You are on page 1of 20

Hepatic Stellate Cells and Liver Fibrosis

Juan E. Puche,1,2 Yedidya Saiman,1 and Scott L. Friedman*1

ABSTRACT
Hepatic stellate cells are resident perisinusoidal cells distributed throughout the liver, with a
remarkable range of functions in normal and injured liver. Derived embryologically from sep-
tum transversum mesenchyme, their precursors include submesothelial cells that invade the liver
parenchyma from the hepatic capsule. In normal adult liver, their most characteristic feature is
the presence of cytoplasmic perinuclear droplets that are laden with retinyl (vitamin A) esters.
Normal stellate cells display several patterns of intermediate filaments expression (e.g., desmin,
vimentin, and/or glial fibrillary acidic protein) suggesting that there are subpopulations within
this parental cell type. In the normal liver, stellate cells participate in retinoid storage, vasoregu-
lation through endothelial cell interactions, extracellular matrix homeostasis, drug detoxification,
immunotolerance, and possibly the preservation of hepatocyte mass through secretion of mi-
togens including hepatocyte growth factor. During liver injury, stellate cells activate into alpha
smooth muscle actin-expressing contractile myofibroblasts, which contribute to vascular distor-
tion and increased vascular resistance, thereby promoting portal hypertension. Other features of
stellate cell activation include mitogen-mediated proliferation, increased fibrogenesis driven by
connective tissue growth factor, and transforming growth factor beta 1, amplified inflammation
and immunoregulation, and altered matrix degradation. Evolving areas of interest in stellate cell
biology seek to understand mechanisms of their clearance during fibrosis resolution by either
apoptosis, senescence, or reversion, and their contribution to hepatic stem cell amplification, re-
generation, and hepatocellular cancer.  C 2013 American Physiological Society. Compr Physiol

3:1473-1492, 2013.

Introduction to one that is proliferative fibrogenic and contractile (62, 82,


83, 234, 241, 289). While HSCs are a major source of myofi-
Liver fibrosis is a common outcome of virtually all chronic broblasts, mounting evidence also implicates portal fibrob-
hepatic insults including viral hepatitis (i.e., hepatitis B lasts as a source when the injury is to bile ducts rather than
and C), alcoholic or obesity-associated steatohepatitis (i.e., hepatocytes (53, 162).
nonalcoholic steatohepatitis (NASH)), parasitic disease From early studies focusing on the role of the stellate
(i.e., schistosomiasis), metabolic disorders (i.e., Wilson’s), cells solely as a source of ECM during liver injury, a sus-
hemochromatosis and other storage diseases, congenital tained effort has subsequently uncovered broadening roles
abnormalities, autoimmune and chronic inflammatory condi- of the cell type as a source of regenerative cytokines, an
tions (i.e., sarcoidosis), and drug toxicity, among others (102). immunomodulatory cell with a range of activities and one
Fibrosis was described in 1951 as a passive process with that serves roles far beyond those envisioned at the time of its
“no direct evidence of fibrous tissue proliferation but the sug- isolation 35 years ago (65). From these studies realistic hopes
gestion of connective tissue appearance just by stromal con- have emerged for exploiting features of stellate cell activation
densation due to hepatocyte cell collapse.” However, in 1978, and hepatic inflammation in devising effective antifibrotic and
a growing body of evidence prompted the World Health Orga- regenerative therapies for patients with chronic liver disease
nization to update its definition of fibrosis as “the presence of and fibrosis.
excess collagen due to new fiber formation (3).” It is virtually This article builds upon a comprehensive review of stel-
axiomatic now that liver fibrosis results from enhanced pro- late cell biology in 2008 by one of the authors (S.L.F.) (65),
duction of extracellular matrix (ECM) due to accumulation
and activation of myofibroblasts in the context of ongoing
or repetitive liver damage. Early studies focused on meth-
* Correspondence to scott.friedman@mssm.edu
ods to isolate and grow primary hepatic stellate cells (HSCs)
1 Division of Liver Diseases, Icahn School of Medicine at Mount Sinai
in culture establishing them as one of the main sources of
Hospital, New York, New York
myofibroblasts in liver parenchymal disease resulting from 2 University CEU-San Pablo, School of Medicine, Institute of Applied
hepatocyte as oppsed to biliary injury (49, 71). Immediately Molecular Medicine (IMMA), Madrid, Spain
thereafter, the concept of stelllate cell “activation” emerged, Published online, October 2013 (comprehensivephysiology.com)
which represents a transdifferentiation of the cell during liver DOI: 10.1002/cphy.c120035
injury from a quiescent state that is rich in vitamin A droplets, Copyright  C American Physiological Society.

Volume 3, October 2013 1473


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

Figure 1 Appearance of hepatic stellate cells and the sinusoidal microenvironment in normal and injured
liver. In normal liver, stellate cells (shown in blue) are laden with perinuclear retinoid droplets and preserve
the differentiated function of surrounding cells, including hepatocytes and sinusoidal endothelial cells. In liver
injury, the cells multiply, lose vitamin A and become embedded within dense extracellular matrix. This leads
to deterioration of hepatocyte function manifested as loss of microvilli, and decreased size and number of
endothelial fenestrations. Reprinted, with permission, from (68).

by highlighting the accelerating pace of progress and increas- endothelial cells (LSECs) while the outer surface, facing
ingly nuanced understanding of a cell type that is unique not the space of Disse, has numerous micro-projections which
only in liver, but throughout mammalian biology. The fas- contact with hepatocytes and may function in detecting
cination with stellate cells may explain why the published chemotactic signals, and transmitting them to the cell’s
literature contains more than 2800 articles about this cell type mechanical apparatus to generate a contractile force that
only in the last ten years (Pubmed search using the keyword regulates blood flow (180). Actin filaments and microtubules
“hepatic stellate cell” from 2002-2012). are distributed in both the periphery and the core of the cell’s
processes, respectively, and could be responsible for their
extension and retraction (115, 148, 204, 246, 252). While no
Hepatic Stellate Cell Biology, Origin electron dense basement membrane can be identified within
the perisinusoidal space of Disse, basement membrane
and Ultrastructure proteins, including type IV collagen, nidogen, and laminin,
HSCs are resident nonparenchymal cells located in the suben- have been identified, which are thought be functionally
dothelial space of Disse, between the basolateral surface of important in preserving differentiated hepatocellular function
hepatocytes and the antiluminal side of sinusoidal endothe- and stellate cell quiescence (20, 29, 70).
lial cells (Fig. 1). This privileged location, together with their HSCs were first described by Kupffer in 1876 (303), using
dendritic cytoplasmic processes, facilitates their direct con- a gold chloride method for detection of neuronal components
tact with hepatocytes, endothelial cells, other stellate cells, in the liver. Their star-shaped characteristics led Kupffer to
and Kupffer cells up to 140 μm away (86, 126). This intimate call them “sternzellen” (“star cell,” in German). However, in
contact between stellate cells and their neighboring cell types 1898, a misinterpretation based on India ink staining led him
may facilitate intercellular transport of soluble mediators and to conclude that they were actually “special endothelial cells
cytokines. In addition, stellate cells are directly adjacent to of the sinusoids with phagocytic capacity,” thereby confusing
nerve endings (19, 299), which is consistent with reports them with macrophages (now often called “Kupffer cells”
identifying neurotrophin receptors (36, 137, 284, 333), and (304)). Ironically, recent studies have indeed established that
with functional studies confirming neurohumoral responsive- stellate cells have phagocytic properties (35, 129, 190, 301).
ness of stellate cells (158, 203, 249). Interestingly, apart from Since Kupffer first identified stellate cells nearly 150
the different patterns of distribution (pericentral vs. periportal years ago (303), a number of new techniques for their detec-
predominances) among species (28, 65, 305), stellate cells tion and isolation have been developed based on their lipid
only represent ∼10% of the total number of resident cells in droplet content, cytoskeletal features, and cell surface mark-
normal liver (86) and ∼1.5% of total liver volume, a low pro- ers (288). These approaches have aided in further defining
portion of the total cell number in contrast to their remarkably their ultrastructure and enhancing their purity during isola-
divergent functions in normal and diseased liver. tion, for example, by using vitamin A fluorescence to isolate
Prominent dendritic cytoplasmic processes from stellate the cells using flow cytometry (45). In normal liver, stel-
cells contact hepatocytes and endothelial cells (65, 266, late cells have spindle-shaped cell bodies with perikaryons
302, 304). These subendothelial processes have three cell within recesses between neighboring parenchymal cells, with
surfaces: inner, outer, and lateral. The inner one is smooth and oval or elongated nuclei. Vitamin A lipid droplets are con-
adheres to the adluminal (basal) surface of the liver sinusoidal spicuous features of the cytoplasm. Stellate cells also have

