvement in both pro- and anti-
fibrogenic functions depending on biological context or undetermined role in HSC activation. Pharmacological intervention to each candidate target is shown in parenthesis.

Given that population-level viral eradication will not be achieved in the 3.2.1. Membrane receptor signaling pathways
next several decades despite the development of highly effective
antivirals, directly targeting fibrosis pathways irrespective of viral clear- 3.2.1.1. TGFβ. Activated HSCs secrete latent TGFβ in response to liver in-
ance will remain a viable antifibrotic strategy. Viral genes and proteins jury, which forms an autocrine positive feedback loop driving
can directly or indirectly promote HSC activation. HBV e antigen directly fibrogenesis via SMAD2/SMAD3, whereas SMAD7 suppresses the acti-
induces activation and proliferation of rat HSCs in vitro via the TGFβ vation [7,107,108]. Apoptotic body-engulfing macrophages produce
pathway, and core and X proteins similarly activate human LX-2 cells TGFβ and activate HSCs [30,109–111]. Liver-targeted TGFβ1 expression
via PDGFβ signaling [94,95]. Although HCV cannot infect HSCs, viral induces fibrosis and chemical carcinogenesis in mice [112,113].
core and non-structural proteins directly induce inflammatory and Therapeutic strategies, e.g., blocking circulating TGFβ1, antagonizing
pro-fibrogenic pathways in HSCs [96,97]. HCV core protein may its receptors, or blocking its activation at the cell surface, have been
promote EMT of infected hepatocytes through TGFβ signaling [98]. challenging due to adverse effects of systemic TGFβ antagonism, includ-
HCV core also increases ECM protein, thrombospondin-1 (TSP-1) and ing inflammation and cancer [6,7,114]. Knockdown of transient
transcription factors, which increase bioavailability and expression of receptor potential melastatin7 (TRPM7) attenuates TGFβ1-induced
TGFβ [99,100]. Elevated expression of IL-34 and macrophage colony- expression of myofibroblast markers, increases the ratio of MMPs/tissue
stimulating factor (M-CSF) from HCV-infected hepatocytes stimulates inhibitor of metalloproteinases (TIMPs), and decrease SMAD2/SMAD3-
maturation of peripheral monocytes into macrophages, which contrib- associated collagen production [115,116]. Ca(2 +)/calmodulin-
utes to HSC activation by enhancing TGFβ and PDGFβ signaling, dependent protein kinase II (CaMKII) mediates TGFβ-induced HSC pro-
blocking antifibrotic NK cell-derived IFNγ, and inducing MMP-1 [101]. liferation, but not collagen production [117]. CUG-binding protein 1
Human immunodeficiency virus (HIV) co-infection is clinically known (CUGBP1), induced by TGFβ, binds the 3′- untranslated region (UTR)
to accelerate fibrosis progression in HCV-infected individuals, and it of antifibrogenic IFNγ mRNA and promotes its decay [118].
can be slowed down by antiretroviral therapies [102,103]. HIV can di-
rectly infect HSCs, and viral envelop protein, gp120, promotes collagen
I expression via CXCR4 [104,105]. Exposure to HIV induces TGFβ, ROS, 3.2.1.2. PDGF. Cell proliferative/inflammatory cytokines, PDGF, FGF,
and NF-κB in a hepatocyte/LX-2 co-culture reporter cell model [106]. TGFα, p38 MAPK, and JNK, converge on MAPK/extracellular signal-
regulated kinase (ERK) pathway activation [6,119]. PDGF is a potent
chemoattractant for HSCs and myofibroblasts [7]. PDGFβ receptor
3.2. Molecular dysregulation in activated HSCs (PDGFRβ) expression is induced during initiation of HSC activation,
and enhances inflammatory and fibrogenic responses to chemical
Functional studies have elucidated complex dysregulation of molec- injury via ERK, AKT, and NF-κB pathways, without affecting
ular pathways in HSCs/myofibroblasts for their perpetuated activation hepatocarcinogenesis in mice [120,121]. Small molecule tyrosine kinase
through multiple signaling pathways and other mechanisms as summa- inhibitors, including imatinib, sorafenib, nilotinib, and sunitinib,
rized below (Fig. 2). suppress proliferative and fibrogenic properties of activated HSCs in
32 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

organotypic ex vivo culture of fibrotic human liver tissues [122–125]. PPARγ repression and promote liver regeneration [152]. A selective in-
CD248/endosialin is a PDGF downstream regulator of HSC proliferation hibitor of the cAMP-response element-binding protein (CBP)/β-catenin
during chronic liver injury [126]. A tyrosine phosphatase SHP-1 inhibits interaction, PRI-724, inhibits HSC activation and reduces liver fibrosis in
proliferation of activated HSCs via PDGF signaling [127]. Knockdown of mice [153]. In contrast, β-catenin-dependent canonical Wnt activation
protein tyrosine phosphatase receptor type O (PTPRO) can suppress is required to maintain quiescent state of HSCs in vitro [154]. Roof
PDGFβ-induced HSC proliferation and activation by reducing phosphor- plate-specific spondin (RSPO) proteins, potent Wnt pathway agonists,
ylation of ERK and AKT in mice [128]. PDGF-C transgenic mice exhibit suppressed HSC activation, which is impaired by DKK1 [155]. In activat-
HSC activation and develop hepatic steatosis, fibrosis, and cancer ed rat HSCs, non-canonical Wnt pathway activation dominates and con-
despite limited hepatocyte injury and lack of extensive fibrosis [129]. fers HSC survival, accompanied with Wnt5a over-expression [156]. A
Pigment epithelium-derived factor (PEDF)-derived 34-mer peptide natural Wnt5a inhibitor, secreted frizzled-related protein 5 (Sfrp5), de-
(Asp44-Asn77) down-regulates PDGF receptor expression and blocks creases HSC activation and liver fibrosis [157]. These findings collective-
HSC activation [130]. ly suggest that involvement of Wnt pathway in HSC activation is highly
complex and biological context-dependent.
3.2.1.3. CTGF. CTGF/CCN2 is a central fibrogenic effector of TGFβ
pathway, and its knockout suppresses fibrogenesis in CCl4-treated 3.2.1.9. Hedgehog (Hh) signaling. Hh ligand expression from hepatocytes
mice [131]. A human monoclonal antibody to CTGF, FG-3019, was tested induces HSC activation and promotes liver fibrosis and hepatocar-
in HBV fibrosis patients, although its clinical utility as antifibrotic thera- cinogenesis in a rodent model [158,159]. Smoothened (Smo), an obligate
py was not clear due to prominent effect of anti-HBV drug, entecavir intermediate in Hh signaling pathway, is required not only for fibrogenic
(NCT01217632) [132,133]. activation of HSCs, but also for epithelial regeneration in mice [37]. Hh
signals transduced via activated Smo are transmitted to Gli transcription
3.2.1.4. EGF. EGF receptor is overexpressed and phosphorylated in acti- factors that induce target gene expression [160]. Gli1-positive
vated HSCs, and its pharmacological inhibition reduces hepatic fibrosis perivascular mesenchymal cells give rise to 39% of αSMA-positive
and HCC nodules in rodent models, although EGF receptor expression myofibroblasts after tissue injury [161]. Forskolin, a hedgehog signaling
is not restricted to HSCs [134]. Genetic knockout of EGF receptor in inhibitor, attenuates CCl4-induced liver fibrosis in rats [162]. Pharmaco-
macrophages reduces HCC development in mice [135]. logical inhibition of Hh signaling by vismodegib (GDC-0449) attenuates
liver fibrosis in BDL and NASH models [163,164]. Inhibition of Notch or
3.2.1.5. Rho kinases (ROCKs). ROCKs and family proteins, including Rac1, Hh signaling in myofibroblasts leads to mesenchymal-to-epithelial like
promote reactive oxygen species (ROS) production and enhance con- transdifferentiation with reduced fibrosis in mice [165].
tractility of HSCs, which can be inhibited by kinase inhibitors and
other measures such as adenosine [136–139]. HSC-specific inhibition 3.2.1.10. Neurochemical receptors. The cannabinoid system is involved in
of Rho-kinase reduces portal pressure in cirrhotic rats without major neuro/immune-modulatory function, and has two G protein-coupled
toxicity [140]. receptors, CB1 (pro-fibrogenic) and CB2 (hepatoprotective) [166]. Sys-
temic antagonism of CB1 can induce serious psychiatric adverse effects
3.2.1.6. Integrins. Integrins are heterodimeric transmembrane proteins as seen during trials of rimonabant. Newer generation antagonists with
composed of α/β subunits that sense cell adhesion, transmit signals improved target specificity and limited central nervous system penetra-
from ECM, and promote fibrogenesis by activating latent TGFβ [6,7]. tion such as AM4113 and AM6545 have been developed based on the
Genetic deletion of integrin αv as well as its pharmacological blockade crystal structure of CB1 [167,168]. A germline genetic variant in the
by a small molecule, CWHM 12, in activated HSCs reduces CCl4-induced CNR2 gene (encoding CB2), CB2-63QQ, was found to be associated
hepatic fibrosis, whereas no such effect is observed for integrins β3, β5, with more severe necroinflammation in patients with chronic hepatitis
β6, and β8 [51]. Integrin β1 induces PAK proteins and YAP signaling in C [169]. A selective CB 2 agonist, JWH-133, reduces fibrosis by inducing
myofibroblasts and promote liver fibrosis [141]. Integrin α1β6 induces HSC quiescence/apoptosis and reducing IL-17 production by Th17 cells
ROS accumulation through Rac1-NADPH oxidase 1 enzyme complex [166,170]. Nutritional products such as curcumin (diferuloylmethane),
in HSCs [142]. a bioactive polyphenolic ingredient of turmeric, may have potential to
inhibit CB1 as well as virtually all cellular pathways involved in HSC ac-
3.2.1.7. Renin-angiotensin system. Angiotensin II, secreted by HSCs, binds tivation [171,172]. Antagonism of a serotonin receptor, 5-hydroxytryp-
to angiotensin II type 1 receptor (AT1R) and promotes liver fibrosis via tamine 2B receptor (5-HT2B), on HSCs attenuates fibrosis and enhances
JAK2, which is associated with portal hypertension [143,144]. Blocking hepatic regeneration [173]. Inhibition of 5-HT2A similarly suppresses
the renin–angiotensin system by angiotensin-converting enzyme (ACE) fibrogenic gene expression and induces apoptosis in HSCs [174].
inhibitors or angiotensin receptor blockers (ARBs) such as losartan may
be an effective antifibrotic therapy as suggested by experimental and clin- 3.2.1.11. Toll like receptors (TLRs). TLRs are pattern recognition receptors
ical studies [145,146]. The AT1R pathway has a synergistic activity with (PRRs) that detect microbial infection and exposure as PAMPs, including
cannabinoid signaling through receptor heterodimerisation during HSC bacterial protein products, transcripts, and genomic DNA [8]. Liver is
activation, suggesting that combination of ACE/ARB and CB1 receptor an- consistently exposed to PAMPs from gastrointestinal and oral microflo-
tagonist may be an option of antifibrotic therapy [147]. Swertiamarin ra, and it is known that plasma and portal level of endotoxin such as li-
(Swe), an extract of Gentiana manshurica Kitag, significantly inhibits an- popolysaccharide (LPS), a Gram-negative bacteria constituent, is
giotensin II-induced HSC proliferation and activation via AT1R [148]. elevated in cirrhotic patients due to bacterial dysbiosis and overgrowth,
as well as gut leakiness [175]. The composition of the gut microbiome
3.2.1.8. Wnt/β-catenin signaling. The Wnt pathway regulates a variety of from cirrhosis patients was highly comparable across different popula-
cellular processes such as development, differentiation, and prolifera- tions, suggesting its influence on pathogenesis of fibrotic liver diseases
tion, and can have both activating and inhibitory effects on HSCs via TLRs [176]. HSCs express TLR4 and TLR9, and TLR2 is induced by
[149]. siRNA-mediated β-catenin knockdown reduces collagen I and III TLR4, upon activation by pro-inflammatory and pro-fibrogenic path-
expression, inhibits cell proliferation, and induces apoptosis of HSCs in ways and cytokines including NF-κB, JNK, TGFβ, IL-6, IL-8, CCL2 (MCP-
vitro [150]. Wnt signaling inhibition by Dickkopf-1 (DKK1) abrogates 1), MIP-2, intercellular cell adhesion molecule 1 (ICAM1), vascular cell ad-
epigenetic repression and restores peroxisome proliferator-activated hesion molecule 1 (VCAM1), and E-selectin [177]. LPS-induced TLR4 acti-
receptor-γ (PPARγ) activity, and reduces fibrosis [151]. Conversely, vation downregulates Bambi (a TGFβ pseudoreceptor) and sensitizes
HSC-derived delta-like homolog 1 (DLK1) activates HSCs via epigenetic quiescent HSCs to TGFβ activation mediated by MyD88-NF-κB pathway,
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 33

and enhances liver fibrogenesis in CCl4 and BDL mice [178]. TLR4 activa- knockout as well as inhibition by BGB324 results in attenuated HSC ac-
tion is required for HCC promotion, but not initiation in mice [179]. tivation and reduced fibrosis in CCl4 mice.
TLR9 senses DNA from apoptotic hepatocytes to stop chemotaxis at the
site of dying hepatocytes and stimulate collagen synthesis [180]. Bacterial 3.2.2. Nuclear receptor signaling pathways
DNA activates TLR9 on resident and recruited macrophages to release IL-
1β, which activates HSCs and enhances lipid accumulation, insulin resis- 3.2.2.1. Farnesoid-X-receptor (FXR). FXR signaling, activated by bile acids,
tance, and injury in hepatocytes, and collectively promotes fibrogenesis enhances insulin sensitivity, and fatty acid beta-oxidation, and reduces
and hepatic inflammation [181]. TLR2 stimulates TNFα production from gluconeogenesis and lipogenesis in hepatocytes [89]. A synthetic lipo-
intestinal monocytes that activates the RhoA pathway in intestinal epithe- philic bile acid, obeticholic acid (OCA), has an FXR agonistic activity
lial cells to increase intestinal permeability and bacterial translocation to stronger than the chemically similar ursodeoxycholic acid, and reduced
the liver [182]. TLR2 deficiency leads to attenuated liver fibrosis in cho- liver fibrosis in non-cirrhotic NASH patients in a phase 2 trial (FLINT
line-deficient amino acid-defined (CDAA) diet-fed NASH mice [183]. trial) [203]. OCA is currently being evaluated in a phase 3 trial in fibrotic
Exosome-mediated TLR3 activation in HSCs promotes liver fibrosis by in- NASH patients (REGENERATE trial, NCT02548351). In addition, the fol-
ducing IL-17A production by γδ T cells in mice [184]. lowing agents targeting FXR signaling are being evaluated in clinical tri-
als: GS-9674, synthetic non-steroidal FXR agonist; INT-767, bile acid
3.2.1.12. Lysophospholipid pathway. Lysophosphatidic acid (LPA) is a analogue agonizing FXR and TGR5 (transmembrane G-protein bile
pleiotropic growth factor-like lysophospholipid, involved in lipid ho- acid receptor expressed on sinusoidal and bile duct epithelial cells)
meostasis. Hepatocyte-derived autotaxin (ATX) catalyzes production [89,204]. Although FXR expression and OCA response are limited in
of LPA that activates LPA receptor 1 (LPAR1) on HSCs, and increased HSCs, induction of FXR signaling and resulting suppression of sphingo-
ATX in various chronic liver diseases has been association with poorer sine-1-phosphate receptor 2 (S1PR2) signaling by dihydroartemisinine
prognosis and cancer risk, [185,186]. Pharmacological and/or genetic in- (DHA), an artemisinin derivative, restricts HSC contraction and
hibition of ATX and/or LPAR1 attenuates HSC activation and reduces fibrogenesis, and improves portal hypertension in rodent models
liver fibrosis and cancer in rodent models [185,186]. [205–207].