1474 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

well-developed juxtanuclear small Golgi complex and rough Extrahepatic stellate cells have also been described
endoplasmic reticulum spaces indicating active biosynthe- throughout many organs in particular pancreas, where they
sis of secreted polypeptides or proteins (56). The presence retain their characteristic shape and markers (267). Stellate
of active lysosomes in stellate cell cytoplasm described for cells in other tissues typically store vitamin A and synthe-
decades (33, 129) is now recognized to indicate a high capac- size and secrete ECM components. Best characterized among
ity for the cells to undergo autophagy during cellular acti- these are pancreatic stellate cells, which clearly contribute
vation (104, 105). Endosomes and multivesicular bodies are to pancreatic fibrosis and tumor stroma (4, 5, 57). Pancre-
also present in HSCs (267) and contribute to the generation atic stellate cells reportedly generate stem cells that may also
of vitamin A-containing lipid droplets. directly contribute to liver regeneration through differentia-
In contrast to the well-established origins of other liver tion into hepatocytes and duct-forming cholangiocytes across
populations (endoderm-derived hepatocytes and cholangio- tissue boundaries, but this observation requires confirmation
cytes, mesoderm-derived endothelial cells and fibroblasts, (153).
and ectoderm-derived neurons), the ontogeny of this enig-
matic cell has been controversial for decades, because the
cells express a pattern of cytoskeletal markers reflecting a Role of Hepatic Stellate Cells in
range of origins including ectoderm (e.g., glial fibrillary acidic Normal Liver
protein (GFAP), nestin, neurotrophins and their receptors,
nerve growth factor (NGF), brain-derived neurotrophic factor, Because of their recognized role in hepatic fibrosis, most
synaptophysin, and N-CAM) and mesoderm (e.g., vimentin, studies of HSCs have focused primarily on their behavior
desmin, alpha smooth muscle actin, hematopoietic markers) during liver injury, but have neglected their contribution to
(80, 186). normal liver homeostasis. With an increased availability of
More recently, elegant developmental studies have estab- tools to selectively express transgenes in stellate cells, their
lished that stellate cells are traced to mesothelial cells role in normal liver development and function is now being
(likely derived from septum transversum mesenchyme), illuminated (Fig. 2).
which appear to give rise to cells that invade the hepatic
bud (8, 11, 165). However, due to the limited labeling effi- Contribution of stellate cells to
ciency of the mesothelium, the proportion of stellate cells liver development
derived from the mesothelium is not known and other cel-
lular sources may also be possible. There is ample evidence As noted above, stellate cells can be identified within the
that HSCs are remarkably heterogeneous in their content of progenitor cell niche (near the canals of Hering) in normal,
retinoid, cytoskeletal phenotype, potential for activation, and developing, and regenerating liver (248, 326). Additionally,
even their capacity to revert to a quiescent state after liver murine fetal liver-derived Thy1+ cells, which express classi-
injury resolves (45, 80, 143, 170). For example, there is a cal markers of HSCs (α-SMA, desmin, and vimentin), pro-
subpopulation of stellate cells that may lack typical cytoskele- mote maturation of hepatic progenitors through cell-cell con-
tal markers (16, 235, 257) depending on the lobular location. tact in culture (12). Pleiotrophin, a morphogen secreted by
Pericentral areas are rich in stellate cells with longer cytoplas- stellate cell precursors (ALCAM+ submesothelial cells) dur-
mic processes (i.e., more astrocytic morphology), predomi- ing liver development, may contribute to liver organogenesis
nant GFAP expression (instead of desmin), decreased number and regeneration (9, 10). Quiescent stellate cells also express
and size of lipid droplets, and more differentiated. Peripor- epimorphin, a mesenchymal morphogenic protein involved
tal stellate cells are typically desmin positive with shorter
cytoplasmic processes (with a more contractile phenotype),
Extracellular matrix
contain more and larger lipid droplets and may be less differ- homeostasis
entiated (82, 195, 304, 305). Vasoregulation Normal liver
While the origin of stellate cells is becoming less con- development
troversial, still uncertain is the possibility that stellate cells
are pluripotent and can give rise to multiple cell lineages.
This phenomenon has been reported in at least two studies
(154, 324), but based only on cell culture findings, where
modulation of cellular phenotype is notoriously promiscuous
Quiescent
and may not reflect events in vivo. On the other hand, it is Preservation of stellate cell
increasingly clear that HSCs are intimately associated with hepatocyte mass Retinoid metabolism
the progenitor cell niche and typically surround cells that
have stem cell-like properties (39, 40, 89, 219, 311). These
Drug metabolism
findings indicate a strong likelihood that stellate cells support and detoxification
stem cell expansion, although underlying mechanisms and
mediators that drive this interaction are not known. Figure 2 Roles of hepatic stellate cells in normal liver.

Volume 3, October 2013 1475


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

in differentiation of rat hepatic stem-like cells by a puta- collagens (type I, III, IV, V, VI, XIV, and XVIII), elastin,
tive epithelial-mesenchymal contact that promotes bile duct structural glycoproteins (laminin, fibronectin, nidogen/
epithelial morphogenesis (184), which involves the RhoA and entactin, tenascin, osteopontin, and secreted acidic proteins
C/EBPβ pathways. These findings complement evidence of rich in cysteine), proteoglycans (heparan sulfate, chondroitin
paracrine interactions between bile duct epithelium and either 4-sulfate, chondroitin 6-sulfate, and dermatan sulfate, synde-
stellate cells or portal fibroblasts both in culture (156, 166) can, biglycan, and decorin), and the free glycosaminoglycan
and in vivo (141, 156, 285). While not limited to stellate cell- hyaluronan (74, 136, 259). While quantitatively modest in the
bile duct crosstalk, components of the Notch (254) and Wnt latter location, ECM in the space of Disse has an important
pathways (178, 330), purinergic signaling (52), chemokines role in preserving liver homeostasis and has a unique spa-
(156) and the Dlk1 protein (298, 330) are also important in tial expression pattern. Type IV collagen is primarily located
hepatic development. Stellate cell precursors, isolated from between LSECs and stellate cells while type I and III fibrillar
fetal liver based on UV fluorescence in flow cytometry (157), collagens are located between HCSs and hepatocytes in nor-
display extensive proliferative activity, and a high capacity mal liver (20, 70). Primarily, the ECM composition can affect
to express hepatocyte growth factor (HGF), CXCL12, and the behavior of surrounding liver cells through cell surface
homeobox transcription factors, supporting their potential receptors, especially integrins, of which stellate cells express
contributions to both hepatic development and hematopoiesis α1β1, α1β2 (227), αvβ6 (222), α5 β1 (110), αvβ6 (216,
(157). A recent study has identified stellate cells in zebrafish 222), as well as integrin linked kinase (269).
using a reporter gene driven by the Hand2 promoter (325). The three cell types surrounding the space of Disse (hep-
Use of this model will further clarify the stellate cell’s contri- atocytes, endothelial cells, and stellate cells) each produce
butions to hepatic development and homeostasis. ECM components in normal liver. While all of them express
collagen type I, hepatocytes mainly produce fibronectin (233),
endothelial cells express collagen IV, and quiescent stellate
Retinoid metabolism (see section on cells secrete laminin and collagen types III and IV (83, 171),
Perpetuation, Section V, below) among several other ECM proteins.
Vitamin A (retinoid) is primarily stored in the liver in The maintenance of ECM homeostasis requires turnover
mammals, and among liver cell types stellate cells are the in which production of new components is offset by parallel
primary cellular depot (303). Normally, dietary retinoids are rates of degradation. Matrix metalloproteinases (MMPs) are
absorbed by the gut and transported in chylomicron remnants the primary effectors of ECM degradation, whose activity is
as retinyl esters to hepatocytes, where they are hydrolyzed regulated in turn by tissue inhibitors of metalloproteinases
into free retinol. Retinol is then transferred to stellate (TIMPs) (6, 145). Several liver cell populations (i.e., Kupf-
cells, where they are reesterified (101). Importantly, these fer cells, myofibroblast, and hepatocytes) can produce both
droplets contain not only retinoids, but also triglycerides, MMPs and TIMPs (6, 145), however a subgroup of this fam-
phospholipids, cholesterol, and free fatty acids, among others ily, the A Disintegrin and Metalloproteinase-domain proteins
(188, 320). Recent studies have identified a family of proteins (ADAMs) may elude TIMP action and contribute to trans-
that coat lipid droplets known as perilipins (27, 28, 253). One forming growth factor beta (TGF-β) activation (26), the most
perilipin, adipose-differentiation-related protein, is expressed potent stimulus for collagen I production by stellate cells (91,
by stellate cells and its levels are reduced as the cells activate 116, 296). While most ADAMs are expressed by more than
and lose retinoid droplets (161). The contributions of these one liver cell type, at least two (ADAM 13 and 28) are pro-
lipid droplets go beyond the simple storage of Vitamin A, and duced solely by stellate cells (194, 261, 273, 314).
extend to the regulation of stellate cell activation (206, 207),
possibly through the impact of lipids in fueling autophagy
(103, 104). The importance of these retinoid mechanisms Secretion of mediators
in fibrosis however has been challenged, however, as mice While stellate cell-derived molecules are a major driving
deficient in lecithin retinol acetyl transferase (LRAT), the force in hepatic fibrosis they may also play an important
enzyme that catalyzes the esterification of retinol into retinyl role in preserving liver homeostasis and promoting regener-
esters nonetheless undergo fibrosis (144) following toxic ation, although the data do not fully support such a role yet.
liver injury, perhaps indicating an alternate or more complex The specific spatial and temporal expression patterns of these
role for retinoid metabolism in hepatic and stellate cell molecules may therefore be important to promoting proper
homeostasis. hepatic development and regeneration after injury. In steady
state conditions, stellate cells are reported to secrete a range
of molecules detailed in the following sections.
Extracellular matrix homeostasis
In normal liver, the ECM comprises ∼0.5% of the total liver
Growth factors
weight (245) and is distributed between portal triads, central
veins and Glisson’s capsule, with only a small portion present HGF is the most potent mitogen for hepatocytes (256). Quies-
in the space of Disse (81). Normal ECM components include cent stellate cells can produce HGF, but interestingly, during