3.2.1.13. Adipokines. Adipokines, e.g., leptin and adiponectin, are soluble 3.2.2.2. PPARs. PPARs, consisting of α, β/δ, and γ isoforms that share
factors secreted by adipose tissue, which contribute to the pathogenesis common target DNA sequences but with distinct ligand selectivity and
of obesity- and metabolic disorder-related complications through their tissue distribution, serve as receptors for a wide range of fatty acids
pro- or anti-inflammatory properties [187]. Leptin and adiponectin and derivatives; they transcriptionally regulate various metabolic pro-
exert pro- and anti-fibrogenic function, respectively, via their receptors cesses to maintain lipid and energy homeostasis [208]. PPARγ has
on HSCs [188,189]. Leptin reduces expression of sterol regulatory ele- been extensively studied as a potential antifibrogenic target in HSCs
ment-binding protein-1c (SREBP-1c) and activates HSCs through involve- [209]. PPARγ suppresses angiogenic PDGFRβ signaling and TGFβ1 via
ment of β-catenin pathway [190]. Adiponectin attenuates liver fibrosis by the β-catenin pathway in HSCs [210,211]. Lipid-induced hepatocyte-de-
inducing nitric oxide (NO) production via adipoR2-AMP-activated protein rived extracellular vesicles regulate HSCs via miR-128-3p targeting
kinase (AMPK)-JNK/ErK1/2-NF-κB pathway and TIMP-1, and inhibits pro- PPARγ [212]. Histone H3K9 demethylase JMJD1A modulates HSCs acti-
liferation and migration of activated HSCs [191,192]. Adiponectin induces vation by epigenetic regulation of PPARγ [213]. Leptin inhibits PPARγ
expression of aquaglyceroporin (known to be down-regulated in associa- expression via GATA binding protein 2 in HSCs [214,215]. Adiponectin
tion with obesity) and inactivates HSCs [193]. Adiponectin-derived syn- regulates HSC activation via PPARγ-dependent and independent mech-
thetic short peptide, ADP355, suppresses focal adhesion kinase (FAK) anisms [216]. Curcumin regulates PPARγ coactivator-1α expression by
activity and TGFβ1/SMAD2 signaling and promotes AMPK and STAT3 sig- AMPK pathway in HSCs in vitro [217]. Activation of PPARγ/p53
naling, and inactivates HSCs in mice [194,195]. signaling is required for curcumin to induce HSC senescence [218].
Thiazolidinediones (TZDs), synthetic insulin sensitizing PPARγ agonists
3.2.1.14. Vascular adhesion protein 1 (VAP-1). VAP-1 (encoded by AOC3) such as pioglitazone and rosiglitazone, improved steatosis and lobular
is a membrane sialoglycoprotein expressed by LSECs that promotes inflammation but not fibrosis in non-cirrhotic NASH patients in a
lymphocyte recruitment [196]. VAP-1 is also expressed on activated phase 3 trial (PIVENS trial) [219]. Safety concerns about TZDs such as
HSCs and associated with hepatic inflammation and fibrosis in NASH the cardiovascular toxicity of rosiglitazone and weight gain due to redis-
patients and mouse models [197]. A selective VAP-1 inhibitor, PXS- tribution of body fat have led to suboptimal acceptance of these agents
4728A, was tested in a phase 1 study. as NASH treatments [89]. A dual PPARα/δ agonist, elafibranor (GFT-
505), improved histological NASH with no worsening of fibrosis
3.2.1.15. Tumor necrosis factor-like weak inducer of apoptosis (TWEAK). in non-cirrhotic NASH patients in a phase 2 trial, which may suggest in-
TWEAK (encoded by TNFSF12) is secreted from macrophages, monocytes, direct suppression of hepatocyte lipotoxicity-driven fibrogenesis
and T lymphocytes upon liver injury, and signals through its receptor, (GOLDEN-505 trial) [220]. A follow-up phase 3 trial is currently
FGF-inducible 14 (Fn14), to activate liver-resident progenitor cells, recruiting patients with fibrotic NASH patients (NCT02704403). A dual
while hepatocyte replication is impaired in chronic liver diseases [198]. PPARα/γ agonist, saroglitazar, is being evaluated in a phase 3 trial for
TWEAK also leads to liver fibrosis progression by directly promoting no worsening fibrosis in NASH patients (Indian trial registry, CTRI/
HSC proliferation. TWEAK, and by enhancing SIRT1 expression and de- 2015/10/006236). Novel selective agonists of PPARα (pemafibrate, K-
creasing p53 acetylation, which assumedly inhibits HSC senescence 877), PPARγ (INT-131), PPARδ (HPP-593), PPARα/γ (DSP-8658) and a
[199,200]. pan-PPAR agonist (IVA337) are under evaluation.

3.2.1.16. G protein-coupled receptor 91 (GPR91). Sirtuin 3 (SIRT3), 3.2.2.3. Liver X receptors (LXRs). LXRs are lipid-activated nuclear recep-
an NAD(+)-dependent protein deacetylase decreased in NAFLD, tors involved in cholesterol transport, lipogenesis, and anti-inflammato-
suppresses αSMA expression via GPR91 on HSCs in methionine- and ry signaling. LXR activation suppresses HSC activation in mice [221].
choline-deficient (MCD) diet-fed NAFLD mice [201]. Leptin-induced activation of p38 MAPK leads to inhibition of LXR activ-
ity and downregulation of SREBP-1c expression, and increases collagen
3.2.1.17. Axl/growth arrest specific 6 (Gas6). Axl is a receptor tyrosine type I expression in HSCs [222]. The therapeutic benefit of LXR agonism
kinase, activated by its ligand, Gas6. Serum Gas6 level is increased as al- may be limited due to induction of de novo lipogenesis by this mecha-
coholic liver disease or chronic hepatitis C progresses [202]. Axl nism [5].
34 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

3.2.2.4. Vitamin D receptor (VDR). In HSCs, TGFβ1 causes a redistribution is reduced in fibrotic liver, and its genetic knockout in HSCs promoted
of genome-wide VDR-binding sites (VDR cistrome) at SMAD3 ECM production and cell proliferation via MAPK pathway [217]. NR4A2
profibrotic target genes via chromatin remodeling, and VDR ligands re- over-expression suppresses ECM production and invasive ability, and in-
duce TGFβ/SMAD3 target gene expression and inhibit fibrosis [223]. duces cell cycle arrest and apoptosis in vitro. Moreover delivery by adeno-
p62/SQSTM1 suppresses fibrogenic and carcinogenic HSC activation by viral infection in dimethylnitrosamine (DMN)-treated fibrotic rats
promoting VDR signaling, which is abrogated by VDR knockout in significantly reduces liver fibrosis [244].
mice [224]. In human primary HSCs, vitamin D treatment similarly sup-
pressed TGFβ-induced fibrogenic transcription, but A1012G single nu- 3.2.3.5. Kruppel-like factors (KLFs). KLF6, a zinc finger transcription factor
cleotide polymorphism (G allele) within the VDR gene reduces VDR and tumor suppressor, directly represses collagen I and PDGFRβ expres-
expression and abolishes the vitamin D effect [225]. sion and increases apoptosis of HSCs, and KLF6 haploinsufficiency in-
creases fibrosis in rats [245]. Statins up-regulates KLF2 expression in
3.2.2.5. Retinoid receptors. Retinoic acid, an active metabolite of vitamin sinusoidal endothelial cells, and induces vasoprotective gene expres-
A, regulates various physiological processes such as energy metabolism sion, and quiescence of HSCs through a KLF2-nitric oxide-guanylate
in metabolically active cells through retinoic acid receptors (RARs) and cyclase-mediated paracrine mechanism [246].
retinoid X receptors (RXRs) that form homo/heterodimers [226].
Retinoid receptor signaling in hepatocytes is implicated in liver carcino- 3.2.3.6. Aryl hydrocarbon receptor (AhR). AhR is a ligand-activated tran-
genesis, and a synthetic acyclic retinoid, peretinoin, has been assessed scription factor involved in the regulation of xenobiotic-metabolizing
as a potential cancer chemoprevention measure [227,228]. In HSCs, enzymes such as cytochrome P450, and AhR-deficient mice develop
all-trans retinoic acid (ATRA) (an RAR ligand) moderately inhibits ex- liver fibrosis [247]. HSC activation, i.e., TGFβ activation, PPARγ reduc-
pression of procollagen I, III, IV, fibronectin, laminin, αSMA, TGFβ, and tion, and increased collagen secretion, induced by AhR deletion requires
IL-6, but with no effect on cell proliferation, whereas 9-cis (RXR ligand) vitamin A in mice [248]. 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)
moderately inhibits proliferation without affecting the fibrogenic gene inhibits lipid droplet storage, and enhances cell proliferation and
expression in rodents [229,230]. These receptors, also together with αSMA expression, but not collagen I expression, in the LX-2 human
PPARγ signaling, have additive inhibitory effects on HSC activation HSC line [249].
[231,232]. RXR can form heterodimer with VDR to transmit signal
from p62 in HSCs to promote fibrogenesis and inflammation, and sup- 3.2.3.7. Gα-interacting vesicle-associated protein (GIV). GIV/Girdin serves
port carcinogenesis [224]. as a cellular hub that transmits signals from multiple receptors such as
PDGF and TGFβ, promotes pro-fibrogenic signaling and inhibit anti-
3.2.2.6. Rev-Erb receptors. Rev-Erbs, consisting of Rev-Erbα and Rev- fibrogenic pathways [250].
Erbβ, are heme-sensing receptors involved in regulation of metabolism,
development, immunity, and circadian rhythm, and are potentially 3.2.3.8. Yes associated protein (YAP). YAP is a transcription co-activator,
druggable [233]. Rev-Erbα is up-regulated in both nucleus and cyto- which is downstream and suppressed by Hippo pathway. YAP also
plasm in activated HSCs in association with increased contractility and transmits signal from integrin β1 in myofibroblasts, and promotes fibrosis
fibrogenic gene expression, which can be antagonized by a synthetic [141]. YAP knockdown inhibits HSC activation, and pharmacological inhi-
ligand, SR6452 [234]. Another synthetic Rev.-Erbα ligand, SR9009, sim- bition of YAP suppresses fibrogenesis in mice [251]. YAP triggers accumu-
ilarly reduces fibrogenic gene expression and proliferation in HSCs, ac- lation of reactive-appearing ductular cells, increases myofibroblasts, and
companied by decreased autophagosome formation [235]. induces fibrosis in murine NASH model [252].

3.2.3. Transcription factors 3.2.3.9. Embryonic stem cell-expressed RAS (ERAS). ERAS is a RAS family
GTPase, which plays an important role in maintaining quiescent HSCs
3.2.3.1. Sex-determining region Y-box 9 (SOX9). SOX9 has been implicated with correlated activation of AKT, STAT3, mTORC2, and Hippo signaling
in control of liver progenitor cells and organ development [236]. A re- pathways and inactivation of FOXO1 and YAP [253].
cent study reported that leptin stimulates SOX9 expression, and SOX9
suppresses PPARγ expression by binding to its promoter region to stim- 3.2.4. Epigenetic transcriptional dysregulation
ulate HSC activation [237]. SOX9 induces osteopontin, and contributes Global epigenomic and local epigenetic transcriptional dysregula-
to liver fibrosis in human [238]. tion in HSC can occur as response or adaptation to environmental and/
or etiological stressors, and has been implicated in its fibrogenic
3.2.3.2. GATA binding protein 4 (GATA4). Conditional knockout or activation [254].
haploinsufficiency of a zinc finger transcription factor, GATA4, in hepatic
mesenchymal cells leads to fibrogenic HSC activation by directly regu- 3.2.4.1. Genomic DNA methylation, histone modification, and chromatin
lating an antifibrogenic transcription factor, Lhx2, and accelerates fibro- remodeling. Genome-wide characterization of DNA methylation (5-mC)
sis in mice, and GATA4 expression is reduced in human cirrhotic livers and hydroxymethylation (5-hmC) of cytosine-phosphoguanine dinucleo-
[239,240]. tides (CpGs) during HSC activation in rat and human has revealed modu-
lation of regulatory enzymes, e.g., induction of DNA methyltransferase
3.2.3.3. Myocardin-related transcription factor A (MRTF-A). MRTF-A (DNMT) 3a/b and suppression of ten-eleven translocation methylcytosine
induces histone H3K4 di/tri-methylation, recruits histone methyltrans- dioxygenase (TET) 2 and 3, accompanied by decreased levels of 5-hmC
ferase complex (COMPASS), and recruits and interacts with regulatory [255]. HSC activation can be suppressed by inhibitors of DNMT1 (e.g., 5-
epigenetic factors, including ASH2, WDR5, and SET1, and activates ex- Aza-2′-deoxycytidine) and enhancer of zeste 2 polycomb repressive com-
pression of fibrogenic genes, including TGFβ target genes, which can plex 2 subunit (EZH2), a polycomb group-protein complex catalyzing
be suppressed by estradiol [241,242]. trimethylation of histone 3 lysine 27 (H3K27me3) (e.g., 3-
deazaneplanocin), supporting their regulatory role in the activation
3.2.3.4. Nuclear receptor subfamily 4 group A member 1/2 (NR4A1/2). [256–258]. DNMT1-mediated DNA hypermethylation of phosphatase
NR4A1 and NR4A2 are orphan nuclear receptor family members that and tension homolog (PTEN) leads to HSC activation [259]. Methyl-CpG
act as transcription factors, regulating cell differentiation, proliferation binding protein 2 (MeCP2), released from translational repression by
and apoptosis. NR4A1 suppresses TGFβ target genes, and its small miR-132, induces HSC activation and liver fibrosis by binding promoter
molecule agonist, Csn-B, reduces fibrosis in mice [243]. NR4A2 expression of PPARγ gene and suppresses its expression [256]. In parallel, MeCP2
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 35

stimulates EZH2 expression and H3K27 methylation to form a repressive providing critical ATP substrates [280,281]. Inhibition of autophagy sup-
chromatin structure in the 3′ exons of PPARγ, which can be pharmacolog- presses HSC activation and proliferation [282]. Autophagy in HSCs can
ically inhibited. ASH1, a transcriptional repressor induced by MeCP2, be induced by oxidative stress and ER stress in response to chronic liver
binds regulatory regions of αSMA, collagen I, TIMP1, and TGFβ1 in activat- injury [283]. The unfolded protein response (UPR) is a highly conserved
ed HSCs, and depletion of ASH1 suppresses fibrogenic gene/protein ex- ER stress effector pathway transmitting signal through either of protein
pression [260]. Histone H3K9 demethylase JMJD1A suppresses HSC kinase R-like ER kinase (PERK) and inositol-requiring enzyme 1 (IRE1)/
activation and liver fibrosis via the H3K9me2 mark in the PPARγ gene X-box transcriptional factor (XBP1) pathways that link ER stress, UPR,
promoter [213]. Low-molecular-weight fibroblast growth factor 2 and fibrogenic HSC activation in distinct ways: the TGFβ pathway is stim-
(FGF2(lmw)) attenuates hepatic fibrosis by epigenetic down-regulation ulated in the PERK-mediated, but not IRE1/XBP1-mediated, UPR during
of Delta-like1 [261]. The bromodomain-containing protein 4 (BRD4), a HSC activation [284–286]. Advanced glycation end product (AGE)-in-
member of bromodomain and extraterminal (BET) proteins highly duced HSC activation requires autophagy in HCV-related fibrosis [287].
enriched at enhancers of pro-fibrogenic genes, can be pharmacologically Transient receptor potential vanilloid 4 inhibits apoptosis in rat HSC
inhibited by its inhibitor, JQ1 to suppress HSC activation [262]. The line, HSC-T6, through induction of autophagy [288]. Rev.-Erb-activating
PPARγ promoter methylation detected in plasma cell-free circulating synthetic ligand, SR9009, and TGFβ similarly affect autophagy, but differ-
DNA stratifies mild and severe fibrosis in patients with NAFLD and alco- entially regulate HSCs [235]. Rev.-Erbα and LXR, along with PPARγ, are
holic liver disease, although it was more correlated with methylation in critical to preserve adipogenic phenotype of HSCs [221,234]. Deletion of
hepatocytes than in myofibroblasts [263]. a collagen-specific molecular chaperone, heat shock protein 47 (Hsp47),
causes ER stress-mediated apoptosis of HSCs [289]. Vitamin A-coupled li-
3.2.4.2. Dietary factors. Ethanol induces an H3K4 methyltransferase, posome nanoparticles loaded with Hsp47-targeting siRNA, ND-L02-
lysine methyltransferase 2A (KMT2A), and transcriptionally activates s0201, resolves fibrosis in rodent models, and is now under clinical eval-
promoter of multiple ECM genes, including elastin [264]. Ethanol can uation in a phase 1b/2 trial (NCT02227459) [290]. Reticulon 4B (RTN4B,
also induce H3K9 acetylation in HSCs in time- and dose-dependent Nogo-B), associated with ER, is over-expressed in activated HSCs [291,
manner [265]. 292]. Hypoxia-inducible factor-1α (HIF1α) regulates autophagy to acti-
vate HSCs [293]. Dimethyl α-ketoglutarate reduces CCl4-induced liver fi-
3.2.4.3. Inheritable epigenetic adaptation. Ancestral history of liver brosis through inhibition of autophagy in HSCs [294]. Suppression of
damage is associated with epigenetic suppressive adaptation that autophagy in hepatocytes by antagonism of IL-17A inhibits liver fibrosis
leads to higher PPARγ expression, fewer myofibroblasts, and lower ex- by restoring the IL-10/STAT3 pathway in BDL and TAA-treated mice
pression of profibrogenic factors such as TGFβ in offspring rats [266]. [295]. Potentiation of HSC activation by extracellular ATP is dependent
on P2X7R-mediated NLRP3 inflammasome activation [296].
3.2.4.4. Non-coding RNAs. Recent studies have elucidated dysregulation of
non-coding RNAs, especially microRNAs (miRs), involved in HSC 3.2.5.2. Retinoids/vitamin A. The loss of cytoplasmic retinoids in activated
activation and inactivation by post-transcriptional regulation of mRNA HSCs is linked to retinyl ester hydrolysis that generates fatty acids as part
degradation and translation. High-throughput miRNA profiling and of the autophagic response, although a causal relationship between reti-
other technologies have identified pro-fibrogenic miRNAs over-expressed noid metabolism and HSC activation is not established [282]. Retinoids
in activated HSCs, including miR-9a-5p, miR-17-5p, miR-21, miR-27, miR- exacerbate rat liver fibrosis by activating latent TGFβ in HSCs [297]. Con-
31, miR-33a, miR-34a/c, miR-125, miR-126, miR-130a/b, miR-181b, miR- version of retinol to retinoic acid leads to TGFβ and collagen expression
214-5p, miR-195, miR-199a/b, miR-221, and miR-222, each of which in- via alcohol dehydrogenase 3, and retinol released from HSCs suppresses
duces proliferation, collagen secretion, and/or migration via a variety of NK cells and contributes to HSC survival [298]. Vitamin A is required in
pathways, and anti-fibrotic miRNAs, including miR-16, miR-19b, miR- AhR deficiency-mediated murine HSC activation [248]. Murine HSCs,
29, miR-30, miR-101, miR-122, miR-133a, miR-144, miR-146a, miR-150, lacking retinoid droplets due to genetic knockout of LRAT, can be activat-
miR-155, miR-192, miR-195, miR-335, miR-454, and miR-483, although ed, suggesting that retinoid droplets are not indispensable for activation
direction of change could depend on biological and/or experimental con- [299]. Vitamin A and insulin are required for the maintenance of HSC qui-
texts [267–269]. miR-378 was recently shown to limit HSC activation by escence [300]. A retinoic acid receptor β2 agonist reduces HSC activation
suppressing Gli3 [270]. miR-145 inhibits HSC activation and proliferation in NAFLD [301]. Perilipin 2, adipose TG lipase (ATGL), and comparative
by targeting zinc finger E-box-binding homeobox 2 (ZEB2), a key media- gene identification-58 (CGI-58) are lipid droplet-associated proteins in
tor of epithelial-to-mesenchymal transition, via Wnt/β-catenin pathway rat HSC line, HSC-T6 [302]. LXRs balance lipid stores in HSCs through
[271]. Interferon-induced miR-195 down-regulates cyclin E1 and sup- Rab18, a retinoid responsive lipid droplet protein [303]. Autophagy regu-
presses HSC proliferation [272]. miR-200a over-expression inhibits lates turnover of lipid droplets via ROS-dependent Rab25 activation in
αSMA expression and suppresses TGFβ-induced proliferation through HSC [304]. Alcohol dehydrogenases (ADHs) oxidize retinol to
β-catenin, SIRT1/Notch1, and NRF2/KEAP1 pathways [273–275]. miR- retinaldehyde, which is metabolized to retinoic acid. Loss of ADH3 en-
214, delivered to HSCs by exosome, binds 3′-UTR of CTGF gene, and sup- hances NK cell activation and IFNγ production, increases HSC apoptosis,
presses its expression [130]. The role of long non-coding RNAs (lncRNAs) and attenuates liver fibrosis in CCl4 and BDL mice [298]. Patatin like phos-
is less studied compared to miRNAs. lncRNA, H19, which can be silenced pholipase domain containing 3 (PNPLA3) is highly expressed in HSCs and
by MeCP2, represses expression of insulin-like growth factor 1 receptor hepatocytes, and hydrolyzes retinol palmitate to retinol and palmitic acid
(IGF1R) and suppresses HSC proliferation [276]. Circulating cell-free with influence from retinol availability, and its over-expression is associ-
miRNAs, e.g., miR-122, miR-138, miR-143, and miR-185, have been ated with reduced cytoplasmic lipid droplets. A single nucleotide poly-
assessed as potential non-invasive biomarkers of HSC activation, liver fi- morphism (SNP) in the PNPLA3 gene, I148M, is associated with its
brosis, and/or prognosis in patients mostly affected with HBV or HCV reduced expression, more advanced NAFLD fibrosis, and HCC, and may
[277–279]. have functional implication in HSC activation [305].

3.2.5. Dysregulation of cellular homeostasis and stress 3.2.5.3. Protein ubiquitination. Ubiquitination is a post-translational modi-
fication, ligating ubiquitin to a substrate protein to regulate its
3.2.5.1. Autophagy, endoplasmic reticulum (ER) stress. Intracellular energy degradation, subcellular localization, and interaction with other proteins.
homeostasis is dysregulated during the process of HSC activation to Ubiquitin C-terminal hydrolase 1 (UCHL1), a deubiquitinase, is up-regu-
meet the energy demand. Autophagy is a highly regulated stress response lated following HSC activation, and promote their proliferation in CCl4
pathway induced by chronic liver injury that fuels HSC activation by and BDL mice [306]. Small molecule UCHL1 inhibitor, LDN-57444,
36 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

attenuates PDGF-induced HSC proliferation by modulating phosphoryla- 5. Summary and future direction
tion of a cell cycle regulator, retinoblastoma protein (Rb).
As summarized in this review, comprehensive characterization of
3.2.5.4. Bone morphogenetic protein 6 (BMP6). BMP6 is an iron homeosta- molecular dysregulation of HSCs and their functional assessment in
sis regulator. BMP6 deficiency promotes hepatic inflammation and engineered animal models have elucidated multiple candidate thera-
fibrosis in MCD plus high-fat diet-fed mice, but not in other non-NASH peutic targets and strategies, many of which are currently under evalu-
models, and recombinant BMP6 protein inhibits HSC activation [307]. ation in patients with chronic liver diseases [324]. However, both
fibrogenesis and fibrosis regression involve highly heterogeneous and
4. Deactivation and elimination of fibrogenic HSCs evolving cell populations whose interactions have not yet fully charac-
terized. Single cell/molecule-resolution genome-wide molecular char-
Termination of chronic liver injury results in regression of liver fibro- acterization, combined with experimental perturbations and special
sis accompanied with resolution of cytokine-rich inflammatory tissue reconstruction may overcome the limitations of the current strategies
microenvironment as well as reduced or loss of activated HSCs [5]. that rely upon single gene/protein marker-based cell type determina-
These observations suggest that deactivating and/or reducing fibrogenic tion [325–327]. Newly emerging experimental systems such as multi-
HSCs could be an antifibrotic strategy irrespective of the cause of liver cell type and/or 3D culture will enable more physiological modeling of
injury. fibrogenic molecular dysregulation [328]. Animal models based on
chemical, surgical, and/or genetic interventions have been criticized
4.1. Induction of non-fibrogenic state for their physiological and/or clinical irrelevance [329]. However, unbi-
ased omics-based characterization could enable identification of models
4.1.1. Reversion, transdifferentiation that recapitulate specific molecular dysregulation of interest [186]. In
Genetic cell lineage tracing studies in mice have showed that activat- addition to these advancements, systematic strategies to translate the
ed HSCs can be reverted to a quiescent-like state upon withdrawal of experimental model-based findings to clinically meaningful long-term
causative agents, although the reverted cells are sensitized to reactiva- consequences in human fibrotic liver diseases will substantially increase
tion by exposure to fibrogenic stimuli [308,309]. More recent study the chance of success in the clinical evaluation of candidate antifibrotic
demonstrated that in vivo transcriptional reprogramming by ectopic therapies [186].
expression of FOXA3, GATA4, HNF1A, and HNF4A transdifferentiates acti-
vated HSCs into hepatocyte-like non-fibrogenic cells, namely induced
hepatocytes (iHeps), and reduces liver fibrosis [310]. Acknowledgements

4.1.2. Senescence This work is supported by Japan Society for the Promotion of Science,
Cellular senescence is a genetically controlled physiological mecha- program for advancing strategic international networks to accelerate
nism that restricts cell division within a finite proliferative capacity to the circulation of talented researchers (to T.H.), NIH/NIDDK DK099558,
prevent accumulation of genetically damaged cells. Senescence mediat- European Union ERC-2014-AdG-671231HEPCIR, Irma T. Hirschl Trust,
ed by the p53/p21 pathway limits proliferation of activated HSCs, and and US Department of Defense W81XWH-16-1-0363 (to Y.H.). Some
reduces liver fibrosis in mice [311,312]. Thus, induction of HSC senes- cell and organelle images in the figures are from Togo picture gallery,
cence might serve as an antifibrotic strategy. A CCN family matricellular Database Center for Life Science (DBCLS) (licensed under creative
protein, cysteine-rich protein 61 (CCN1/CYR61), attenuates TGFβ sig- commons 4.0).
naling and induces senescence and apoptosis of activated HSCs [142,
313,314]. Proteomic analysis has revealed that CYR61 and Wnt5a are References
among several matrix proteins secreted from a human HSC line, LX-2
[315]. A celecoxib derivative, OSU-03012, inhibits the proliferation and [1] D.C. Rockey, P.D. Bell, J.A. Hill, Fibrosis—a common pathway to organ injury and
failure, N. Engl. J. Med. 372 (2015) 1138–1149.
activation of HSCs by inducing senescence [316]. Soluble egg antigens [2] G.B.D.R.F. Collaborators, M.H. Forouzanfar, L. Alexander, H.R. Anderson, V.F.
(SEA) of Schistosoma japonicum induce senescence of activated HSCs Bachman, S. Biryukov, et al., Global, regional, and national comparative risk assess-
by activating FoxO3a/SKP2/p27 pathway [317]. HSC senescence can ment of 79 behavioural, environmental and occupational, and metabolic risks or
clusters of risks in 188 countries, 1990–2013: a systematic analysis for the Global
also be induced by phytochemicals such as curcumin by activating
Burden of Disease Study 2013, Lancet 386 (2015) 2287–2323.
PPARγ/p53 signaling [218]. RAR/RXR and PPARγ may play a role in [3] E.A. Tsochatzis, J. Bosch, A.K. Burroughs, Liver cirrhosis, Lancet 383 (2014)
inducing HSC senescence as antifibrotic therapy [318]. 1749–1761.
[4] P.S. Ge, B.A. Runyon, Treatment of patients with cirrhosis, N. Engl. J. Med. 375
(2016) 767–777.
4.2. Induction of HSC death or killing [5] Y.A. Lee, M.C. Wallace, S.L. Friedman, Pathobiology of liver fibrosis: a translational
success story, Gut 64 (2015) 830–841.
Activated HSCs are resistant to cell death stimuli by activating surviv- [6] S.L. Friedman, D. Sheppard, J.S. Duffield, S. Violette, Therapy for fibrotic diseases:
nearing the starting line, Sci. Transl. Med. 5 (2013) (167sr161).
al signals via the NF-κB pathway and inducing anti-apoptotic proteins [7] S.L. Friedman, Hepatic stellate cells: protean, multifunctional, and enigmatic cells
such as Bcl-2 [5]. Therapeutic induction of susceptibility to cell death sig- of the liver, Physiol. Rev. 88 (2008) 125–172.
nals may reduce the number of ECM-producing transdifferentiated HSCs [8] F. Heymann, F. Tacke, Immunology in the liver—from homeostasis to disease, Nat.
Rev. Gastroenterol. Hepatol. 13 (2016) 88–110.
and serve as an option of antifibrotic therapy. Indeed, liver fibrosis can [9] D.N. D'Ambrosio, J.L. Walewski, R.D. Clugston, P.D. Berk, R.A. Rippe, W.S. Blaner,
be reduced in vivo by the NF-κB inhibitor, BAY 11–7082, as well as by Distinct populations of hepatic stellate cells in the mouse liver have different ca-
proteasome inhibitors, bortezomib and MG132, via NF-κB target gene pacities for retinoid and lipid storage, PLoS One 6 (2011), e24993.
[10] T. Li, Z. Shi, D.C. Rockey, Preproendothelin-1 expression is negatively regulated by
A1 [319]. Inhibition of neurochemical receptors such as CB1 and 5HT IFNgamma during hepatic stellate cell activation, Am. J. Physiol. Gastrointest. Liver
can also induce HSC apoptosis. IFNγ is a hallmark of activated NK and Physiol. 302 (2012) G948–G957.
NKT cells that kill senescent HSCs and limits liver inflammation and fi- [11] R. Dobie, N.C. Henderson, Homing in on the hepatic scar: recent advances in cell-
specific targeting of liver fibrosis, F1000Research 5 (2016).
brosis [8,320]. IFNγ delivery specifically targeting activated HSCs by
[12] S.N. Greenhalgh, K.P. Conroy, N.C. Henderson, Cre-ativity in the liver: transgenic
PDGFRβ-mediated uptake suppresses HSCs and reduces hepatic fibrosis approaches to targeting hepatic nonparenchymal cells, Hepatology 61 (2015)
[321]. Nilotinib induces apoptosis and autophagic cell death of activated 2091–2099.
HSCs via inhibition of histone deacetylases [322]. A multi-kinase inhibi- [13] K. Poelstra, J. Prakash, L. Beljaars, Drug targeting to the diseased liver, J. Control. Re-
lease 161 (2012) 188–197.
tor, sorafenib, similarly induces autophagic cell death and apoptosis in [14] N. Kawada, Cytoglobin as a marker of hepatic stellate cell-derived myofibroblasts,
HSCs through the JNK and Akt signaling pathways [323]. Front. Physiol. 6 (2015) 329.
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 37