1476 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

liver damage the expression of this growth factor is downreg- cells. A carefully regulated pathway exists for activation of
ulated in stellate cells by the action of transforming growth latent endothelin in stellate cells (138, 164).
factor β (TGF-β) (232). This temporal expression of HGF Human stellate cells have also functional receptors for
may, therefore, explain the decreased rate of hepatic regener- adrenomedullin (ADM), a peptide produced by most contrac-
ation in a fibrotic/injured liver. tile cells, which modulates the contractile effect of ET-1 (88).
TGF-β is among most potent cytokines that regulate stel- Moreover, cultured human stellate cells secrete ADM in base-
late cell phenotype (21, 51). In normal liver TGF-β iso- line conditions, and its production is markedly increased by
form expression (TGF-β1,2,3 ) is shared between hepatocytes, cytokines (88), a surprising finding given that stellate cell acti-
Kupffer cells and stellate cells (21). While TGF-β1 is more vation promotes cellular contraction. ADM can also attenuate
highly expressed by Kupffer cells than HSCs, TGF-β3 is only activation of stellate cells by inhibiting TGF-β1 production
expressed by stellate cells (48). Regardless, TGF-β is secreted and TGF-β-induced MMP-2 expression partially through the
in its latent form, and requires a further activation of the latent ERK pathway (312). These results suggest that ADM regu-
molecule to exert its action. The predominant pathways of lates stellate cell activation and contractility in an autocrine
TGF-β activation diverge among different tissues. In liver, manner.
integrins, fibrinogen, and urokinase-type plasminogen acti- A more comprehensive assessment of protein production
vator, among others, can activate latent TGF-β during liver by stellate cells was generated by an unbiased proteomic anal-
injury, which eventually induces stellate cell activation (25, ysis of the cells and their surrounding ECM (13, 127, 236).
227, 278). On the other hand, it can be inactivated by binding In initial work by Kristensen, the patterns of protein expres-
to the proteoglycan decorin (15). The recent elucidation of sion were compared between quiescent, in vivo activated and
the latent TGF-β structure (271) could yield important new in vitro activated stellate cells, by two-dimensional-gel elec-
approaches to selectively blocking its activation in vivo. trophoresis. From the 300 identified proteins, 83 were found to
Vascular endothelial growth factor (VEGF) is also be secreted, including collagen α1 (I), α1 (III), and α2 (I); α1-
expressed by quiescent stellate cells (316). Its potent mito- antitrypsin; calcyclin, calgizzarin, and galectin-1; proteases
genic effect toward sinusoidal and endothelial cells under- including plasminogen activator inhibitor-1 and cathepsin A,
scores the key role that stellate cells play in communication B, and D; ganglioside GM2; among others. A more recent
and control of liver homeostasis. This important paracrine sig- analysis by Ji et al. has emphasized the importance of stellate
naling pathway between stellate cells and sinusoidal endothe- cells in also generating immunoregulatory molecules, con-
lium, mediated by VEGF and soluble guanylate cyclase, may sistent with their function in conferring immune tolerance in
be critical for sinusoidal homeostasis in normal liver and liver (130, 290, 318).
regeneration (309, 310, 319).
Other growth factors synthesized by stellate cells are Drug metabolism and detoxification
insulin-like growth factors (IGF-I and IGF-II), transforming HSCs express both alcohol- and acetaldehyde-
growth factor α (TGF-α), EGF, stem cell factor, and fibrob- dehydrogenases, but not cytochrome P450-2E1 (34)
last growth factors (both acidic and basic FGF), although their and it is likely that their contribution to ethanol detoxification
contributions may be more critical during liver development is minimal compared to hepatocytes. Apart from P450-2E1,
and regeneration (39, 40, 75, 177, 182, 191, 247, 298, 332). other isoforms of cytochrome p450 are expressed by stellate
cells, and are downregulated during cellular activation (321);
Neurotrophins and their receptors however, their roles in cellular quiescence and activation are
NGF, brain-derived neurotrophin, neurotrophin 3, neu- unknown. Some cytochrome p450 isoforms have been iden-
rotrophin 4/5, the low-affinity NGF receptor p75 and the high- tified in stellate cells (160), implicating their participation in
affinity tyrosine kinase receptors B and C are all expressed xenobiotic detoxification and oxidant stress response.
by HSCs (36), and/or their precursors (284). A number of
potential functions of these pathways are suggested by their
activities in other tissues; however, to date, their best known Role of Hepatic Stellate Cells in Liver
function in liver is in contributing to stellate cell activation Injury and Fibrosis
and tissue repair (36, 137, 215).
The framework for understanding stellate cell activation was
established several years ago (63), and remains a practical
Other mediators
and relevant template for characterizing the cell’s response to
Endothelin-1 (ET-1) is a potent vasoconstrictor produced pri- injury. A common consequence of liver injury is parenchy-
marily by endothelial cells in normal liver (322), but also mal damage with an increase in apoptotic bodies, Kupffer
by stellate cells. Interestingly, during liver injury, endothe- cell activation, production of oxidative species, and ECM
lial cells decrease their production and stellate cells become remodeling (65), which all function as triggers for cellu-
the dominant source of ET-1, which correlates with stel- lar activation. Activation of stellate cells comprises two
late cell activation (138, 221, 242, 270) highlighting the well-established phases: initiation (also called “preinflamma-
complex interplay of ET-1 between stellate and endothelial tory stage”) and perpetuation, which can be followed by a