[15] D.Y. Zhang, N. Goossens, J. Guo, M.-C. Tsai, H.-I. Chou, C. Altunkaynak, et al., A he- [45] T. McHedlidze, M. Waldner, S. Zopf, J. Walker, A.L. Rankin, M. Schuchmann, et al.,
patic stellate cell gene expression signature associated with outcomes in hepatitis Interleukin-33-dependent innate lymphoid cells mediate hepatic fibrosis, Immuni-
C cirrhosis and hepatocellular carcinoma after curative resection, Gut 65 (2016) ty 39 (2013) 357–371.
1754–1764. [46] T. Lan, T. Kisseleva, D.A. Brenner, Deficiency of NOX1 or NOX4 prevents liver in-
[16] J. Ji, T. Eggert, A. Budhu, M. Forgues, A. Takai, H. Dang, et al., Hepatic stellate cell and flammation and fibrosis in mice through inhibition of hepatic stellate cell activa-
monocyte interaction contributes to poor prognosis in hepatocellular carcinoma, tion, PLoS One 10 (2015), e0129743.
Hepatology 62 (2015) 481–495. [47] J.X. Jiang, S. Venugopal, N. Serizawa, X. Chen, F. Scott, Y. Li, et al., Reduced nicotin-
[17] S.L. Friedman, F.J. Roll, Isolation and culture of hepatic lipocytes, Kupffer cells, and amide adenine dinucleotide phosphate oxidase 2 plays a key role in stellate cell ac-
sinusoidal endothelial cells by density gradient centrifugation with Stractan, Anal. tivation and liver fibrogenesis in vivo, Gastroenterology 139 (2010) 1375–1384.
Biochem. 161 (1987) 207–218. [48] T. Aoyama, Y.H. Paik, S. Watanabe, B. Laleu, F. Gaggini, L. Fioraso-Cartier, et al., Nic-
[18] M.R. Ebrahimkhani, I. Mohar, I.N. Crispe, Cross-presentation of antigen by diverse otinamide adenine dinucleotide phosphate oxidase in experimental liver fibrosis:
subsets of murine liver cells, Hepatology 54 (2011) 1379–1387. GKT137831 as a novel potential therapeutic agent, Hepatology 56 (2012)
[19] S. De Minicis, E. Seki, H. Uchinami, J. Kluwe, Y. Zhang, D.A. Brenner, et al., Gene ex- 2316–2327.
pression profiles during hepatic stellate cell activation in culture and in vivo, Gas- [49] N.O. Ku, P. Strnad, H. Bantel, M.B. Omary, Keratins: biomarkers and modulators of
troenterology 132 (2007) 1937–1946. apoptotic and necrotic cell death in the liver, Hepatology 64 (2016) 966–976.
[20] M. Fausther, J.R. Goree, E.G. Lavoie, A.L. Graham, J. Sevigny, J.A. Dranoff, Establish- [50] R.G. Wells, The role of matrix stiffness in regulating cell behavior, Hepatology 47
ment and characterization of rat portal myofibroblast cell lines, PLoS One 10 (2008) 1394–1400.
(2015), e0121161. [51] N.C. Henderson, T.D. Arnold, Y. Katamura, M.M. Giacomini, J.D. Rodriguez, J.H.
[21] L. Xu, A.Y. Hui, E. Albanis, M.J. Arthur, S.M. O'Byrne, W.S. Blaner, et al., Human he- McCarty, et al., Targeting of alphav integrin identifies a core molecular pathway
patic stellate cell lines, LX-1 and LX-2: new tools for analysis of hepatic fibrosis, Gut that regulates fibrosis in several organs, Nat. Med. 19 (2013) 1617–1624.
54 (2005) 142–151. [52] J. Gortzen, R. Schierwagen, J. Bierwolf, S. Klein, F.E. Uschner, P.F. van der Ven, et al., In-
[22] N. Shibata, T. Watanabe, T. Okitsu, M. Sakaguchi, M. Takesue, T. Kunieda, et al., Es- terplay of matrix stiffness and c-SRC in hepatic fibrosis, Front. Physiol. 6 (2015) 359.
tablishment of an immortalized human hepatic stellate cell line to develop [53] S.R. Caliari, M. Perepelyuk, B.D. Cosgrove, S.J. Tsai, G.Y. Lee, R.L. Mauck, et al., Stiff-
antifibrotic therapies, Cell Transplant. 12 (2003) 499–507. ening hydrogels for investigating the dynamics of hepatic stellate cell
[23] Y. Kim, V. Ratziu, S.G. Choi, A. Lalazar, G. Theiss, Q. Dang, et al., Transcriptional mechanotransduction during myofibroblast activation, Sci. Rep. 6 (2016) 21387.
activation of transforming growth factor beta1 and its receptors by the [54] S.R. Caliari, M. Perepelyuk, E.M. Soulas, G.Y. Lee, R.G. Wells, J.A. Burdick, Gradually
Kruppel-like factor Zf9/core promoter-binding protein and Sp1. Potential softening hydrogels for modeling hepatic stellate cell behavior during fibrosis re-
mechanisms for autocrine fibrogenesis in response to injury, J. Biol. Chem. gression, Integr. Biol. (Camb.) 8 (2016) 720–728.
273 (1998) 33750–33758. [55] C. Bonnans, J. Chou, Z. Werb, Remodelling the extracellular matrix in development
[24] S.K. Meurer, M. Alsamman, H. Sahin, H.E. Wasmuth, T. Kisseleva, D.A. Brenner, and disease, Nat. Rev. Mol. Cell Biol. 15 (2014) 786–801.
et al., Overexpression of endoglin modulates TGF-beta1-signalling pathways in a [56] E. Borkham-Kamphorst, C.R. van Roeyen, E. Van de Leur, J. Floege, R. Weiskirchen,
novel immortalized mouse hepatic stellate cell line, PLoS One 8 (2013), e56116. CCN3/NOV small interfering RNA enhances fibrogenic gene expression in primary
[25] R. Borojevic, A.N. Monteiro, S.A. Vinhas, G.B. Domont, P.A. Mourao, H. Emonard, hepatic stellate cells and cirrhotic fat storing cell line CFSC, J. Cell Commun. Signal.
et al., Establishment of a continuous cell line from fibrotic schistosomal granulo- 6 (2012) 11–25.
mas in mice livers, In Vitro Cell. Dev. Biol. 21 (1985) 382–390. [57] V. Barry-Hamilton, R. Spangler, D. Marshall, S. McCauley, H.M. Rodriguez, M. Oyasu,
[26] J. Guo, J. Loke, F. Zheng, F. Hong, S. Yea, M. Fukata, et al., Functional linkage of cir- et al., Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a
rhosis-predictive single nucleotide polymorphisms of Toll-like receptor 4 to hepat- pathologic microenvironment, Nat. Med. 16 (2010) 1009–1017.
ic stellate cell responses, Hepatology 49 (2009) 960–968. [58] F. Tacke, H.W. Zimmermann, Macrophage heterogeneity in liver injury and fibrosis,
[27] P. Greenwel, M. Schwartz, M. Rosas, S. Peyrol, J.A. Grimaud, M. Rojkind, Character- J. Hepatol. 60 (2014) 1090–1096.
ization of fat-storing cell lines derived from normal and CCl4-cirrhotic livers. Dif- [59] J.S. Duffield, S.J. Forbes, C.M. Constandinou, S. Clay, M. Partolina, S. Vuthoori, et al.,
ferences in the production of interleukin-6, Lab. Investig. 65 (1991) 644–653. Selective depletion of macrophages reveals distinct, opposing roles during liver in-
[28] I. Mederacke, C.C. Hsu, J.S. Troeger, P. Huebener, X. Mu, D.H. Dapito, et al., Fate trac- jury and repair, J. Clin. Invest. 115 (2005) 56–65.
ing reveals hepatic stellate cells as dominant contributors to liver fibrosis indepen- [60] A. Sica, P. Invernizzi, A. Mantovani, Macrophage plasticity and polarization in liver
dent of its aetiology, Nat. Commun. 4 (2013) 2823. homeostasis and pathology, Hepatology 59 (2014) 2034–2042.
[29] K. Iwaisako, C. Jiang, M. Zhang, M. Cong, T.J. Moore-Morris, T.J. Park, et al., Origin of [61] A. Lujambio, L. Akkari, J. Simon, D. Grace, D.F. Tschaharganeh, J.E. Bolden, et al.,
myofibroblasts in the fibrotic liver in mice, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) Non-cell-autonomous tumor suppression by p53, Cell 153 (2013) 449–460.
E3297–E3305. [62] P. Ramachandran, A. Pellicoro, M.A. Vernon, L. Boulter, R.L. Aucott, A. Ali, et al., Dif-
[30] T. Kisseleva, H. Uchinami, N. Feirt, O. Quintana-Bustamante, J.C. Segovia, R.F. ferential Ly-6C expression identifies the recruited macrophage phenotype, which
Schwabe, et al., Bone marrow-derived fibrocytes participate in pathogenesis of orchestrates the regression of murine liver fibrosis, Proc. Natl. Acad. Sci. U. S. A.
liver fibrosis, J. Hepatol. 45 (2006) 429–438. 109 (2012) E3186–E3195.
[31] T. Kisseleva, D.A. Brenner, The phenotypic fate and functional role for bone mar- [63] P.G. Traber, H. Chou, E. Zomer, F. Hong, A. Klyosov, M.I. Fiel, et al., Regression of fi-
row-derived stem cells in liver fibrosis, J. Hepatol. 56 (2012) 965–972. brosis and reversal of cirrhosis in rats by galectin inhibitors in thioacetamide-in-
[32] T. Kisseleva, The origin of fibrogenic myofibroblasts in fibrotic liver, Hepatology 65 duced liver disease, PLoS One 8 (2013), e75361.
(2017) 1039–1043. [64] S.A. Harrison, S.R. Marri, N. Chalasani, R. Kohli, W. Aronstein, G.A. Thompson, et al.,
[33] Y. Li, J. Wang, K. Asahina, Mesothelial cells give rise to hepatic stellate cells and Randomised clinical study: GR-MD-02, a galectin-3 inhibitor, vs. placebo in pa-
myofibroblasts via mesothelial-mesenchymal transition in liver injury, Proc. Natl. tients having non-alcoholic steatohepatitis with advanced fibrosis, Aliment.
Acad. Sci. U. S. A. 110 (2013) 2324–2329. Pharmacol. Ther. 44 (2016) 1183–1198.
[34] K. Taura, K. Miura, K. Iwaisako, C.H. Osterreicher, Y. Kodama, M. Penz-Osterreicher, [65] K.R. Karlmark, H.W. Zimmermann, C. Roderburg, N. Gassler, H.E. Wasmuth, T.
et al., Hepatocytes do not undergo epithelial-mesenchymal transition in liver fibro- Luedde, et al., The fractalkine receptor CX(3)CR1 protects against liver fibrosis by
sis in mice, Hepatology 51 (2010) 1027–1036. controlling differentiation and survival of infiltrating hepatic monocytes,
[35] D. Scholten, R. Weiskirchen, Questioning the challenging role of epithelial-to-mes- Hepatology 52 (2010) 1769–1782.
enchymal transition in liver injury, Hepatology 53 (2011) 1048–1051. [66] A. Pellicoro, R.L. Aucott, P. Ramachandran, A.J. Robson, J.A. Fallowfield, V.K.
[36] A.S. Chu, R. Diaz, J.J. Hui, K. Yanger, Y. Zong, G. Alpini, et al., Lineage tracing demon- Snowdon, et al., Elastin accumulation is regulated at the level of degradation by
strates no evidence of cholangiocyte epithelial-to-mesenchymal transition in mu- macrophage metalloelastase (MMP-12) during experimental liver fibrosis,
rine models of hepatic fibrosis, Hepatology 53 (2011) 1685–1695. Hepatology 55 (2012) 1965–1975.
[37] G.A. Michelotti, G. Xie, M. Swiderska, S.S. Choi, G. Karaca, L. Kruger, et al., Smooth- [67] L. Hammerich, F. Heymann, F. Tacke, Role of IL-17 and Th17 cells in liver diseases,
ened is a master regulator of adult liver repair, J. Clin. Invest. 123 (2013) Clin. Dev. Immunol. 2011 (2011) 345803.
2380–2394. [68] F. Meng, K. Wang, T. Aoyama, S.I. Grivennikov, Y. Paik, D. Scholten, et al., Interleu-
[38] M.C. Wallace, S.L. Friedman, D.A. Mann, Emerging and disease-specific mecha- kin-17 signaling in inflammatory, Kupffer cells, and hepatic stellate cells
nisms of hepatic stellate cell activation, Semin. Liver Dis. 35 (2015) 107–118. exacerbates liver fibrosis in mice, Gastroenterology 143 (2012) 765–776 (e761-
[39] T. Luedde, N. Kaplowitz, R.F. Schwabe, Cell death and cell death responses in liver 763).
disease: mechanisms and clinical relevance, Gastroenterology 147 (2014) 765–783 [69] X. Kong, D. Feng, H. Wang, F. Hong, A. Bertola, F.S. Wang, et al., Interleukin-22
(e764). induces hepatic stellate cell senescence and restricts liver fibrosis, Hepatology 56
[40] H. Shen, L. Sheng, Z. Chen, L. Jiang, H. Su, L. Yin, et al., Mouse hepatocyte overex- (2012) 1150–1159.
pression of NF-kappaB-inducing kinase (NIK) triggers fatal macrophage-depen- [70] D.H. Lu, X.Y. Guo, S.Y. Qin, W. Luo, X.L. Huang, M. Chen, et al., Interleukin-22 ame-
dent liver injury and fibrosis, Hepatology 60 (2014) 2065–2076. liorates liver fibrogenesis by attenuating hepatic stellate cell activation and down-
[41] A. Pellicoro, P. Ramachandran, J.P. Iredale, J.A. Fallowfield, Liver fibrosis and repair: regulating the levels of inflammatory cytokines, World J. Gastroenterol. 21 (2015)
immune regulation of wound healing in a solid organ, Nat. Rev. Immunol. 14 1531–1545.
(2014) 181–194. [71] W. Gao, Y.C. Fan, J.Y. Zhang, M.H. Zheng, Emerging role of interleukin 22 in hepati-
[42] C. Brenner, L. Galluzzi, O. Kepp, G. Kroemer, Decoding cell death signals in liver in- tis B virus infection: a double-edged sword, J. Clin. Transl. Hepatol. 1 (2013)
flammation, J. Hepatol. 59 (2013) 583–594. 103–108.
[43] J.K. Chan, J. Roth, J.J. Oppenheim, K.J. Tracey, T. Vogl, M. Feldmann, et al., Alarmins: [72] B. Langhans, B. Kramer, M. Louis, H.D. Nischalke, R. Huneburg, A. Staratschek-Jox,
awaiting a clinical response, J. Clin. Invest. 122 (2012) 2711–2719. et al., Intrahepatic IL-8 producing Foxp3(+)CD4(+) regulatory T cells and
[44] M.I. Arshad, C. Piquet-Pellorce, M. Samson, IL-33 and HMGB1 alarmins: sensors of fibrogenesis in chronic hepatitis C, J. Hepatol. 59 (2013) 229–235.
cellular death and their involvement in liver pathology, Liver Int. 32 (2012) [73] L. Hammerich, J.M. Bangen, O. Govaere, H.W. Zimmermann, N. Gassler, S. Huss,
1200–1210. et al., Chemokine receptor CCR6-dependent accumulation of gammadelta T cells
38 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

in injured liver restricts hepatic inflammation and fibrosis, Hepatology 59 (2014) other liver events in HIV and hepatitis C virus-coinfected individuals from 2001
630–642. to 2014: a multicohort study, Clin. Infect. Dis. 63 (2016) 821–829.
[74] T.I. Novobrantseva, G.R. Majeau, A. Amatucci, S. Kogan, I. Brenner, S. Casola, et al., [103] N. Brau, M. Salvatore, C.F. Rios-Bedoya, A. Fernandez-Carbia, F. Paronetto, J.F.
Attenuated liver fibrosis in the absence of B cells, J. Clin. Invest. 115 (2005) Rodriguez-Orengo, et al., Slower fibrosis progression in HIV/HCV-coinfected
3072–3082. patients with successful HIV suppression using antiretroviral therapy, J. Hepatol.
[75] M. Thapa, R. Chinnadurai, V.M. Velazquez, D. Tedesco, E. Elrod, J.H. Han, et al., Liver 44 (2006) 47–55.
fibrosis occurs through dysregulation of MyD88-dependent innate B-cell activity, [104] A.C. Tuyama, F. Hong, Y. Saiman, C. Wang, D. Ozkok, A. Mosoian, et al., Human im-
Hepatology 61 (2015) 2067–2079. munodeficiency virus (HIV)-1 infects human hepatic stellate cells and promotes
[76] S. Radaeva, R. Sun, B. Jaruga, V.T. Nguyen, Z. Tian, B. Gao, Natural killer cells amelio- collagen I and monocyte chemoattractant protein-1 expression: implications for
rate liver fibrosis by killing activated stellate cells in NKG2D-dependent and tumor the pathogenesis of HIV/hepatitis C virus-induced liver fibrosis, Hepatology 52
necrosis factor-related apoptosis-inducing ligand-dependent manners, Gastroen- (2010) 612–622.
terology 130 (2006) 435–452. [105] F. Hong, Y. Saiman, C. Si, A. Mosoian, M.B. Bansal, X4 human immunodeficiency
[77] J. Jiao, K. Ooka, H. Fey, M.I. Fiel, A.H. Rahmman, K. Kojima, et al., Interleukin-15 re- virus type 1 gp120 promotes human hepatic stellate cell activation and collagen
ceptor alpha on hepatic stellate cells regulates hepatic fibrogenesis in mice, J. I expression through interactions with CXCR4, PLoS One 7 (2012), e33659.
Hepatol. 65 (2016) 344–353. [106] S. Salloum, J.A. Holmes, R. Jindal, S.S. Bale, C. Brisac, N. Alatrakchi, et al., Exposure to
[78] B. Langhans, A.W. Alwan, B. Kramer, A. Glassner, P. Lutz, C.P. Strassburg, et al., Reg- human immunodeficiency virus/hepatitis C virus in hepatic and stellate cell lines
ulatory CD4+ T cells modulate the interaction between NK cells and hepatic stel- reveals cooperative profibrotic transcriptional activation between viruses and cell
late cells by acting on either cell type, J. Hepatol. 62 (2015) 398–404. types, Hepatology 64 (2016) 1951–1968.
[79] W.K. Syn, K.M. Agboola, M. Swiderska, G.A. Michelotti, E. Liaskou, H. Pang, et al., [107] M. Shi, J. Zhu, R. Wang, X. Chen, L. Mi, T. Walz, et al., Latent TGF-beta structure and
NKT-associated hedgehog and osteopontin drive fibrogenesis in non-alcoholic activation, Nature 474 (2011) 343–349.
fatty liver disease, Gut 61 (2012) 1323–1329. [108] S. Dooley, P. ten Dijke, TGF-beta in progression of liver disease, Cell Tissue Res. 347
[80] B. Gao, S. Radaeva, Natural killer and natural killer T cells in liver fibrosis, Biochim. (2012) 245–256.
Biophys. Acta 1832 (2013) 1061–1069. [109] Z. Szondy, Z. Sarang, P. Molnar, T. Nemeth, M. Piacentini, P.G. Mastroberardino,
[81] A. Wehr, C. Baeck, F. Heymann, P.M. Niemietz, L. Hammerich, C. Martin, et al., Che- et al., Transglutaminase 2−/− mice reveal a phagocytosis-associated crosstalk be-
mokine receptor CXCR6-dependent hepatic NK T cell accumulation promotes in- tween macrophages and apoptotic cells, Proc. Natl. Acad. Sci. U. S. A. 100 (2003)
flammation and liver fibrosis, J. Immunol. 190 (2013) 5226–5236. 7812–7817.
[82] J. Poisson, S. Lemoinne, C. Boulanger, F. Durand, R. Moreau, D. Valla, et al., Liver si- [110] C. Li, Y. Kong, H. Wang, S. Wang, H. Yu, X. Liu, et al., Homing of bone marrow mes-
nusoidal endothelial cells: physiology and role in liver diseases, J. Hepatol. 66 enchymal stem cells mediated by sphingosine 1-phosphate contributes to liver fi-
(2017) 212–227. brosis, J. Hepatol. 50 (2009) 1174–1183.
[83] G. Xie, X. Wang, L. Wang, L. Wang, R.D. Atkinson, G.C. Kanel, et al., Role of differen- [111] N.M. Meindl-Beinker, K. Matsuzaki, S. Dooley, TGF-beta signaling in onset and pro-
tiation of liver sinusoidal endothelial cells in progression and regression of hepatic gression of hepatocellular carcinoma, Dig. Dis. 30 (2012) 514–523.
fibrosis in rats, Gastroenterology 142 (2012) 918–927 (e916). [112] N. Sanderson, V. Factor, P. Nagy, J. Kopp, P. Kondaiah, L. Wakefield, et al., Hepatic
[84] M.J. Williams, A.D. Clouston, S.J. Forbes, Links between hepatic fibrosis, ductular re- expression of mature transforming growth factor beta 1 in transgenic mice results
action, and progenitor cell expansion, Gastroenterology 146 (2014) 349–356. in multiple tissue lesions, Proc. Natl. Acad. Sci. U. S. A. 92 (1995) 2572–2576.
[85] C. Kordes, I. Sawitza, S. Gotze, D. Herebian, D. Haussinger, Hepatic stellate cells con- [113] J. Schnur, P. Nagy, A. Sebestyen, Z. Schaff, S.S. Thorgeirsson, Chemical
tribute to progenitor cells and liver regeneration, J. Clin. Invest. 124 (2014) hepatocarcinogenesis in transgenic mice overexpressing mature TGF beta-1 in
5503–5515. liver, Eur. J. Cancer 35 (1999) 1842–1845.
[86] X. Wang, A. Lopategi, X. Ge, Y. Lu, N. Kitamura, R. Urtasun, et al., Osteopontin in- [114] C. Trautwein, S.L. Friedman, D. Schuppan, M. Pinzani, Hepatic fibrosis: concept to
duces ductular reaction contributing to liver fibrosis, Gut 63 (2014) 1805–1818. treatment, J. Hepatol. 62 (2015) S15–S24.
[87] Y. Wen, S. Jeong, Q. Xia, X. Kong, Role of osteopontin in liver diseases, Int. J. Biol. Sci. [115] L. Fang, C. Huang, X. Meng, B. Wu, T. Ma, X. Liu, et al., TGF-beta1-elevated TRPM7
12 (2016) 1121–1128. channel regulates collagen expression in hepatic stellate cells via TGF-beta1/
[88] T. Kodama, T. Takehara, H. Hikita, S. Shimizu, W. Li, T. Miyagi, et al., Thrombocyto- Smad pathway, Toxicol. Appl. Pharmacol. 280 (2014) 335–344.
penia exacerbates cholestasis-induced liver fibrosis in mice, Gastroenterology 138 [116] L. Fang, S. Zhan, C. Huang, X. Cheng, X. Lv, H. Si, et al., TRPM7 channel regulates
(2010) 2487–2498 (2498 e2481-2487). PDGF-BB-induced proliferation of hepatic stellate cells via PI3K and ERK pathways,
[89] Y. Rotman, A.J. Sanyal, Current and upcoming pharmacotherapy for non-alcoholic Toxicol. Appl. Pharmacol. 272 (2013) 713–725.
fatty liver disease, Gut 66 (2017) 180–190. [117] P. An, Y. Tian, M. Chen, H. Luo, Ca(2+)/calmodulin-dependent protein kinase II
[90] S. De Minicis, C. Rychlicki, L. Agostinelli, S. Saccomanno, C. Candelaresi, L. Trozzi, mediates transforming growth factor-beta-induced hepatic stellate cells prolifera-
et al., Dysbiosis contributes to fibrogenesis in the course of chronic liver injury in tion but not in collagen alpha1(I) production, Hepatol. Res. 42 (2012) 806–818.
mice, Hepatology 59 (2014) 1738–1749. [118] X. Wu, X. Wu, Y. Ma, F. Shao, Y. Tan, T. Tan, et al., CUG-binding protein 1 regulates
[91] M. Mazagova, L. Wang, A.T. Anfora, M. Wissmueller, S.A. Lesley, Y. Miyamoto, et al., HSC activation and liver fibrogenesis, Nat. Commun. 7 (2016) 13498.
Commensal microbiota is hepatoprotective and prevents liver fibrosis in mice, [119] F. Hong, H. Chou, M.I. Fiel, S.L. Friedman, Antifibrotic activity of sorafenib in exper-
FASEB J. 29 (2015) 1043–1055. imental hepatic fibrosis: refinement of inhibitory targets, dosing, and window of
[92] K. Tomita, T. Teratani, T. Suzuki, M. Shimizu, H. Sato, K. Narimatsu, et al., Free cho- efficacy in vivo, Dig. Dis. Sci. 58 (2013) 257–264.
lesterol accumulation in hepatic stellate cells: mechanism of liver fibrosis aggrava- [120] P. Kocabayoglu, A. Lade, Y.A. Lee, A.C. Dragomir, X. Sun, M.I. Fiel, et al., Beta-PDGF
tion in nonalcoholic steatohepatitis in mice, Hepatology 59 (2014) 154–169. receptor expressed by hepatic stellate cells regulates fibrosis in murine liver injury,
[93] J.D. Stanaway, A.D. Flaxman, M. Naghavi, C. Fitzmaurice, T. Vos, I. Abubakar, et al., but not carcinogenesis, J. Hepatol. 63 (2015) 141–147.
The global burden of viral hepatitis from 1990 to 2013: findings from the Global [121] P. Czochra, B. Klopcic, E. Meyer, J. Herkel, J.F. Garcia-Lazaro, F. Thieringer, et al.,
Burden of Disease Study 2013, Lancet 388 (2016) 1081–1088. Liver fibrosis induced by hepatic overexpression of PDGF-B in transgenic mice, J.
[94] Y. Zan, Y. Zhang, P. Tien, Hepatitis B virus e antigen induces activation of rat hepatic Hepatol. 45 (2006) 419–428.
stellate cells, Biochem. Biophys. Res. Commun. 435 (2013) 391–396. [122] Y. Kim, M.I. Fiel, E. Albanis, H.I. Chou, W. Zhang, G. Khitrov, et al., Anti-fibrotic ac-
[95] Q. Bai, J. An, X. Wu, H. You, H. Ma, T. Liu, et al., HBV promotes the proliferation of tivity and enhanced interleukin-6 production by hepatic stellate cells in response
hepatic stellate cells via the PDGF-B/PDGFR-beta signaling pathway in vitro, Int. to imatinib mesylate, Liver Int. 32 (2012) 1008–1017.
J. Mol. Med. 30 (2012) 1443–1450. [123] I.M. Westra, D. Oosterhuis, G.M. Groothuis, P. Olinga, Precision-cut liver slices as a
[96] A. Florimond, P. Chouteau, P. Bruscella, J. Le Seyec, E. Merour, N. Ahnou, et al., model for the early onset of liver fibrosis to test antifibrotic drugs, Toxicol. Appl.
Human hepatic stellate cells are not permissive for hepatitis C virus entry and rep- Pharmacol. 274 (2014) 328–338.
lication, Gut 64 (2015) 957–965. [124] Y. Liu, Z. Wang, S.Q. Kwong, E.L. Lui, S.L. Friedman, F.R. Li, et al., Inhibition of PDGF,
[97] R. Bataller, Y.H. Paik, J.N. Lindquist, J.J. Lemasters, D.A. Brenner, Hepatitis C virus TGF-beta, and Abl signaling and reduction of liver fibrosis by the small molecule
core and nonstructural proteins induce fibrogenic effects in hepatic stellate cells, Bcr-Abl tyrosine kinase antagonist Nilotinib, J. Hepatol. 55 (2011) 612–625.
Gastroenterology 126 (2004) 529–540. [125] S. Tugues, G. Fernandez-Varo, J. Munoz-Luque, J. Ros, V. Arroyo, J. Rodes, et al.,
[98] S. Battaglia, N. Benzoubir, S. Nobilet, P. Charneau, D. Samuel, A.L. Zignego, et al., Antiangiogenic treatment with sunitinib ameliorates inflammatory infiltrate, fi-
Liver cancer-derived hepatitis C virus core proteins shift TGF-beta responses brosis, and portal pressure in cirrhotic rats, Hepatology 46 (2007) 1919–1926.
from tumor suppression to epithelial-mesenchymal transition, PLoS One 4 [126] A. Wilhelm, V. Aldridge, D. Haldar, A.J. Naylor, C.J. Weston, D. Hedegaard, et al.,
(2009), e4355. CD248/endosialin critically regulates hepatic stellate cell proliferation during
[99] N. Benzoubir, C. Lejamtel, S. Battaglia, B. Testoni, B. Benassi, C. Gondeau, et al., HCV chronic liver injury via a PDGF-regulated mechanism, Gut 65 (2016) 1175–1185.
core-mediated activation of latent TGF-beta via thrombospondin drives the [127] E. Tibaldi, F. Zonta, L. Bordin, E. Magrin, E. Gringeri, U. Cillo, et al., The tyrosine
crosstalk between hepatocytes and stromal environment, J. Hepatol. 59 (2013) phosphatase SHP-1 inhibits proliferation of activated hepatic stellate cells by
1160–1168. impairing PDGF receptor signaling, Biochim. Biophys. Acta 1843 (2014) 288–298.
[100] L.D. Presser, S. McRae, G. Waris, Activation of TGF-beta1 promoter by hepatitis C [128] X. Zhang, Z. Tan, Y. Wang, J. Tang, R. Jiang, J. Hou, et al., PTPRO-associated hepatic
virus-induced AP-1 and Sp1: role of TGF-beta1 in hepatic stellate cell activation stellate cell activation plays a critical role in liver fibrosis, Cell. Physiol. Biochem.
and invasion, PLoS One 8 (2013), e56367. 35 (2015) 885–898.
[101] L. Preisser, C. Miot, H. Le Guillou-Guillemette, E. Beaumont, E.D. Foucher, E. Garo, [129] J.S. Campbell, S.D. Hughes, D.G. Gilbertson, T.E. Palmer, M.S. Holdren, A.C. Haran,
et al., IL-34 and macrophage colony-stimulating factor are overexpressed in hepa- et al., Platelet-derived growth factor C induces liver fibrosis, steatosis, and hepato-
titis C virus fibrosis and induce profibrotic macrophages that promote collagen cellular carcinoma, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 3389–3394.
synthesis by hepatic stellate cells, Hepatology 60 (2014) 1879–1890. [130] T.H. Tsai, S.C. Shih, T.C. Ho, H.I. Ma, M.Y. Liu, S.L. Chen, et al., Pigment epithelium-de-
[102] L.I. Gjaerde, L. Shepherd, E. Jablonowska, A. Lazzarin, M. Rougemont, K. Darling, rived factor 34-mer peptide prevents liver fibrosis and hepatic stellate cell activation
et al., Trends in incidences and risk factors for hepatocellular carcinoma and through down-regulation of the PDGF receptor, PLoS One 9 (2014), e95443.
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 39