Volume 3, October 2013 1477


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

potential third phase, resolution, if the liver injury resolves HGF, and then inhibited stellate cell expression of the type I
(58, 64, 69). During resolution of fibrosis, loss of activated collagen gene in a Met signal-dependent manner (147); and,
stellate cells occurs through numerous pathways and there is (iii) activation of human stellate cells in culture is suppressed
now evidence that stellate cells can not only undergo apopto- by human platelets or platelet-derived ATP via the adenosine-
sis, but are also able to either become senescent or revert to a cAMP signaling pathway (114).
quiescent phenotype (67, 143, 294). The elucidation of molec- Hepatocytes are the main target for most forms of liver
ular mechanisms underlying these events may accelerate the injury including viral infection, alcohol, and obesity, among
discovery of potential antifibrotic drug targets. others (198). Following injury, damaged hepatocytes become
a major source of lipid peroxides and apoptotic bodies
that initiate stellate activation through a process mediated
Initiation of stellate cell activation by Fas and TRAIL (32). The contribution of hepatocyte-
Stellate cell initiation promotes changes in gene expression derived apoptotic bodies to stellate cell activation is indepen-
and phenotype that render the cells susceptible to the changing dent of the inflammatory response, since in cultured stellate
environment and stimuli in the injured liver, thereby promot- cells, addition of hepatocyte apoptotic fragments are directly
ing the transition to the perpetuation phase. The earliest sig- fibrogenic (33), and can also activate Kupffer cells (31).
nals triggering the initiation of stellate cell activation result Hepatocytes also express P450-2E1, an important enzyme
from paracrine stimulation by neighboring cell populations involved in the metabolism of xenobiotics as ethanol, and a
(endothelial cells, platelets, immune cells, and hepatocytes) potent source of ROS (193) that can stimulate stellate cell
and changes in its surrounding ECM (Fig. 3). fibrogenesis (193).
As endothelial cells comprise the vascular lining of the Changes in the composition and stiffness of ECM also
liver’s sinusoids, they play a vital role in these early stages. impact on stellate cell responses (84, 121, 315) suggesting
They induce the activation of stellate cells by secreting a feed-forward loop where stellate cell mediated changes in
fibronectin (125) and by activating latent transforming TGF-β ECM further drive stellate cell activation. Early changes in
(149), as well as through secretion of a range of mediators that transcription factor activity in response to ECM, as well as
modulate inflammation and participate in cellular crosstalk soluble signals set the stage for a broad phenotypic transi-
(275, 319). Fibronectin’s effects are largely promigratory and tion of the cells, and a large number of nuclear factors have
not fibrogenic, suggesting that stellate cell migration is an been implicated [see (173) for review]. Additionally, path-
important first step in responding to injury (210). Platelets also ways of translational, transcriptional and posttranscriptional
contribute by secreting TGF-β, as well as EGF and platelet- regulatory control (including epigenetic pathways and miR-
derived growth factor (PDGF), the most potent stellate cell NAs) contribute to this process (18, 41, 99, 163, 172, 174,
mitogen identified (14, 24). Overall however, the resident 218, 243, 297).
hepatic macrophage population may be the main source of While the initial presumption that transcription factors
PDGF, as well as other paracrine mediators that drive stellate largely stimulate stellate cell activation, it appears equally
cell activation (106, 287, 308). true that other factors repress activation, and their activity is
There is an increasingly nuanced understanding of downregulated during cellular activation. Three key examples
how inflammatory and immune cells regulate stellate cell include Lhx2 (306), KLF6 (85), and Foxf1 (133), in which
responses and activation. In particular, T cells, dendritic cells each contribute to preservation of a quiescent phenotype, such
(DC), and macrophage subsets all have well defined inter- that their loss or downregulation derepresses the activation
actions with stellate cells [see (76, 131, 152, 179, 313) for program.
reviews]. Among these, recent studies have characterized a
specific macrophage subset in rodents, Ly-6Clo , that are vital Perpetuation of stellate cell
for regression of hepatic fibrosis (231). On the other hand, activation—mechanisms and implications
different macrophages can drive stellate cell function includ-
After the initial liver injury, stellate cells initiate activation
ing stimulation of matrix synthesis, cell proliferation, and
followed by a process of perpetuation, leading to accumula-
retinoid release by secreting TGF-β, TNF-α, and MMP-9, and
tion of ECM and culminating in the formation of scar tissue.
production of reactive oxygen species (ROS) and lipid perox-
Perpetuation of stellate cell activation is a tightly orches-
ides (230). Moreover, ROS produced by hepatic macrophages
trated process that includes a number of functional responses
can initiate downstream signals that include osteopontin, an
including proliferation, fibrogenesis, chemotaxis, contractil-
ECM protein that can induce collagen (300) and perpetuate
ity, matrix degradation, retinoid loss, and cytokine/chemokine
the activated stellate cell phenotype. Recent studies further
expression (Fig. 4).
implicate an inhibitory role of platelets in blocking stellate
cell activation based on the following: (i) transgenic throm-
Proliferation
bocytopenic mice develop exacerbated liver fibrosis, with
increased expression of type I collagen α1 and α2, during PDGF is most potent stellate cell mitogen during liver
cholestasis (147); (ii) in vitro experiments reveal that, upon injury. An increase in available PDGF and stellate cell
exposure to stellate cells, platelets became activated, released responsiveness due to increased expression of PDGF receptor

1478 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

(A) Normal liver

Portal triad
Bile duct Hepatocytes

HSC

Sinusoidal space of Disse


Terminal
Portal vein Sinusoidal endothelial cells KC hepatic
vein

Hepatic
arteriole

(B) Fibrotic liver

HSC activation Loss of hepatocyte


and proliferation microvilli

Loss of endothelial fenestrations Distortion


of veins

Increase in fibril-forming
collagen in space of Disse

Fibril-forming collagens (types I, III, and V)


Basement membrane collagens (type IV and VI)
Glycoconjugates (laminin, fibronectin, glycosaminoglycans, and tensacin)

Hernandez-Gea V, Friedman SL. 2011.


Annu. Rev. Pathol. Mech. Dis. 6:425–56
Figure 3 Matrix and cellular alteration in hepatic fibrosis. Normal liver parenchyma contains epithelial cells (hepatocytes) and
nonparenchymal cells: fenestrated sinusoidal endothelium, hepatic stellate cells (HSCs), and Kupffer cells (KCs). (A) Sinusoids are
separated from hepatocytes by a low-density basement membrane-like matrix confined to the space of Disse, which ensures metabolic
exchange. Upon injury, the stellate cells become activated and secrete large amounts of extracellular matrix (ECM), resulting in
progressive thickening of the septa. (B) Deposition of ECM in the space of Disse leads to the loss of both endothelial fenestrations and
hepatocyte microvilli, which results in both the impairment of normal bidirectional metabolic exchange between portal venous flow
and hepatocytes and the development of portal hypertension.

Volume 3, October 2013 1479


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

Initiation Perpetuation

Injury Proliferation
Contractility
Oxidative stress
Apoptotic bodies PDGF
ET–1
LPS VEGF NO
Paracrine stimuli FGF Fibrogenesis
TGF-β1/
CTGF

MMP-2&9; MT-1-MMP

TIMP-1,2 Altered matrix


Reversion degradation
PDGF
Chemokines

Resolution
Chemokines Adenosine
TLR ligands
HSC
TIMP-1,2 chemotaxis
TRAIL
Fas T cells
B cells
NK cells
Apoptosis NK-T cells
Inflammatory
signaling

Figure 4 Pathways of hepatic stellate cell activation and loss during liver injury and resolution.
Features of stellate cell activation can be distinguished between those that stimulate initiation and those
that contribute to perpetuation. Initiation is provoked by soluble stimuli that include oxidant stress signals
(reactive oxygen intermediates), apoptotic bodies, lipopolysaccharide (LPS), and paracrine stimuli from
neighboring cell types including hepatic macrophages (Kupffer cells), sinusoidal endothelium, and
hepatocytes. Perpetuation follows, characterized by a number of specific phenotypic changes including
proliferation, contractility, fibrogenesis, altered matrix degradation, chemotaxis, and inflammatory
signaling. During resolution of hepatic fibrosis, there is both programmed cell death (apoptosis) to
clear fibrogenic cells, as well as reversion to a more quiescient phenotype. FGF, fibroblast growth
factor; ET-1, endothelin-1; NK, natural killer; NO, nitric oxide; MT, membrane type. Reprinted, with
permission, from (66).