[131] C. Hao, Y. Xie, M. Peng, L. Ma, Y. Zhou, Y. Zhang, et al., Inhibition of connective tis- [158] S.I. Chung, H. Moon, H.L. Ju, K.J. Cho, D.Y. Kim, K.H. Han, et al., Hepatic expression of
sue growth factor suppresses hepatic stellate cell activation in vitro and prevents Sonic Hedgehog induces liver fibrosis and promotes hepatocarcinogenesis in a
liver fibrosis in vivo, Clin. Exp. Med. 14 (2014) 141–150. transgenic mouse model, J. Hepatol. 64 (2016) 618–627.
[132] P. Hauff, U. Gottwald, M. Ocker, Early to phase II drugs currently under investiga- [159] Y. Chen, S.S. Choi, G.A. Michelotti, I.S. Chan, M. Swiderska-Syn, G.F. Karaca, et al.,
tion for the treatment of liver fibrosis, Expert Opin. Investig. Drugs 24 (2015) Hedgehog controls hepatic stellate cell fate by regulating metabolism, Gastroenter-
309–327. ology 143 (2012) 1319–1329 (e1311-1311).
[133] K.E. Lipson, C. Wong, Y. Teng, S. Spong, CTGF is a central mediator of tissue remod- [160] T. Gorojankina, Hedgehog signaling pathway: a novel model and molecular
eling and fibrosis and its inhibition can reverse the process of fibrosis, Fibrogenesis mechanisms of signal transduction, Cell. Mol. Life Sci. 73 (2016) 1317–1332.
Tissue Repair 5 (2012) S24. [161] R. Kramann, R.K. Schneider, D.P. DiRocco, F. Machado, S. Fleig, P.A. Bondzie, et al.,
[134] B.C. Fuchs, Y. Hoshida, T. Fujii, L. Wei, S. Yamada, G.Y. Lauwers, et al., Epidermal Perivascular Gli1 + progenitors are key contributors to injury-induced organ
growth factor receptor inhibition attenuates liver fibrosis and development of he- fibrosis, Cell Stem Cell 16 (2015) 51–66.
patocellular carcinoma, Hepatology 59 (2014) 1577–1590. [162] N.N. El-Agroudy, R.N. El-Naga, R.A. El-Razeq, E. El-Demerdash, Forskolin, a hedge-
[135] H. Lanaya, A. Natarajan, K. Komposch, L. Li, N. Amberg, L. Chen, et al., EGFR has a hog signalling inhibitor, attenuates carbon tetrachloride-induced liver fibrosis in
tumour-promoting role in liver macrophages during hepatocellular carcinoma for- rats, Br. J. Pharmacol. 173 (2016) 3248–3260.
mation, Nat. Cell Biol. 16 (2014) 972–981 (971-977). [163] A. Pratap, S. Singh, V. Mundra, N. Yang, R. Panakanti, J.D. Eason, et al., Attenuation
[136] S.S. Choi, J.K. Sicklick, Q. Ma, L. Yang, J. Huang, Y. Qi, et al., Sustained activation of of early liver fibrosis by pharmacological inhibition of smoothened receptor signal-
Rac1 in hepatic stellate cells promotes liver injury and fibrosis in mice, Hepatology ing, J. Drug Target. 20 (2012) 770–782.
44 (2006) 1267–1277. [164] P. Hirsova, S.H. Ibrahim, S.F. Bronk, H. Yagita, G.J. Gores, Vismodegib suppresses
[137] M. Fukushima, M. Nakamuta, M. Kohjima, K. Kotoh, M. Enjoji, N. Kobayashi, et al., TRAIL-mediated liver injury in a mouse model of nonalcoholic steatohepatitis,
Fasudil hydrochloride hydrate, a Rho-kinase (ROCK) inhibitor, suppresses collagen PLoS One 8 (2013), e70599.
production and enhances collagenase activity in hepatic stellate cells, Liver Int. 25 [165] G. Xie, G. Karaca, M. Swiderska-Syn, G.A. Michelotti, L. Kruger, Y. Chen, et al., Cross-
(2005) 829–838. talk between Notch and Hedgehog regulates hepatic stellate cell fate in mice,
[138] M.A. Sohail, A.Z. Hashmi, W. Hakim, A. Watanabe, A. Zipprich, R.J. Groszmann, et al., Hepatology 58 (2013) 1801–1813.
Adenosine induces loss of actin stress fibers and inhibits contraction in hepatic [166] A. Mallat, F. Teixeira-Clerc, S. Lotersztajn, Cannabinoid signaling and liver thera-
stellate cells via Rho inhibition, Hepatology 49 (2009) 185–194. peutics, J. Hepatol. 59 (2013) 891–896.
[139] T. Murata, S. Arii, T. Nakamura, A. Mori, T. Kaido, H. Furuyama, et al., Inhibitory ef- [167] A.B. Gueye, Y. Pryslawsky, J.M. Trigo, N. Poulia, F. Delis, K. Antoniou, et al., The CB1
fect of Y-27632, a ROCK inhibitor, on progression of rat liver fibrosis in association neutral antagonist AM4113 retains the therapeutic efficacy of the inverse agonist
with inactivation of hepatic stellate cells, J. Hepatol. 35 (2001) 474–481. rimonabant for nicotine dependence and weight loss with better psychiatric toler-
[140] S. Klein, M.M. Van Beuge, M. Granzow, L. Beljaars, R. Schierwagen, S. Kilic, et al., ability, Int. J. Neuropsychopharmacol. 19 (2016).
HSC-specific inhibition of Rho-kinase reduces portal pressure in cirrhotic rats [168] T. Hua, K. Vemuri, M. Pu, L. Qu, G.W. Han, Y. Wu, et al., Crystal structure of the
without major systemic effects, J. Hepatol. 57 (2012) 1220–1227. human cannabinoid receptor CB1, Cell 167 (2016) 750–762 (e714).
[141] K. Martin, J. Pritchett, J. Llewellyn, A.F. Mullan, V.S. Athwal, R. Dobie, et al., PAK pro- [169] N. Coppola, R. Zampino, G. Bellini, M. Macera, A. Marrone, M. Pisaturo, et al.,
teins and YAP-1 signalling downstream of integrin beta-1 in myofibroblasts pro- Association between a polymorphism in cannabinoid receptor 2 and severe
mote liver fibrosis, Nat. Commun. 7 (2016) 12502. necroinflammation in patients with chronic hepatitis C, Clin. Gastroenterol.
[142] K.H. Kim, C.C. Chen, R.I. Monzon, L.F. Lau, Matricellular protein CCN1 promotes re- Hepatol. 12 (2014) 334–340.
gression of liver fibrosis through induction of cellular senescence in hepatic [170] A. Guillot, N. Hamdaoui, A. Bizy, K. Zoltani, R. Souktani, E.S. Zafrani, et al., Cannabi-
myofibroblasts, Mol. Cell. Biol. 33 (2013) 2078–2090. noid receptor 2 counteracts interleukin-17-induced immune and fibrogenic
[143] M. Granzow, R. Schierwagen, S. Klein, B. Kowallick, S. Huss, M. Linhart, et al., Angio- responses in mouse liver, Hepatology 59 (2014) 296–306.
tensin-II type 1 receptor-mediated Janus kinase 2 activation induces liver fibrosis, [171] S. El Swefy, R.A. Hasan, A. Ibrahim, M.F. Mahmoud, Curcumin and hemopressin
Hepatology 60 (2014) 334–348. treatment attenuates cholestasis-induced liver fibrosis in rats: role of CB1 receptors,
[144] S. Klein, J. Rick, J. Lehmann, R. Schierwagen, I.G. Schierwagen, L. Verbeke, et al., Naunyn Schmiedeberg's Arch. Pharmacol. 389 (2016) 103–116.
Janus-kinase-2 relates directly to portal hypertension and to complications in ro- [172] Y. Tang, Curcumin targets multiple pathways to halt hepatic stellate cell
dent and human cirrhosis, Gut 66 (2017) 145–155. activation: updated mechanisms in vitro and in vivo, Dig. Dis. Sci. 60 (2015)
[145] Z.A. Salama, A. Sadek, A.M. Abdelhady, S.K. Darweesh, S.A. Morsy, G. Esmat, 1554–1564.
Losartan may inhibit the progression of liver fibrosis in chronic HCV patients, [173] M.R. Ebrahimkhani, F. Oakley, L.B. Murphy, J. Mann, A. Moles, M.J. Perugorria, et al.,
Hepatobiliary Surg. Nutr. 5 (2016) 249–255. Stimulating healthy tissue regeneration by targeting the 5-HT(2)B receptor in
[146] T. Namisaki, R. Noguchi, K. Moriya, M. Kitade, Y. Aihara, A. Douhara, et al., Beneficial chronic liver disease, Nat. Med. 17 (2011) 1668–1673.
effects of combined ursodeoxycholic acid and angiotensin-II type 1 receptor [174] D.C. Kim, D.W. Jun, Y.I. Kwon, K.N. Lee, H.L. Lee, O.Y. Lee, et al., 5-HT2A receptor
blocker on hepatic fibrogenesis in a rat model of nonalcoholic steatohepatitis, J. antagonists inhibit hepatic stellate cell activation and facilitate apoptosis, Liver
Gastroenterol. 51 (2016) 162–172. Int. 33 (2013) 535–543.
[147] R. Rozenfeld, A. Gupta, K. Gagnidze, M.P. Lim, I. Gomes, D. Lee-Ramos, et al., AT1R- [175] H. Tilg, P.D. Cani, E.A. Mayer, Gut microbiome and liver diseases, Gut 65 (2016)
CB(1)R heteromerization reveals a new mechanism for the pathogenic properties 2035–2044.
of angiotensin II, EMBO J. 30 (2011) 2350–2363. [176] N. Qin, F. Yang, A. Li, E. Prifti, Y. Chen, L. Shao, et al., Alterations of the human gut
[148] S. Li, Q. Wang, Y. Tao, C. Liu, Swertiamarin attenuates experimental rat hepatic microbiome in liver cirrhosis, Nature 513 (2014) 59–64.
fibrosis by suppressing angiotensin II-angiotensin type 1 receptor-extracellu- [177] S. Kiziltas, Toll-like receptors in pathophysiology of liver diseases, World J. Hepatol.
lar signal-regulated kinase signaling, J. Pharmacol. Exp. Ther. 359 (2016) 8 (2016) 1354–1369.
247–255. [178] E. Seki, S. De Minicis, C.H. Osterreicher, J. Kluwe, Y. Osawa, D.A. Brenner, et al., TLR4
[149] C.G. Miao, Y.Y. Yang, X. He, C. Huang, Y. Huang, L. Zhang, et al., Wnt signaling in enhances TGF-beta signaling and hepatic fibrosis, Nat. Med. 13 (2007) 1324–1332.
liver fibrosis: progress, challenges and potential directions, Biochimie 95 (2013) [179] D.H. Dapito, A. Mencin, G.Y. Gwak, J.P. Pradere, M.K. Jang, I. Mederacke, et al., Pro-
2326–2335. motion of hepatocellular carcinoma by the intestinal microbiota and TLR4, Cancer
[150] W.S. Ge, Y.J. Wang, J.X. Wu, J.G. Fan, Y.W. Chen, L. Zhu, Beta-catenin is Cell 21 (2012) 504–516.
overexpressed in hepatic fibrosis and blockage of Wnt/beta-catenin signaling [180] A. Watanabe, A. Hashmi, D.A. Gomes, T. Town, A. Badou, R.A. Flavell, et al., Apopto-
inhibits hepatic stellate cell activation, Mol. Med. Rep. 9 (2014) 2145–2151. tic hepatocyte DNA inhibits hepatic stellate cell chemotaxis via toll-like receptor 9,
[151] J.H. Cheng, H. She, Y.P. Han, J. Wang, S. Xiong, K. Asahina, et al., Wnt antagonism Hepatology 46 (2007) 1509–1518.
inhibits hepatic stellate cell activation and liver fibrosis, Am. J. Physiol. Gastrointest. [181] K. Miura, Y. Kodama, S. Inokuchi, B. Schnabl, T. Aoyama, H. Ohnishi, et al., Toll-like
Liver Physiol. 294 (2008) G39–G49. receptor 9 promotes steatohepatitis by induction of interleukin-1beta in mice,
[152] N.L. Zhu, K. Asahina, J. Wang, A. Ueno, R. Lazaro, Y. Miyaoka, et al., Hepatic stellate Gastroenterology 139 (2010) 323–334 (e327).
cell-derived delta-like homolog 1 (DLK1) protein in liver regeneration, J. Biol. [182] P. Hartmann, M. Haimerl, M. Mazagova, D.A. Brenner, B. Schnabl, Toll-like receptor
Chem. 287 (2012) 10355–10367. 2-mediated intestinal injury and enteric tumor necrosis factor receptor I contribute
[153] Y. Osawa, K. Oboki, J. Imamura, E. Kojika, Y. Hayashi, T. Hishima, et al., Inhibition of to liver fibrosis in mice, Gastroenterology 143 (2012) 1330–1340 (e1331).
cyclic adenosine monophosphate (cAMP)-response element-binding protein [183] K. Miura, L. Yang, N. van Rooijen, D.A. Brenner, H. Ohnishi, E. Seki, Toll-like receptor
(CREB)-binding protein (CBP)/beta-catenin reduces liver fibrosis in mice, 2 and palmitic acid cooperatively contribute to the development of nonalcoholic
EBioMedicine 2 (2015) 1751–1758. steatohepatitis through inflammasome activation in mice, Hepatology 57 (2013)
[154] C. Kordes, I. Sawitza, D. Haussinger, Canonical Wnt signaling maintains the quies- 577–589.
cent stage of hepatic stellate cells, Biochem. Biophys. Res. Commun. 367 (2008) [184] W. Seo, H.S. Eun, S.Y. Kim, H.S. Yi, Y.S. Lee, S.H. Park, et al., Exosome-mediated
116–123. activation of toll-like receptor 3 in stellate cells stimulates interleukin-17 produc-
[155] Y. Xinguang, Y. Huixing, W. Linlin, W. Wanxin, W. Xiaojun, Y. Linghua, RSPOs tion by gammadelta T cells in liver fibrosis, Hepatology 64 (2016) 616–631.
facilitated HSC activation and promoted hepatic fibrogenesis, Oncotarget 7 (2016) [185] E. Kaffe, A. Katsifa, N. Xylourgidis, I. Ninou, M. Zannikou, V. Harokopos, et al., Hepa-
63767–63778. tocyte autotaxin expression promotes liver fibrosis and cancer, Hepatology 65
[156] L. Corbett, J. Mann, D.A. Mann, Non-canonical Wnt predominates in activated rat (2017) 1369–1383.
hepatic stellate cells, influencing HSC survival and paracrine stimulation of Kupffer [186] S. Nakagawa, L. Wei, W.M. Song, T. Higashi, S. Ghoshal, R.S. Kim, et al., Molecular
cells, PLoS One 10 (2015), e0142794. liver cancer prevention in cirrhosis by organ transcriptome analysis and
[157] N. Chatani, Y. Kamada, T. Kizu, S. Ogura, K. Furuta, M. Egawa, et al., Secreted lysophosphatidic acid pathway inhibition, Cancer Cell 30 (2016) 879–890.
frizzled-related protein 5 (Sfrp5) decreases hepatic stellate cell activation and [187] N. Ouchi, J.L. Parker, J.J. Lugus, K. Walsh, Adipokines in inflammation and metabolic
liver fibrosis, Liver Int. 35 (2015) 2017–2026. disease, Nat. Rev. Immunol. 11 (2011) 85–97.
40 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

[188] X. Ding, N.K. Saxena, S. Lin, A. Xu, S. Srinivasan, F.A. Anania, The roles of leptin and contributes to hepatic stellate cell activation, Biochim. Biophys. Acta 1842 (2014)
adiponectin: a novel paradigm in adipocytokine regulation of liver fibrosis and 2367–2377.
stellate cell biology, Am. J. Pathol. 166 (2005) 1655–1669. [216] M.S. Shafiei, S. Shetty, P.E. Scherer, D.C. Rockey, Adiponectin regulation of stellate
[189] N.K. Saxena, F.A. Anania, Adipocytokines and hepatic fibrosis, Trends Endocrinol. cell activation via PPARgamma-dependent and -independent mechanisms, Am. J.
Metab. 26 (2015) 153–161. Pathol. 178 (2011) 2690–2699.
[190] X. Zhai, K. Yan, J. Fan, M. Niu, Q. Zhou, Y. Zhou, et al., The beta-catenin pathway [217] X. Zhai, H. Qiao, W. Guan, Z. Li, Y. Cheng, X. Jia, et al., Curcumin regulates peroxi-
contributes to the effects of leptin on SREBP-1c expression in rat hepatic stellate some proliferator-activated receptor-gamma coactivator-1alpha expression by
cells and liver fibrosis, Br. J. Pharmacol. 169 (2013) 197–212. AMPK pathway in hepatic stellate cells in vitro, Eur. J. Pharmacol. 746 (2015)
[191] Z. Dong, L. Su, S. Esmaili, T.J. Iseli, M. Ramezani-Moghadam, L. Hu, et al., 56–62.
Adiponectin attenuates liver fibrosis by inducing nitric oxide production of hepatic [218] H. Jin, N. Lian, F. Zhang, L. Chen, Q. Chen, C. Lu, et al., Activation of PPARgamma/P53
stellate cells, J. Mol. Med. (Berl.) 93 (2015) 1327–1339. signaling is required for curcumin to induce hepatic stellate cell senescence, Cell
[192] M. Ramezani-Moghadam, J. Wang, V. Ho, T.J. Iseli, B. Alzahrani, A. Xu, et al., Death Dis. 7 (2016), e2189.
Adiponectin reduces hepatic stellate cell migration by promoting tissue inhibitor [219] A.J. Sanyal, N. Chalasani, K.V. Kowdley, A. McCullough, A.M. Diehl, N.M. Bass, et al.,
of metalloproteinase-1 (TIMP-1) secretion, J. Biol. Chem. 290 (2015) 5533–5542. Pioglitazone, vitamin E, or placebo for nonalcoholic steatohepatitis, N. Engl. J. Med.
[193] M. Tardelli, V. Moreno-Viedma, M. Zeyda, B.K. Itariu, F.B. Langer, G. Prager, et al., 362 (2010) 1675–1685.
Adiponectin regulates aquaglyceroporin expression in hepatic stellate cells altering [220] V. Ratziu, S.A. Harrison, S. Francque, P. Bedossa, P. Lehert, L. Serfaty, et al.,
their functional state, J. Gastroenterol. Hepatol. 32 (2017) 253–260. Elafibranor, an agonist of the peroxisome proliferator-activated receptor-alpha
[194] P. Kumar, T. Smith, K. Rahman, J.E. Mells, N.E. Thorn, N.K. Saxena, et al., Adiponectin and -delta, induces resolution of nonalcoholic steatohepatitis without fibrosis
modulates focal adhesion disassembly in activated hepatic stellate cells: implica- worsening, Gastroenterology 150 (2016) 1147–1159 (e1145).
tion for reversing hepatic fibrosis, FASEB J. 28 (2014) 5172–5183. [221] S.W. Beaven, K. Wroblewski, J. Wang, C. Hong, S. Bensinger, H. Tsukamoto, et al.,
[195] H. Wang, H. Zhang, Z. Zhang, B. Huang, X. Cheng, D. Wang, et al., Adiponectin- Liver X receptor signaling is a determinant of stellate cell activation and suscepti-
derived active peptide ADP355 exerts anti-inflammatory and anti-fibrotic activities bility to fibrotic liver disease, Gastroenterology 140 (2011) 1052–1062.
in thioacetamide-induced liver injury, Sci. Rep. 6 (2016) 19445. [222] K. Yan, X. Deng, X. Zhai, M. Zhou, X. Jia, L. Luo, et al., p38 mitogen-activated protein ki-
[196] P.F. Lalor, S. Edwards, G. McNab, M. Salmi, S. Jalkanen, D.H. Adams, Vascular adhe- nase and liver X receptor-alpha mediate the leptin effect on sterol regulatory element
sion protein-1 mediates adhesion and transmigration of lymphocytes on human binding protein-1c expression in hepatic stellate cells, Mol. Med. 18 (2012) 10–18.
hepatic endothelial cells, J. Immunol. 169 (2002) 983–992. [223] N. Ding, R.T. Yu, N. Subramaniam, M.H. Sherman, C. Wilson, R. Rao, et al., A vitamin
[197] C.J. Weston, E.L. Shepherd, L.C. Claridge, P. Rantakari, S.M. Curbishley, J.W. D receptor/SMAD genomic circuit gates hepatic fibrotic response, Cell 153 (2013)
Tomlinson, et al., Vascular adhesion protein-1 promotes liver inflammation and 601–613.
drives hepatic fibrosis, J. Clin. Invest. 125 (2015) 501–520. [224] A. Duran, E.D. Hernandez, M. Reina-Campos, E.A. Castilla, S. Subramaniam, S.
[198] B.J. Dwyer, J.K. Olynyk, G.A. Ramm, J.E. Tirnitz-Parker, TWEAK and LTbeta signaling Raghunandan, et al., p62/SQSTM1 by binding to vitamin D receptor inhibits hepat-
during chronic liver disease, Front. Immunol. 5 (2014) 39. ic stellate cell activity, fibrosis, and liver cancer, Cancer Cell 30 (2016) 595–609.
[199] A. Wilhelm, E.L. Shepherd, A. Amatucci, M. Munir, G. Reynolds, E. Humphreys, et al., [225] A. Beilfuss, J.P. Sowa, S. Sydor, M. Beste, L.P. Bechmann, M. Schlattjan, et al., Vitamin
Interaction of TWEAK with Fn14 leads to the progression of fibrotic liver disease by D counteracts fibrogenic TGF-beta signalling in human hepatic stellate cells both
directly modulating hepatic stellate cell proliferation, J. Pathol. 239 (2016) 109–121. receptor-dependently and independently, Gut 64 (2015) 791–799.
[200] F. Zhang, M. Zhang, A. Wang, M. Xu, C. Wang, G. Xu, et al., TWEAK increases SIRT1 [226] R. Zhang, Y. Wang, R. Li, G. Chen, Transcriptional factors mediating retinoic acid sig-
expression and promotes p53 deacetylation affecting human hepatic stellate cell nals in the control of energy metabolism, Int. J. Mol. Sci. 16 (2015) 14210–14244.
senescence, Cell Biol. Int. 41 (2017) 147–154. [227] M. Shimizu, Y. Shirakami, T. Hanai, K. Imai, A. Suetsugu, K. Takai, et al., Pharmaceu-
[201] Y.H. Li, D.H. Choi, E.H. Lee, S.R. Seo, S. Lee, E.H. Cho, Sirtuin 3 (SIRT3) regulates tical and nutraceutical approaches for preventing liver carcinogenesis: chemopre-
alpha-smooth muscle actin (alpha-SMA) production through the succinate dehy- vention of hepatocellular carcinoma using acyclic retinoid and branched-chain
drogenase-G protein-coupled receptor 91 (GPR91) pathway in hepatic stellate amino acids, Mol. Nutr. Food Res. 58 (2014) 124–135.
cells, J. Biol. Chem. 291 (2016) 10277–10292. [228] K. Okita, N. Izumi, O. Matsui, K. Tanaka, S. Kaneko, H. Moriwaki, et al., Peretinoin
[202] C. Barcena, M. Stefanovic, A. Tutusaus, L. Joannas, A. Menendez, C. Garcia-Ruiz, after curative therapy of hepatitis C-related hepatocellular carcinoma: a random-
et al., Gas6/Axl pathway is activated in chronic liver disease and its targeting ized double-blind placebo-controlled study, J. Gastroenterol. 50 (2015) 191–202.
reduces fibrosis via hepatic stellate cell inactivation, J. Hepatol. 63 (2015) 670–678. [229] K. Hellemans, I. Grinko, K. Rombouts, D. Schuppan, A. Geerts, All-trans and 9-cis
[203] B.A. Neuschwander-Tetri, R. Loomba, A.J. Sanyal, J.E. Lavine, M.L. Van Natta, M.F. retinoic acid alter rat hepatic stellate cell phenotype differentially, Gut 45 (1999)
Abdelmalek, et al., Farnesoid X nuclear receptor ligand obeticholic acid for non-cir- 134–142.
rhotic, non-alcoholic steatohepatitis (FLINT): a multicentre, randomised, placebo- [230] S. Hisamori, C. Tabata, Y. Kadokawa, K. Okoshi, R. Tabata, A. Mori, et al., All-trans-
controlled trial, Lancet 385 (2015) 956–965. retinoic acid ameliorates carbon tetrachloride-induced liver fibrosis in mice
[204] J.P. Arab, S.J. Karpen, P.A. Dawson, M. Arrese, M. Trauner, Bile acids and nonalcohol- through modulating cytokine production, Liver Int. 28 (2008) 1217–1225.
ic fatty liver disease: molecular insights and therapeutic perspectives, Hepatology [231] E. Sharvit, S. Abramovitch, S. Reif, R. Bruck, Amplified inhibition of stellate cell ac-
65 (2017) 350–362. tivation pathways by PPAR-gamma, RAR and RXR agonists, PLoS One 8 (2013),
[205] L. Verbeke, R. Farre, J. Trebicka, M. Komuta, T. Roskams, S. Klein, et al., Obeticholic e76541.
acid, a farnesoid X receptor agonist, improves portal hypertension by two distinct [232] R. Bruck, S. Weiss, H. Aeed, M. Pines, Z. Halpern, I. Zvibel, Additive inhibitory effect of
pathways in cirrhotic rats, Hepatology 59 (2014) 2286–2298. experimentally induced hepatic cirrhosis by agonists of peroxisome proliferator acti-
[206] W. Xu, C. Lu, F. Zhang, J. Shao, S. Zheng, Dihydroartemisinin restricts hepatic vator receptor gamma and retinoic acid receptor, Dig. Dis. Sci. 54 (2009) 292–299.
stellate cell contraction via an FXR-S1PR2-dependent mechanism, IUBMB Life 68 [233] D.J. Kojetin, T.P. Burris, REV-ERB and ROR nuclear receptors as drug targets, Nat.
(2016) 376–387. Rev. Drug Discov. 13 (2014) 197–216.
[207] W. Xu, C. Lu, F. Zhang, J. Shao, S. Yao, S. Zheng, Dihydroartemisinin counteracts [234] T. Li, A.L. Eheim, S. Klein, F.E. Uschner, A.C. Smith, E. Brandon-Warner, et al., Novel
fibrotic portal hypertension via farnesoid X receptor-dependent inhibition of role of nuclear receptor Rev-erbalpha in hepatic stellate cell activation: potential
hepatic stellate cell contraction, FEBS J. 284 (2017) 114–133. therapeutic target for liver injury, Hepatology 59 (2014) 2383–2396.
[208] B. Gross, M. Pawlak, P. Lefebvre, B. Staels, PPARs in obesity-induced T2DM, [235] P.G. Thomes, E. Brandon-Warner, T. Li, T.M. Donohue Jr., L.W. Schrum, Rev-erb ag-
dyslipidaemia and NAFLD, Nat. Rev. Endocrinol. 13 (2017) 36–49. onist and TGF-beta similarly affect autophagy but differentially regulate hepatic
[209] H. Tsukamoto, H. She, S. Hazra, J. Cheng, T. Miyahara, Anti-adipogenic regulation stellate cell fibrogenic phenotype, Int. J. Biochem. Cell Biol. 81 (2016) 137–147.
underlies hepatic stellate cell transdifferentiation, J. Gastroenterol. Hepatol. 21 [236] C. Yin, Molecular mechanisms of Sox transcription factors during the development
(Suppl. 3) (2006) S102–S105. of liver, bile duct, and pancreas, Semin. Cell Dev. Biol. 63 (2017) 68–78.
[210] F. Zhang, D. Kong, L. Chen, X. Zhang, N. Lian, X. Zhu, et al., Peroxisome proliferator- [237] H. Qiao, Q. Cao, Y. Fu, W. Guan, F. Cheng, J. Wu, et al., Sex-determining region Y-box 9
activated receptor-gamma interrupts angiogenic signal transduction by acts downstream of NADPH oxidase to influence the effect of leptin on PPARgamma1
transrepression of platelet-derived growth factor-beta receptor in hepatic stellate expression in hepatic stellate cells, Biochim. Biophys. Acta 1862 (2016) 2186–2196.
cells, J. Cell Sci. 127 (2014) 305–314. [238] J. Pritchett, E. Harvey, V. Athwal, A. Berry, C. Rowe, F. Oakley, et al., Osteopontin is a
[211] J. Qian, M. Niu, X. Zhai, Q. Zhou, Y. Zhou, Beta-catenin pathway is required for TGF- novel downstream target of SOX9 with diagnostic implications for progression of
beta1 inhibition of PPARgamma expression in cultured hepatic stellate cells, liver fibrosis in humans, Hepatology 56 (2012) 1108–1116.
Pharmacol. Res. 66 (2012) 219–225. [239] I. Delgado, M. Carrasco, E. Cano, R. Carmona, R. Garcia-Carbonero, L.M. Marin-
[212] D. Povero, N. Panera, A. Eguchi, C.D. Johnson, B.G. Papouchado, L. de Araujo Horcel, Gomez, et al., GATA4 loss in the septum transversum mesenchyme promotes
et al., Lipid-induced hepatocyte-derived extracellular vesicles regulate hepatic stel- liver fibrosis in mice, Hepatology 59 (2014) 2358–2370.
late cell via microRNAs targeting PPAR-gamma, Cell. Mol. Gastroenterol. Hepatol. 1 [240] E. Wandzioch, A. Kolterud, M. Jacobsson, S.L. Friedman, L. Carlsson, Lhx2−/− mice
(2015) 646–663 (e644). develop liver fibrosis, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 16549–16554.
[213] Y. Jiang, S. Wang, Y. Zhao, C. Lin, F. Zhong, L. Jin, et al., Histone H3K9 demethylase [241] W. Tian, C. Hao, Z. Fan, X. Weng, H. Qin, X. Wu, et al., Myocardin related transcrip-
JMJD1A modulates hepatic stellate cells activation and liver fibrosis by epigeneti- tion factor A programs epigenetic activation of hepatic stellate cells, J. Hepatol. 62
cally regulating peroxisome proliferator-activated receptor gamma, FASEB J. 29 (2015) 165–174.
(2015) 1830–1841. [242] W. Tian, Z. Fan, J. Li, C. Hao, M. Li, H. Xu, et al., Myocardin-related transcription fac-
[214] Y. Zhou, X. Jia, J. Qin, C. Lu, H. Zhu, X. Li, et al., Leptin inhibits PPARgamma gene ex- tor A (MRTF-A) plays an essential role in hepatic stellate cell activation by epige-
pression in hepatic stellate cells in the mouse model of liver damage, Mol. Cell. netically modulating TGF-beta signaling, Int. J. Biochem. Cell Biol. 71 (2016) 35–43.
Endocrinol. 323 (2010) 193–200. [243] K. Palumbo-Zerr, P. Zerr, A. Distler, J. Fliehr, R. Mancuso, J. Huang, et al., Orphan
[215] Q. Zhou, W. Guan, H. Qiao, Y. Cheng, Z. Li, X. Zhai, et al., GATA binding protein 2 me- nuclear receptor NR4A1 regulates transforming growth factor-beta signaling and
diates leptin inhibition of PPARgamma1 expression in hepatic stellate cells and fibrosis, Nat. Med. 21 (2015) 150–158.
T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42 41