results in rapid proliferation and an overall increase in by stellate cells also drives stellate cell proliferation in an
the absolute number of stellate cells with a profibrogeneic AKT-dependent manner (61). This mechanism might con-
phenotype (24, 220, 317). Tumor necrosis factor (TNF) alpha tribute to the antiproliferative effects of activated Vitamin
signaling also contributes to PDGF-mediated stellate cell D, 1,25(OH)(2)D(3) (1). Vitamin D receptor is expressed by
proliferation primarily through the TNF receptor 1 (291). quiescent stellate cells (79), and its expression is downregu-
Stellate cells are also responsive to a wide array of factors lated with activation. Consequently, treatment of stellate cells
including, VEGF (327), thrombin, EGF, keratinocytre growth with 1,25(OH)(2)D(3) dampens proliferation via cyclin D1
factor (282), and bFGF (328). Among those, VEGF is one of suppression and decreases expression of type I collagen and
the major cytokines secreted by activated HSCs, which drives TIMP-1 while simultaneously increasing MMP-9 expression.
both angiogenesis and fibrogenesis, as described above (44, Patients with liver disease exhibit vitamin D deficiency, and
135, 159). VEGF production is dependent on the overex- appropriate supplementation might prove to be a useful ther-
pression of COX-2 protein via phospho-p42/44 MAP kinase apeutic intervention (225).
activation (329). Overall, HSCs contribute to both wound MicroRNAs are small 19-24 noncoding RNA sequences
healing and tumor growth, a conclusion underscored by sev- that can regulate posttransciptional gene expression by
eral studies implicating this cell type in the development and sequence-specific binding to the 3 -UTR of mRNAs to pro-
growth of both primary and metastatic tumors (47, 134, 209). mote their degradation. The role of miRNAs in fibrosis pro-
Tissue inhibitors of matrix metalloproteinases (TIMPs) gression is being clarified, as miRNA levels change in livers
are profibrogenic by inhibiting matrix degradation, and pro- of patients with fibrotic disease and in stellate cells during
moting stellate cell survival. Increased TIMP-1 expression activation (93, 99, 112, 196, 197, 205, 243, 307). In stellate

1480 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

cells, miRNAs control proliferation and fibrogenesis liver resident cells and inducing collagen type I production
by regulating protein expression of proproliferative and in HSCs through engagement of the signal transducer and
profibrogenic signaling pathways. In particular, miR-27a and activator of transcription 3 signaling pathway (181).
-27b and miR29b are upregulated in activated stellate cells, As discussed above, (see “Proliferation”) miRNAs play
whereas their suppression leads to decreased proliferation and a significant role in stellate cell biology and can modulate
an increase in lipid droplets indicative of the quiescent pheno- collagen synthesis. MiR-29b binds directly to the 3 -UTR of
type (205, 244). miR-27a -27b directly target the 3 -UTR of collagen-IαI and -IV, thereby inhibiting its translation. miR-
retinoid X receptor α (RXRα) to inhibit its expression. RXRα 29b expression is repressed by TGF-β, and its overexpression
can decrease DNA synthesis, leading to growth arrest in stel- inhibits TGF-β induced collagen expression via a SMAD-
late cells (100). Furthermore, RXRα regulates adipogenesis independent mechanism, while HGF, which exhibits antifi-
by activation of peroxisome proliferator-activated receptor γ brotic effects in stellate cells, induces miR-29b expression
(PPARγ) which is a master regulator of stellate cell activation (244, 266). MiR-21, whose expression is enhanced during
(297). RXRα expression is decreased in stellate cell activa- fibrotic disease, is also controlled by TGF-β signaling. TGF-
tion and its expression in activated stellate cells increases with β functions via 2 distinct mechanisms to increase miR-21
inactivation of miR-27a -27b, indicating a direct interaction production. Smad3 induces miR-21 transcription, while both
between RXRα and miR27a -27b (128). In contrast, miR- Smad2 and Smad3 enhance miR-21 maturation.
195 is downregulated during stellate cell proliferation and
its expression is induced upon treatment with IFN-β, which
Chemotaxis
exhibits antifibrotic effects independent of its antiviral activ-
ity. Treatment with IFN-β downregulates cyclin E1 and upreg- Stellate cell chemotaxis is an important event in the gener-
ulates p21 in a miR-195 specific manner thereby promoting ation of fibrotic septae by allowing activated cells to align
cell cycle arrest and decreased stellate cell proliferation (265). within regions of injury. Stellate cells migrate primarily
towards chemoattractant cytokines (chemokines), and they
express a range of chemokine receptors and their cognate
Fibrogenesis
chemokine ligands. Notably, stellate cells migrate towards
Production of ECM, in particular collagen type I, is a hallmark PDGF (113, 142), VEGF, Ang-1 (199), TGF-β1 , EGF (323),
of activated stellate cells. Production of collagen type I by b-FGF (59), CCL2 (176), and CXCR4 (254), and CXCR3
stellate cells is regulated both transcriptionally and posttran- specific ligands (23). The mechanism of chemotaxis includes
sciptionally (2, 37, 73, 117-119, 167, 175, 214, 238, 279,280, a cytoskeletal remodeling with cell spreading at the tip, move-
295). TGF-β1 is major driver of this process through autocrine ment of the cell body towards the stimulus, and retraction of
and paracrine stimulation of ECM production (see above). trailing protrusions (180). Oxidant signaling contributes to
The other well-characterized fibrogenic cytokine towards stel- these responses. Specifically, numerous chemoattractant sig-
late cells is connective tissue growth factor (CTGF/CCN2). nals (PDGF, VEGF, and CCL2) increase NADPH oxidase-
Levels of CTGF are increased in liver injury and the cytokine dependent intracellular ROS and activation of the ERK1/2
promotes a range of profibrotic activities toward stellate cells, and JNK1/2 pathways (50, 212). Furthermore, generation of
mediated by a G-coupled protein receptor (78, 92, 111). CTGF intracellular superoxide anion or H2 O2 by treatment with
represents a very appealing target for antifibrotic therapy, menadione promotes cell migration even in the absence of
since unlike antagonism of TGF-β1, CTGF inhibition should specific chemoattractants (198).
have no impact on hepatocyte growth or confer a risk of car- Hypoxia is another broad activator of stellate cell migra-
cinogenesis. tion, which functions via two distinct mechanisms. After
There is a growing list of other factors that contribute induction of hypoxic conditions, mitochondrial-generated
to fibrogenesis, including signaling molecules, chemokines, ROS activate the ERK1/2 and JNK1/2 pathways, driving
and cellular stressors (250). For example, osteopontin, an migration. Sustained hypoxia leads to a HIF-1α-dependent
ECM cytokine expressed by stellate cells, activates collagen increased production and secretion of VEGF by stellate cells,
I expression via integrin α(V)β(3) engagement and activa- promoting their mobility (200).
tion of the PI3K/pAkT/NFκB signaling pathways (300). Fur- Since stellate cell mobilization is also required for tis-
thermore, the recent identification of receptors to profibro- sue wound healing, it has been reported that the space of
genic chemokines on stellate cells including CXCR4 (108), Disse microenvironment, per se, is another key factor in reg-
CCR1, CCR5 (262), CXCR2 (281), and CCR2 (263), add ulating the migratory behavior of stellate cells. ECM com-
to the repertoire of signals promoting stellate cell activation. ponents including MMP-2 and type I collagen are able to
The targetability of chemokine receptors with small molecule induce stellate cell migration (208, 323). Cellular fibronectin
inhibitors makes them ideal candidates for antifibrotic ther- containing an alternatively spliced domain A (EIIA) is upreg-
apies. Recently, blockade of IL-17A has been proposed as ulated during liver injury. Signaling specifically by the EIIA
a potential strategy for cirrhosis treatment due to its induc- fibronectin variant though integrin receptor α(9)β(1) on stel-
tion (together with its receptor) in response to liver injury. late cells promotes formation of lamellipodia and cellular
IL-17A may promote fibrosis by activating inflammatory and motility, further implicating ECM signaling in stellate cell

Volume 3, October 2013 1481


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

biology (210). The hyaluronic acid receptor (CD44) is also Retinoid loss
increased in liver injury and repair, promoting both activation
Recently, there has been a renewed focus on the role
and migration of stellate cells (140). Interestingly, a specific
of retinoid loss in stellate cell activation and collagen
splice variant (CD44v6) is responsible for up to 50% of this
production. Stellate cell activation is characterized by the loss
migration, confirming the idea that activated stellate cells may
of perinuclear retinoid droplets (65, 70); however, their func-
depend, to some degree, on CD44v6 and hyaluronic acid for
tion in activation and fibrogenesis is only now being revealed.
migration.
Stellate cells are the largest reserve of retinoids in the body
(∼60%) and conversion of retinol into retinyl ester is a hall-
mark of stellate cell activation. The abundance of vitamin A
Contraction in stellate cells is heterogenous depending on the intralobular
position of the cell (80) and may be indicative of alternate acti-
Stellate cell contraction is thought to be a primary determi-
vation states. Quiescent stellate cells that are isolated based
nant of portal hypertension in patients with end-stage liver
on their collagen 1 expression display an increase in CYP251
disease. The factors leading to portal hypertension include
retinoid catabolizing cytochrome, a decrease in retinyl esters,
increased blood flow, increased intrahepatic resistance, and
and a more activated phenotype compared to cells isolated
disrupted liver architecture. During injury, the hepatic sinu-
based on their buoyancy in gradient centrifugation (45).
soids undergo both morphological and functional changes
LRAT, which catalyzes the esterification of retinol into
mediated by HSCs. Dramatic remodeling occurs, character-
retinyl ester, is the sole acyltransferase found in the liver. With
ized by deposition of collagen matrix, loss of fenestrations
stellate cell activation, LRAT expression is lost. Addition-
and increased density of contractile HSCs (293). Addition-
ally, treatment with IL-1 promotes decreased LRAT expres-
ally, there is an imbalance of vasoactive forces characterized
sion (139). Despite its apparent role in stellate cell activation,
by deficient nitric oxide production and an increase in vaso-
mice deficient in LRAT neither display spontaneous fibrogen-
constrictive substances including ET-1, angiotensinogen II,
esis, nor do they exhibit increased fibrogenesis in liver injury
eicosanoids, atrial natriuretic peptide, somatostatin, and car-
models, indicating that perhaps retinoid loss is a marker of
bon monoxide (237, 239, 240, 292). Together, these factors
activation, but is not crucial for stellate cell activation (144).
lead to an increase in sinusoidal resistance and portal hyper-
However, treatment with retinoic acid can decrease stellate
tenstion.
cell activation as reflected in reduced collagen I, MMP-9, and
The final step in the induction of cellular contraction
α-SMA by inhibiting expression of TGF-β (98).
is phosphorylation of MLC-2, resulting from a calcium-
In contrast to the lack of dependence on LRAT for fibro-
dependent and a calcium-sensitive pathway. In the calcium-
genesis by stellate cells, mice deficient in LRAT are protected
dependent pathway, typically in skeletal and cardiac muscle,
from chemical hepatocarcinogenesis. An increase in active
release of Ca2+ from the endoplasmic/sarcoplasmic reticulum
retinoids due to their lack of conversion to retinol storage
leads to activation of myosin light-chain kinase (MLCK) by
form leads to an overall antiproliferative effect, increased p21
calmodulin and subsequent phosphorylation of MLC. Alter-
levels, and inhibition of tumor progression (144, 272).
natively, in the Ca2+ -sensitive pathway, the dominant path-
An additional link between retinoid metabolism to stel-
way in stellate cell mediated contraction and smooth mus-
late cell activation has emerged through the recognition that
cle cells, Rho-kinase inactivates the myosin binding subunit
this process requires cellular autophagy (104). Specifically,
by phosphorylation, thereby preventing it from dephospho-
hydrolysis of retinyl esters liberates fatty acids that are metab-
rylating MLC and inhibiting contraction. The net effect of
olized by β-oxidation, generating the substrates that are essen-
Rho-kinase activation is increased phosphorylated-MLC and
tial for fueling the energy-intensive pathways of cellular acti-
contraction (168).
vation. The free retinol can be detected in the extracellular
Adenosine plays an important role in stellate cell dif-
milieu under these conditions (72), but pathways enabling
ferentiation, proliferation, and type I collagen production.
its cellular egress are not known. Similar to stellate cells,
Despite these profibrotic effects, however, adenosine inhibits
autophagy contributes to the intracellular catabolism of lipids
stellate cell contraction (as well as chemotaxis) via loss of
in hepatocytes, fibroblasts (274), and neurons (150, 151).
actin stress fibers. Engagement of the A2a receptor by adeno-
Moreover, inhibition of autophagy in hepatocytes leads to
sine promotes PKA activity and Rho A inhibition (276),
reduced rates of β-oxidation and marked lipid accumulation
which establishes a rationale for Rho antagonism as a strat-
in cytosolic lipid droplets (274).
egy for treatment of portal hypertension (132). Adenosine
therefore promotes both injurious and protective effects on
Matrix degradation
stellate cells and understanding these different functions will
be important in the development of antifibrotic agents. Inter- Fibrosis is a dynamic process of matrix production and degra-
estingly, the antifibrotic activities of caffeine, now validated dation. Fibrotic progression is characterized by the replace-
in several large cohorts (43, 187, 211), may reflect the com- ment of normal basement membrane, collagen type IV, with
pound’s effect in reducing adenosine signaling in stellate scar forming collagen type I. Early matrix degradation is an
cells (38). important step in fibrosis and may be essential for stellate cell

1482 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

migration to sites of injury. Stellate cells are the prominent their maturation within the liver. For example, stellate cells
source for MMP-2, MMP-9, and MMP-13 (7, 96, 183, 302), can inhibit the priming of naı̈ve T cells in a cell contact-
which function as either interstitial collagenases (MMP-13) dependent CD54 mechanism. Incubation of stellate cells
or gelatinases (MMP-2 and -9). with DCs and OT-1 T-cells inhibits T cell proliferation and
Despite the dependence of fibrosis progression on MMPs, reduces CD25 and CD44 activation markers (255). The abil-
these molecules paradoxically exhibit antifibrotic properties, ity of stellate cells to inhibit T cell proliferation is depen-
and their expression is suppressed with fibrotic progression. dent on their activation state, since quiescent stellate cells
MMP-2 inhibits collagen type I production by stellate cells do not possess this inhibitory function. Proteomics analy-
and mice lacking MMP-2 therefore exhibit increased fibro- sis reinforces this immune suppressive role (127). In the
sis (229). Furthermore, MMP-2 is able to regulate stellate transplantation setting, endotoxin-stimulated stellate cells are
cell apoptosis by cleavage of the extracellular domain of N- important immune regulators by inducing selective expan-
cadherin, further supporting its antifibrotic role (97). Some of sion of tolerance-promoting Tregs to reduce inflammation
the paradoxical activity of MMPs in vivo may be explained by and alloimmunity (46).
their simultaneous activation of macrophages (87) a feature Pattern recognition receptors, particularly TLR4, con-
often overlooked in interpreting their contribution to fibrosis. tribute significantly to stellate cell responses. Stimulation
Regulation of MMP expression occurs through several mech- of stellate cells with the TLR4 ligands enhances TGF-β
anisms. TIMPs, which are prominently expressed by stellate signaling and production of inflammatory and chemotactic
cells, bind MMPs, inactivating them. Additionally, with pro- cytokines (CCL2, IL-6), leading to a more profibrogenic
gressive fibrosis, MMP-9 and MMP-13 are repressed at the phenotype (94, 185, 213, 264). Additionally, stellate cell
chromatin level and access by the transcription factors NF-κB TLR3 activation confers antifibrotic activities by stimulating
and AP-1 is inhibited. Impaired histone acetylation is associ- interleukin-10 production (30).
ated with permanently silenced genes, and activated stellate During late stage fibrosis when portal hypertension is
cells have a global increase in histone deacetlyase-4 leading present there is an increased bacterial load delivery from the
to decreases in acetylation in the MMP-9 and -13 promoter gut to the liver due to increased intestinal permeability, leading
regions and gene repression (226). to an increase in LPS and TLR4 activation by stellate cells.
While the source of enzymes that degrade ECM in liver Treatment of mice with the intestinal decontaminant rifax-
has been elusive, findings increasingly point to subsets of imin, decreases portal pressure, fibrosis, and angiogenesis in
macrophages that have fibrolytic potential. Indeed, a recent a TLR4-specific manner (331). Specifically, rifaximin treat-
study has characterized a specific subtype of macrophages ment inhibits production of stellate cell derived fibronectin.
that express the surface marker Ly-6C as a recruited cellular LSECs are the key mediators during angiogenesis. Stellate
subset that secretes a range of proteases that have the capacity cell derived fibronectin is able to induce LSEC migration, and
to degrade ECM (230, 231). tubulogenesis, thereby contributing to the pathogenic effects
of LPS (331).
The stellate cell’s capabilities now extend to their role
Immunoregulation
in DC development (109). DCs exposed to stellate cells or
In recent years stellate cells have emerged as a promi- their supernatants express low CD11c, CD86, and major his-
nent determinant of hepatic immunoregulation during injury. tocompatibility complex class II, which elicits an inferior
They express of a battery of chemokines including CCL2 allostimulatory function compared with conventional DC.
(277), CCL5 (260), CXCL2, CCL21 (22), CXCL8, CXCL9, From the complementary point of view, the inflamma-
CXCL10, CXCL12, and CX3CL3 which have been shown to tory response can also modulate stellate cell activation and
recruit neutrophils, macrophages/monocytes, NK/NKT cells, fibrogenesis. Therefore, a positive feedback loop exists in
DCs, and T-cells, thereby establishing their role in immune which inflammatory and fibrogenic cells stimulate each other
cell infiltration. Many of these chemokines also function inde- in amplifying fibrosis. Thus, cell types regulating progression
pendent of immune cells and modulate cellular differentiation, and resolution of fibrosis include macrophages (a pivotal cell
survival, proliferation, and apoptosis. with the potential for both a pro- and antifibrotic capacity
Stellate cells can control immune cell function through by secreting/regulating TGF-β1, PDGF, MMPs, and TIMPs)
three interrelated mechanisms (283, 288): (i) cell surface (230, 231), natural killer cells (antifibrotic activity by inhibit-
expression of chemokines and/or delivery of chemokines to ing and/or killing activated stellate cells) (77, 189), T-cells
endothelial cells promotes lymphocyte adhesion and subse- (responsible for initiating/maintaining the adaptive immune
quent migration and activated stellate cells specifically pro- response and may induce liver injury upon CCL3/MIP1α
mote ICAM-1 and VCAM-1 dependent adhesion and migra- recruitment of CCR1 expressing CD4+ T cells) (286), B-cells
tion (107); (ii) increased expression of stellate cell derived (fibrosis promotion in an antibody- and T cell-independent
chemokines establishes a chemoattractant gradient between manner) (201), DCs (involved in both proinflammatory and
the peripheral blood and the liver, thus driving immune cell immunogenic responses) (42, 131), as well as endothelial cells
migration into the liver; and (iii) stellate cell interaction (antigen-presenting cells for CD4+ and CD8+ T populations
with immune cells has a direct role in promoting/inhibiting with antigen clearance activity) (146).