[244] P. Chen, J. Li, Y. Huo, J. Lu, L. Wan, Q. Yang, et al., Adenovirus-mediated expression induced inhibition of hepatic stellate cell proliferation, J. Cell. Physiol. 226
of orphan nuclear receptor NR4A2 targeting hepatic stellate cell attenuates liver (2011) 2535–2542.
fibrosis in rats, Sci. Rep. 6 (2016) 33593. [273] X. Sun, Y. He, T.T. Ma, C. Huang, L. Zhang, J. Li, Participation of miR-200a in TGF-
[245] Z. Ghiassi-Nejad, V. Hernandez-Gea, C. Woodrell, U.E. Lang, K. Dumic, A. Kwong, beta1-mediated hepatic stellate cell activation, Mol. Cell. Biochem. 388 (2014)
et al., Reduced hepatic stellate cell expression of Kruppel-like factor 6 tumor sup- 11–23.
pressor isoforms amplifies fibrosis during acute and chronic rodent liver injury, [274] J.J. Yang, H. Tao, L.P. Liu, W. Hu, Z.Y. Deng, J. Li, miR-200a controls hepatic stellate
Hepatology 57 (2013) 786–796. cell activation and fibrosis via SIRT1/Notch1 signal pathway, Inflamm. Res. 66
[246] G. Marrone, L. Russo, E. Rosado, D. Hide, G. Garcia-Cardena, J.C. Garcia-Pagan, et al., (2017) 341–352.
The transcription factor KLF2 mediates hepatic endothelial protection and para- [275] J.J. Yang, H. Tao, W. Hu, L.P. Liu, K.H. Shi, Z.Y. Deng, et al., MicroRNA-200a controls
crine endothelial-stellate cell deactivation induced by statins, J. Hepatol. 58 Nrf2 activation by target Keap1 in hepatic stellate cell proliferation and fibrosis,
(2013) 98–103. Cell. Signal. 26 (2014) 2381–2389.
[247] P. Fernandez-Salguero, T. Pineau, D.M. Hilbert, T. McPhail, S.S. Lee, S. Kimura, et al., [276] J.J. Yang, L.P. Liu, H. Tao, W. Hu, P. Shi, Z.Y. Deng, et al., MeCP2 silencing of LncRNA
Immune system impairment and hepatic fibrosis in mice lacking the dioxin-bind- H19 controls hepatic stellate cell proliferation by targeting IGF1R, Toxicology 359-
ing Ah receptor, Science 268 (1995) 722–726. 360 (2016) 39–46.
[248] F. Andreola, D.F. Calvisi, G. Elizondo, S.B. Jakowlew, J. Mariano, F.J. Gonzalez, et al., [277] E. El-Ahwany, F. Nagy, M. Zoheiry, M. Shemis, M. Nosseir, H.A. Taleb, et al., Circulat-
Reversal of liver fibrosis in aryl hydrocarbon receptor null mice by dietary vitamin ing miRNAs as predictor markers for activation of hepatic stellate cells and pro-
A depletion, Hepatology 39 (2004) 157–166. gression of HCV-induced liver fibrosis, Electron. Physician 8 (2016) 1804–1810.
[249] W.A. Harvey, K. Jurgensen, X. Pu, C.L. Lamb, K.A. Cornell, R.J. Clark, et al., Exposure [278] B.B. Li, D.L. Li, C. Chen, B.H. Liu, C.Y. Xia, H.J. Wu, et al., Potentials of the elevated cir-
to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) increases human hepatic stellate culating miR-185 level as a biomarker for early diagnosis of HBV-related liver fi-
cell activation, Toxicology 344-346 (2016) 26–33. brosis, Sci. Rep. 6 (2016) 34157.
[250] I. Lopez-Sanchez, Y. Dunkel, Y.S. Roh, Y. Mittal, S. De Minicis, A. Muranyi, et al., GIV/ [279] W.J. Wang, R.T. Lai, J. Lu, X.G. Xiang, G.D. Zhao, W.L. Tang, et al., Correlation be-
Girdin is a central hub for profibrogenic signalling networks during liver fibrosis, tween circulating miR-122 and prognosis of chronic HBV-related liver failure, J.
Nat. Commun. 5 (2014) 4451. Dig. Dis. 17 (2016) 334–339.
[251] I. Mannaerts, S.B. Leite, S. Verhulst, S. Claerhout, N. Eysackers, L.F. Thoen, et al., The [280] A. Mallat, J. Lodder, F. Teixeira-Clerc, R. Moreau, P. Codogno, S. Lotersztajn, Autoph-
Hippo pathway effector YAP controls mouse hepatic stellate cell activation, J. agy: a multifaceted partner in liver fibrosis, Biomed. Res. Int. 2014 (2014) 869390.
Hepatol. 63 (2015) 679–688. [281] L.F. Thoen, E.L. Guimaraes, L. Dolle, I. Mannaerts, M. Najimi, E. Sokal, et al., A role for
[252] M.V. Machado, G.A. Michelotti, T.A. Pereira, G. Xie, R. Premont, H. Cortez-Pinto, autophagy during hepatic stellate cell activation, J. Hepatol. 55 (2011) 1353–1360.
et al., Accumulation of duct cells with activated YAP parallels fibrosis progression [282] V. Hernandez-Gea, Z. Ghiassi-Nejad, R. Rozenfeld, R. Gordon, M.I. Fiel, Z. Yue, et al.,
in non-alcoholic fatty liver disease, J. Hepatol. 63 (2015) 962–970. Autophagy releases lipid that promotes fibrogenesis by activated hepatic stellate
[253] S. Nakhaei-Rad, H. Nakhaeizadeh, S. Gotze, C. Kordes, I. Sawitza, M.J. Hoffmann, cells in mice and in human tissues, Gastroenterology 142 (2012) 938–946.
et al., The role of embryonic stem cell-expressed RAS (ERAS) in the maintenance [283] V. Hernandez-Gea, M. Hilscher, R. Rozenfeld, M.P. Lim, N. Nieto, S. Werner, et al.,
of quiescent hepatic stellate cells, J. Biol. Chem. 291 (2016) 8399–8413. Endoplasmic reticulum stress induces fibrogenic activity in hepatic stellate cells
[254] T. Hardy, D.A. Mann, Epigenetics in liver disease: from biology to therapeutics, Gut through autophagy, J. Hepatol. 59 (2013) 98–104.
(2016). [284] J.H. Koo, H.J. Lee, W. Kim, S.G. Kim, Endoplasmic reticulum stress in hepatic stellate
[255] A. Page, P. Paoli, E. Moran Salvador, S. White, J. French, J. Mann, Hepatic stellate cell cells promotes liver fibrosis via PERK-mediated degradation of HNRNPA1 and up-
transdifferentiation involves genome-wide remodeling of the DNA methylation regulation of SMAD2, Gastroenterology 150 (2016) 181–193 (e188).
landscape, J. Hepatol. 64 (2016) 661–673. [285] R.S. Kim, D. Hasegawa, N. Goossens, T. Tsuchida, V. Athwal, X. Sun, et al., The XBP1
[256] J. Mann, D.C. Chu, A. Maxwell, F. Oakley, N.L. Zhu, H. Tsukamoto, et al., MeCP2 con- arm of the unfolded protein response induces fibrogenic activity in hepatic stellate
trols an epigenetic pathway that promotes myofibroblast transdifferentiation and cells through autophagy, Sci. Rep. 6 (2016) 39342.
fibrosis, Gastroenterology 138 (2010) 705–714 (714 e701-704). [286] F. Heindryckx, F. Binet, M. Ponticos, K. Rombouts, J. Lau, J. Kreuger, et al., Endoplas-
[257] J. Mann, F. Oakley, F. Akiboye, A. Elsharkawy, A.W. Thorne, D.A. Mann, Regulation mic reticulum stress enhances fibrosis through IRE1alpha-mediated degradation of
of myofibroblast transdifferentiation by DNA methylation and MeCP2: implica- miR-150 and XBP-1 splicing, EMBO Mol. Med. 8 (2016) 729–744.
tions for wound healing and fibrogenesis, Cell Death Differ. 14 (2007) 275–285. [287] Y. He, J. Zhu, Y. Huang, H. Gao, Y. Zhao, Advanced glycation end product (AGE)-in-
[258] A. El Taghdouini, A.L. Sorensen, A.H. Reiner, M. Coll, S. Verhulst, I. Mannaerts, et al., duced hepatic stellate cell activation via autophagy contributes to hepatitis C-relat-
Genome-wide analysis of DNA methylation and gene expression patterns in puri- ed fibrosis, Acta Diabetol. 52 (2015) 959–969.
fied, uncultured human liver cells and activated hepatic stellate cells, Oncotarget 6 [288] L. Zhan, Y. Yang, T.T. Ma, C. Huang, X.M. Meng, L. Zhang, et al., Transient receptor
(2015) 26729–26745. potential vanilloid 4 inhibits rat HSC-T6 apoptosis through induction of autophagy,
[259] E.B. Bian, C. Huang, T.T. Ma, H. Tao, H. Zhang, C. Cheng, et al., DNMT1-mediated Mol. Cell. Biochem. 402 (2015) 9–22.
PTEN hypermethylation confers hepatic stellate cell activation and liver [289] K. Kawasaki, R. Ushioda, S. Ito, K. Ikeda, Y. Masago, K. Nagata, Deletion of the colla-
fibrogenesis in rats, Toxicol. Appl. Pharmacol. 264 (2012) 13–22. gen-specific molecular chaperone Hsp47 causes endoplasmic reticulum stress-me-
[260] M.J. Perugorria, C.L. Wilson, M. Zeybel, M. Walsh, S. Amin, S. Robinson, et al., His- diated apoptosis of hepatic stellate cells, J. Biol. Chem. 290 (2015) 3639–3646.
tone methyltransferase ASH1 orchestrates fibrogenic gene transcription during [290] Y. Sato, K. Murase, J. Kato, M. Kobune, T. Sato, Y. Kawano, et al., Resolution of liver
myofibroblast transdifferentiation, Hepatology 56 (2012) 1129–1139. cirrhosis using vitamin A-coupled liposomes to deliver siRNA against a collagen-
[261] R.L. Pan, L.X. Xiang, P. Wang, X.Y. Liu, L. Nie, W. Huang, et al., Low-molecular- specific chaperone, Nat. Biotechnol. 26 (2008) 431–442.
weight fibroblast growth factor 2 attenuates hepatic fibrosis by epigenetic [291] D. Zhang, T. Utsumi, H.C. Huang, L. Gao, P. Sangwung, C. Chung, et al., Reticulon 4B
down-regulation of Delta-like1, Hepatology 61 (2015) 1708–1720. (Nogo-B) is a novel regulator of hepatic fibrosis, Hepatology 53 (2011) 1306–1315.
[262] N. Ding, N. Hah, R.T. Yu, M.H. Sherman, C. Benner, M. Leblanc, et al., BRD4 is a novel [292] R. Men, M. Wen, X. Dan, Y. Zhu, W. Wang, J. Li, et al., Nogo-B: a potential indicator
therapeutic target for liver fibrosis, Proc. Natl. Acad. Sci. U. S. A. 112 (2015) for hepatic cirrhosis and regulator in hepatic stellate cell activation, Hepatol. Res.
15713–15718. 45 (2015) 113–122.
[263] T. Hardy, M. Zeybel, C.P. Day, C. Dipper, S. Masson, S. McPherson, et al., Plasma DNA [293] J. Deng, Q. Huang, Y. Wang, P. Shen, F. Guan, J. Li, et al., Hypoxia-inducible factor-
methylation: a potential biomarker for stratification of liver fibrosis in non- 1alpha regulates autophagy to activate hepatic stellate cells, Biochem. Biophys.
alcoholic fatty liver disease, Gut (2016) pii: gutjnl-2016-311526. Res. Commun. 454 (2014) 328–334.
[264] A. Page, P.P. Paoli, S.J. Hill, R. Howarth, R. Wu, S.M. Kweon, et al., Alcohol directly [294] J. Zhao, L. Peng, R. Cui, X. Guo, M. Yan, Dimethyl alpha-ketoglutarate reduces CCl4-
stimulates epigenetic modifications in hepatic stellate cells, J. Hepatol. 62 (2015) induced liver fibrosis through inhibition of autophagy in hepatic stellate cells,
388–397. Biochem. Biophys. Res. Commun. 481 (2016) 90–96.
[265] J.S. Kim, S.D. Shukla, Histone h3 modifications in rat hepatic stellate cells by etha- [295] X.W. Zhang, S. Mi, Z. Li, J.C. Zhou, J. Xie, F. Hua, et al., Antagonism of interleukin-17A
nol, Alcohol Alcohol. 40 (2005) 367–372. ameliorates experimental hepatic fibrosis by restoring the IL-10/STAT3-suppressed
[266] M. Zeybel, T. Hardy, Y.K. Wong, J.C. Mathers, C.R. Fox, A. Gackowska, et al., Multi- autophagy in hepatocytes, Oncotarget 8 (2017) 9922–9934.
generational epigenetic adaptation of the hepatic wound-healing response, Nat. [296] S. Jiang, Y. Zhang, J.H. Zheng, X. Li, Y.L. Yao, Y.L. Wu, et al., Potentiation of hepatic
Med. 18 (2012) 1369–1377. stellate cell activation by extracellular ATP is dependent on P2X7R-mediated
[267] M. Kitano, P.M. Bloomston, Hepatic stellate cells and microRNAs in pathogenesis of NLRP3 inflammasome activation, Pharmacol. Res. 117 (2016) 82–93.
liver fibrosis, J. Clin. Med. 5 (2016). [297] M. Okuno, H. Moriwaki, S. Imai, Y. Muto, N. Kawada, Y. Suzuki, et al., Retinoids ex-
[268] F. Yu, Z. Lu, K. Huang, X. Wang, Z. Xu, B. Chen, et al., MicroRNA-17-5p-activated acerbate rat liver fibrosis by inducing the activation of latent TGF-beta in liver stel-
Wnt/beta-catenin pathway contributes to the progression of liver fibrosis, late cells, Hepatology 26 (1997) 913–921.
Oncotarget 7 (2016) 81–93. [298] H.S. Yi, Y.S. Lee, J.S. Byun, W. Seo, J.M. Jeong, O. Park, et al., Alcohol dehydrogenase
[269] M. Coll, A. El Taghdouini, L. Perea, I. Mannaerts, M. Vila-Casadesus, D. Blaya, et al., III exacerbates liver fibrosis by enhancing stellate cell activation and suppressing
Integrative miRNA and gene expression profiling analysis of human quiescent he- natural killer cells in mice, Hepatology 60 (2014) 1044–1053.
patic stellate cells, Sci. Rep. 5 (2015) 11549. [299] J. Kluwe, N. Wongsiriroj, J.S. Troeger, G.Y. Gwak, D.H. Dapito, J.P. Pradere, et al., Ab-
[270] J. Hyun, S. Wang, J. Kim, K.M. Rao, S.Y. Park, I. Chung, et al., MicroRNA-378 limits sence of hepatic stellate cell retinoid lipid droplets does not enhance hepatic fibro-
activation of hepatic stellate cells and liver fibrosis by suppressing Gli3 expression, sis but decreases hepatic carcinogenesis, Gut 60 (2011) 1260–1268.
Nat. Commun. 7 (2016) 10993. [300] A. Yoneda, K. Sakai-Sawada, Y. Niitsu, Y. Tamura, Vitamin A and insulin are re-
[271] D.D. Zhou, X. Wang, Y. Wang, X.J. Xiang, Z.C. Liang, Y. Zhou, et al., MicroRNA-145 quired for the maintenance of hepatic stellate cell quiescence, Exp. Cell Res. 341
inhibits hepatic stellate cell activation and proliferation by targeting ZEB2 through (2016) 8–17.
Wnt/beta-catenin pathway, Mol. Immunol. 75 (2016) 151–160. [301] S.E. Trasino, X.H. Tang, J. Jessurun, L.J. Gudas, A retinoic acid receptor beta2 agonist
[272] Y. Sekiya, T. Ogawa, M. Iizuka, K. Yoshizato, K. Ikeda, N. Kawada, Down-regula- reduces hepatic stellate cell activation in nonalcoholic fatty liver disease, J. Mol.
tion of cyclin E1 expression by microRNA-195 accounts for interferon-beta- Med. (Berl.) 94 (2016) 1143–1151.
42 T. Higashi et al. / Advanced Drug Delivery Reviews 121 (2017) 27–42

[302] T.O. Eichmann, L. Grumet, U. Taschler, J. Hartler, C. Heier, A. Woblistin, et al., ATGL [315] S.T. Rashid, J.D. Humphries, A. Byron, A. Dhar, J.A. Askari, J.N. Selley, et al., Proteomic
and CGI-58 are lipid droplet proteins of the hepatic stellate cell line HSC-T6, J. Lipid analysis of extracellular matrix from the hepatic stellate cell line LX-2 identifies
Res. 56 (2015) 1972–1984. CYR61 and Wnt-5a as novel constituents of fibrotic liver, J. Proteome Res. 11
[303] F. O'Mahony, K. Wroblewski, S.M. O'Byrne, H. Jiang, K. Clerkin, J. Benhammou, et al., (2012) 4052–4064.
Liver X receptors balance lipid stores in hepatic stellate cells through Rab18, a ret- [316] J. Zhang, M. Wang, Z. Zhang, Z. Luo, F. Liu, J. Liu, Celecoxib derivative OSU-03012
inoid responsive lipid droplet protein, Hepatology 62 (2015) 615–626. inhibits the proliferation and activation of hepatic stellate cells by inducing cell se-
[304] Z. Zhang, S. Zhao, Z. Yao, L. Wang, J. Shao, A. Chen, et al., Autophagy regulates turn- nescence, Mol. Med. Rep. 11 (2015) 3021–3026.
over of lipid droplets via ROS-dependent Rab25 activation in hepatic stellate cell, [317] Y. Duan, J. Pan, J. Chen, D. Zhu, J. Wang, X. Sun, et al., Soluble egg antigens of
Redox Biol. 11 (2016) 322–334. Schistosoma japonicum induce senescence of activated hepatic stellate cells by ac-
[305] F.V. Bruschi, T. Claudel, M. Tardelli, A. Caligiuri, T.M. Stulnig, F. Marra, et al., The tivation of the FoxO3a/SKP2/P27 pathway, PLoS Negl. Trop. Dis. 10 (2016),
PNPLA3 I148M variant modulates the fibrogenic phenotype of human hepatic stel- e0005268.
late cells, Hepatology (2017) http://dx.doi.org/10.1002/hep.29041. [318] C. Panebianco, J.A. Oben, M. Vinciguerra, V. Pazienza, Senescence in hepatic stellate
[306] C.L. Wilson, L.B. Murphy, J. Leslie, S. Kendrick, J. French, C.R. Fox, et al., Ubiquitin C- cells as a mechanism of liver fibrosis reversal: a putative synergy between retinoic
terminal hydrolase 1: a novel functional marker for liver myofibroblasts and a acid and PPAR-gamma signalings, Clin. Exp. Med. (2016).
therapeutic target in chronic liver disease, J. Hepatol. 63 (2015) 1421–1428. [319] A. Anan, E.S. Baskin-Bey, S.F. Bronk, N.W. Werneburg, V.H. Shah, G.J. Gores, Protea-
[307] S. Arndt, E. Wacker, C. Dorn, A. Koch, M. Saugspier, W.E. Thasler, et al., Enhanced some inhibition induces hepatic stellate cell apoptosis, Hepatology 43 (2006)
expression of BMP6 inhibits hepatic fibrosis in non-alcoholic fatty liver disease, 335–344.
Gut 64 (2015) 973–981. [320] B. Gao, S. Radaeva, O. Park, Liver natural killer and natural killer T cells:
[308] T. Kisseleva, M. Cong, Y. Paik, D. Scholten, C. Jiang, C. Benner, et al., Myofibroblasts immunobiology and emerging roles in liver diseases, J. Leukoc. Biol. 86 (2009)
revert to an inactive phenotype during regression of liver fibrosis, Proc. Natl. Acad. 513–528.
Sci. U. S. A. 109 (2012) 9448–9453. [321] R. Bansal, J. Prakash, E. Post, L. Beljaars, D. Schuppan, K. Poelstra, Novel engineered
[309] J.S. Troeger, I. Mederacke, G.Y. Gwak, D.H. Dapito, X. Mu, C.C. Hsu, et al., Deactiva- targeted interferon-gamma blocks hepatic fibrogenesis in mice, Hepatology 54
tion of hepatic stellate cells during liver fibrosis resolution in mice, Gastroenterol- (2011) 586–596.
ogy 143 (2012) 1073–1083 (e1022). [322] M.E. Shaker, A. Ghani, G.E. Shiha, T.M. Ibrahim, W.Z. Mehal, Nilotinib induces apo-
[310] G. Song, M. Pacher, A. Balakrishnan, Q. Yuan, H.C. Tsay, D. Yang, et al., Direct ptosis and autophagic cell death of activated hepatic stellate cells via inhibition of
reprogramming of hepatic myofibroblasts into hepatocytes in vivo attenuates histone deacetylases, Biochim. Biophys. Acta 1833 (2013) 1992–2003.
liver fibrosis, Cell Stem Cell 18 (2016) 797–808. [323] H. Hao, D. Zhang, J. Shi, Y. Wang, L. Chen, Y. Guo, et al., Sorafenib induces autoph-
[311] B. Schnabl, C.A. Purbeck, Y.H. Choi, C.H. Hagedorn, D. Brenner, Replicative senes- agic cell death and apoptosis in hepatic stellate cell through the JNK and Akt signal-
cence of activated human hepatic stellate cells is accompanied by a pronounced in- ing pathways, Anti-Cancer Drugs 27 (2016) 192–203.
flammatory but less fibrogenic phenotype, Hepatology 37 (2003) 653–664. [324] R. Weiskirchen, F. Tacke, Liver fibrosis: from pathogenesis to novel therapies, Dig.
[312] V. Krizhanovsky, M. Yon, R.A. Dickins, S. Hearn, J. Simon, C. Miething, et al., Se- Dis. 34 (2016) 410–422.
nescence of activated stellate cells limits liver fibrosis, Cell 134 (2008) [325] Y. Taniguchi, P.J. Choi, G.W. Li, H. Chen, M. Babu, J. Hearn, et al., Quantifying E. coli
657–667. proteome and transcriptome with single-molecule sensitivity in single cells, Sci-
[313] E. Borkham-Kamphorst, C. Schaffrath, E. Van de Leur, U. Haas, L. Tihaa, S.K. Meurer, ence 329 (2010) 533–538.
et al., The anti-fibrotic effects of CCN1/CYR61 in primary portal myofibroblasts are [326] A. Dixit, O. Parnas, B. Li, J. Chen, C.P. Fulco, L. Jerby-Arnon, et al., Perturb-Seq: dis-
mediated through induction of reactive oxygen species resulting in cellular senes- secting molecular circuits with scalable single-cell RNA profiling of pooled genetic
cence, apoptosis and attenuated TGF-beta signaling, Biochim. Biophys. Acta 1843 screens, Cell 167 (2016) 1853–1866 (e1817).
(2014) 902–914. [327] R. Satija, J.A. Farrell, D. Gennert, A.F. Schier, A. Regev, Spatial reconstruction of sin-
[314] S. Klein, J. Klosel, R. Schierwagen, C. Korner, M. Granzow, S. Huss, et al., Atorvastatin gle-cell gene expression data, Nat. Biotechnol. 33 (2015) 495–502.
inhibits proliferation and apoptosis, but induces senescence in hepatic [328] L.A. van Grunsven, 3D in vitro models of liver fibrosis, Adv. Drug Deliv. Rev. (2017).
myofibroblasts and thereby attenuates hepatic fibrosis in rats, Lab. Investig. 92 [329] S. Crespo Yanguas, B. Cogliati, J. Willebrords, M. Maes, I. Colle, B. van den Bossche,
(2012) 1440–1450. et al., Experimental models of liver fibrosis, Arch. Toxicol. 90 (2016) 1025–1048.

You might also like