Volume 3, October 2013 1483


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

Stellate Cells and Resolution of likely to be a significant pathway in fibrosis regression that
involves approximately 50% of the stellate cell population,
Liver Fibrosis and may in part be regulated by changes in PPARγ activity.
Despite the historical conception of liver fibrosis as a passive
and irreversible process due to hepatocyte collapse (223,224), Conclusion
the idea of fibrosis regression was proposed in the 1970s (217,
224) and demonstrated in the 1990s (17, 95). At a certain HSCs are among the most fascinating and versatile of cell
stage of disease, fibrosis may become irreversible, likely due types in mammalian biology. While their contributions to hep-
to significant collagen cross-linking and development of an atic injury and fibrosis has been well established for at least
insoluble and relatively hypocellular matrix, which may coin- two decades, clarifying this role further continues to bene-
cide with the appearance of clinical cirrhosis manifestations fit from remarkable discoveries uncovering the signals that
(124). Since stellate cells are an important contributor to ECM control cell plasticity, survival, and intracellular signaling. A
remodeling in liver fibrosis, three possibilities may account growing range of genes, epigenetic changes, mediators, and
for regression of fibrosis, either their apoptosis, senescence, secretory proteins of stellate cells has been uncovered using
or reversion to their quiescent stage. complementary methods of gene profiling and proteomics,
Apoptosis of stellate cells during liver fibrosis recovery which has led to a more integrated understanding of the cell’s
contributes to a decrement in the number of activated stellate regulation. Concurrently, genetic models including new meth-
cells (120, 123, 251). This situation coincides with a decrease ods to delete genes specifically in stellate cells further clarify
in TIMP-1 expression (122), which protects the cells from their roles in normal and injured liver. In aggregate, these dis-
apoptosis and inhibits the MMPs functions (192), favoring coveries have enhanced the perceived importance of HSCs not
the partial degradation of ECM (55). Of interest, the NF-κB only to the fibrotic response in liver, but also to normal liver
transcription factor plays a role in the stellate cell’s protec- homeostasis that includes blood flow regulation, regeneration,
tion from apoptosis during liver fibrosis resolution, since its and stem cell responses.
inhibition can accelerate recovery from liver fibrosis (202). A key issue in studying HSCs is the choice of model
Moreover, apoptosis can be spontaneously initiated in acti- and species. Studies over the past 25 years have utilized both
vated stellate cells via CD95L (Fas ligand) Bcl-2 and p53 rodent and human cells, as well as several methods to immor-
(90). Finally, other cell populations can contribute to stellate talize cultured cells, including SV40 T antigen gene transfec-
cell apoptosis. Hepatocytes can secrete NGF, which is apop- tion, spontaneous immortalization in low serum, or ectopic
totic towards stellate cells (228) (203). Natural killer cells telomerase gene expression. Whereas human cells have the
can also provoke stellate cell apoptosis through TRAIL and advantage of direct human disease relevance, they lack the
NKG2D, a stellate cell membrane receptor that is specifically capacity for in vivo gene deletion using knockout technology,
expressed by activated stellate cells, rendering them suscepti- which is now widely used for mouse models and isolated
ble to NK action (228). Kupffer cells can also induce stellate stellate cells. This raises an important issue, as yet unan-
cell apoptosis by a caspase-9-mediated mechanism (60). swered, whether the genotypes and phenotypes of rodent and
Recent studies have now established that HSCs can undero human stellate cells in normal and diseased liver are suffi-
p53-mediated senescence in culture (258) and in vivo (155). ciently similar to allow investigators to rely on rodent studies
Initial studies in cultured HSCs suggested that as the cells to understand and predict human disease targets. This concern
reached their replicative limit, they adopt a more inflammatory has become more pertinent based on a recent study indicating
and less fibrogenic phenotype (258). Based on these observa- that rodent models of inflammation bear little resemblance
tions, elegant in vivo approaches confirmed the development to human disease (268). While one approach might be to
of senescence in HSCs in vivo (155, 169). Moreover, this “humanize” mice so that they contain human stellate cells,
response is p53-dependent, because animals in which p53 is methods to achieve this goal would be tedious and would
specifically deleted in stellate cells have more fibrosis after likely require use of immunodeficient animals. An alternative
toxic liver injury (169). Remarkably, this senescence program approach would be to perform comprehensive comparative
regulates the polarization of macrophages towards a pheno- analysis of mRNAs, microRNAs, proteins, and/or epigenet-
type that is protective against experimental carcinogenesis, ics between human and rodent cells from normal and injured
unearthing a noncell autonomous pathway of tumor suppres- liver to determine how faithful rodent cells are to their human
sion by p53 through its impact on the surrounding stromal counterparts. Based on studies to date, there is no evidence
biology (169). that fibrogenic responses by stellate cells between species are
Until recently, reversion from activated to quiescent HSC widely divergent, but the question has not yet been rigorously
was only demonstrated in cultured cells. However, elegant addressed.
studies have now used genetic methods to document reversion Despite progress in the field, several other key issues also
of activated cells to a quiescent phenotype in vivo (67, 143, remain unresolved. The contribution of stellate cells to regen-
294). Reverted cells, however, have a heightened capacity to eration has long been assumed, but recent reports suggest
reactivate, compared to those that have never activated. Quan- that HSC depletion has no effect on regeneration (54, 144).
titatively, reversion of activated stellate cells to quiescence is Furthermore, use of transgenic animals has begun to trace

1484 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

stellate cell fate during fibrosis regression (143, 230, 294) but 14. Bachem MG, Melchior R, Gressner AM. The role of thrombocytes in
underlying mechanisms are still incomplete. Similarly, it is liver fibrogenesis: Effects of platelet lysate and thrombocyte-derived
growth factors on the mitogenic activity and glycosaminoglycan syn-
increasingly clear that stellate cells and the fibrotic stroma thesis of cultured rat liver fat storing cells. J Clin Chem Clin Biochem
they generate in injured liver accelerate the risk of hepato- 27: 555-565, 1989.
15. Baghy K, Iozzo RV, Kovalszky I. Decorin-TGFbeta axis in hepatic
cellular carcinoma, but methods to clarify pathways linking fibrosis and cirrhosis. J Histochem Cytochem 60: 262-268, 2012.
fibrosis to cancer are only now emerging. Also unclear is the 16. Ballardini G, Groff P, Badiali dGL, Schuppan D, Bianchi FB. Ito cell
heterogeneity: Desmin-negative Ito cells in normal rat liver. Hepatology
full range of paracrine interactions in stellate cell crosstalk 19: 440-446, 1994.
with sinusoidal endothelial cells, biliary epithelial cells, and 17. Bataller R, Brenner DA. Liver fibrosis. J Clin Invest 115: 209-218,
2005.
immune cell subsets. Finally, although activation of HSCs 18. Bian EB, Huang C, Ma TT, Tao H, Zhang H, Cheng C, Lv XW, Li
into contractile myofibroblasts is well recognized as the cen- J. DNMT1-mediated PTEN hypermethylation confers hepatic stellate
cell activation and liver fibrogenesis in rats. Toxicol Appl Pharmacol
tral event in hepatic fibrosis, this insight has not yet translated 264: 13-22, 2012.
into an effective antifibrotic therapy for patients with chronic 19. Bioulac-Sage P, Lafon ME, Saric J, Balabaud C. Nerves and perisinu-
soidal cells in human liver. J Hepatol 10: 105-112, 1990.
fibrosing liver diseases. Future studies in animal models will 20. Bissell DM, Arenson DM, Maher JJ, Roll FJ. Support of cultured hep-
further clarify these vital roles, while their translation into atocytes by a laminin-rich gel. Evidence for a functionally significant
subendothelial matrix in normal rat liver. J Clin Invest 79: 801-812,
effective antifibrotic drugs for fibrotic liver disease is sure to 1987.
emerge in the coming years. 21. Bissell DM, Wang SS, Jarnagin WR, Roll FJ. Cell-specific expression
of transforming growth factor-beta in rat liver. Evidence for autocrine
regulation of hepatocyte proliferation. J Clin Invest 96: 447-455,
1995.
Acknowledgements 22. Bonacchi A, Petrai I, Defranco RM, Lazzeri E, Annunziato F, Efsen
E, Cosmi L, Romagnani P, Milani S, Failli P, Batignani G, Liotta
F, Laffi G, Pinzani M, Gentilini P, Marra F. The chemokine CCL21
Work in the authors’ laboratory was supported by NIH Grants modulates lymphocyte recruitment and fibrosis in chronic hepatitis C.
DK56621, AA020709 (to S.L.F.), NIDDK F30-DK090986, Gastroenterology 125: 1060-1076, 2003.
23. Bonacchi A, Romagnani P, Romanelli RG, Efsen E, Annunziato F,
and T32-GM007280 (to Y.S.), and funds from the Alfonso Lasagni L, Francalanci M, Serio M, Laffi G, Pinzani M, Gentilini P,
Martin Escudero Fundation (to J.E.P.). Marra F. Signal transduction by the chemokine receptor CXCR3: Acti-
vation of Ras/ERK, Src, and phosphatidylinositol 3-kinase/Akt controls
cell migration and proliferation in human vascular pericytes. J Biol
Chem 276: 9945-9954, 2001.
References 24. Borkham-Kamphorst E, van Roeyen CR, Ostendorf T, Floege J, Gress-
ner AM, Weiskirchen R. Pro-fibrogenic potential of PDGF-D in liver
1. Abramovitch S, Dahan-Bachar L, Sharvit E, Weisman Y, Ben Tov A, fibrosis. J Hepatol 46: 1064-1074, 2007.
Brazowski E, Reif S. Vitamin D inhibits proliferation and profibrotic 25. Botella LM, Sanchez-Elsner T, Sanz-Rodriguez F, Kojima S, Shi-
marker expression in hepatic stellate cells and decreases thioacetamide- mada J, Guerrero-Esteo M, Cooreman MP, Ratziu V, Langa C, Vary
induced liver fibrosis in rats. Gut 60: 1728-1737, 2011. CP, Ramirez JR, Friedman S, Bernabeu C. Transcriptional activation
2. Anania FA, Womack L, Jiang M, Saxena NK. Aldehydes potentiate of endoglin and transforming growth factor-beta signaling compo-
alpha(2)(I) collagen gene activity by JNK in hepatic stellate cells. Free nents by cooperative interaction between Sp1 and KLF6: Their poten-
Radic Biol Med 30: 846-857, 2001. tial role in the response to vascular injury. Blood 100: 4001-4010,
3. Anthony PP, Ishak KG, Nayak NC, Poulsen HE, Scheuer PJ, Sobin 2002.
LH. The morphology of cirrhosis. Recommendations on definition, 26. Bourd-Boittin K, Bonnier D, Leyme A, Mari B, Tuffery P, Samson M,
nomenclature, and classification by a working group sponsored by the Ezan F, Baffet G, Theret N. Protease profiling of liver fibrosis reveals
World Health Organization. J Clin Pathol 31: 395-414, 1978. the ADAM metallopeptidase with thrombospondin type 1 motif, 1 as
4. Apte MV, Pirola RC, Wilson JS. Pancreatic stellate cells: A starring a central activator of transforming growth factor beta. Hepatology 54:
role in normal and diseased pancreas. Front Physiol 3: 344, 2012. 2173-2184, 2011.
5. Apte MV, Wilson JS. Dangerous liaisons: Pancreatic stellate cells and 27. Brasaemle DL, Wolins NE. Packaging of fat: An evolving model of
pancreatic cancer cells. J Gastroenterol Hepatol 27(Suppl 2): 69-74, lipid droplet assembly and expansion. J Biol Chem 287: 2273-2279,
2012. 2012.
6. Arthur MJ, Friedman SL, Roll FJ, Bissell DM. Lipocytes from normal 28. Buers I, Hofnagel O, Ruebel A, Severs NJ, Robenek H. Lipid
rat liver release a neutral metalloproteinase that degrades basement droplet associated proteins: An emerging role in atherogenesis. His-
membrane (type IV) collagen. J Clin Invest 84: 1076-1085, 1989. tol Histopathol 26: 631-642, 2011.
7. Arthur MJ, Stanley A, Iredale JP, Rafferty JA, Hembry RM, Friedman 29. Burt AD, Griffiths MR, Schuppan D, Voss B, MacSween RN. Ultra-
SL. Secretion of 72 kDa type IV collagenase/gelatinase by cultured structural localization of extracellular matrix proteins in liver biopsies
human lipocytes. Analysis of gene expression, protein synthesis and using ultracryomicrotomy and immuno-gold labelling. Histopathology
proteinase activity. Biochem J 287: 701-707, 1992. 16: 53-58, 1990.
8. Asahina K. Hepatic stellate cell progenitor cells. J Gastroenterol Hep- 30. Byun JS, Suh YG, Yi HS, Lee YS, Jeong WI. Activation of toll-like
atol 27(Suppl 2): 80-84, 2012. receptor 3 attenuates alcoholic liver injury by stimulating Kupffer cells
9. Asahina K, Sato H, Yamasaki C, Kataoka M, Shiokawa M, Katayama and stellate cells to produce interleukin-10 in mice. J Hepatol 58: 342-
S, Tateno C, Yoshizato K. Pleiotrophin/heparin-binding growth- 349, 2013.
associated molecule as a mitogen of rat hepatocytes and its role in 31. Canbay A, Feldstein AE, Higuchi H, Werneburg N, Grambihler A,
regeneration and development of liver. Am J Pathol 160: 2191-2205, Bronk SF, Gores GJ. Kupffer cell engulfment of apoptotic bodies stim-
2002. ulates death ligand and cytokine expression. Hepatology 38: 1188-1198,
10. Asahina K, Tsai SY, Li P, Ishii M, Maxson RE, Jr., Sucov HM, 2003.
Tsukamoto H. Mesenchymal origin of hepatic stellate cells, subme- 32. Canbay A, Higuchi H, Bronk SF, Taniai M, Sebo TJ, Gores GJ. Fas
sothelial cells, and perivascular mesenchymal cells during mouse liver enhances fibrogenesis in the bile duct ligated mouse: A link between
development. Hepatology 49: 998-1011, 2009. apoptosis and fibrosis. Gastroenterology 123: 1323-1330, 2002.
11. Asahina K, Zhou B, Pu WT, Tsukamoto H. Septum transversum- 33. Canbay A, Taimr P, Torok N, Higuchi H, Friedman S, Gores GJ. Apop-
derived mesothelium gives rise to hepatic stellate cells and perivascular totic body engulfment by a human stellate cell line is profibrogenic.
mesenchymal cells in developing mouse liver. Hepatology 53: 983-995, Lab Invest 83: 655-663, 2003.
2011. 34. Casini A, Pellegrini G, Ceni E, Salzano R, Parola M, Robino G, Milani
12. Baba S, Fujii H, Hirose T, Yasuchika K, Azuma H, Hoppo T, Naito S, Dianzani MU, Surrenti C. Human hepatic stellate cells express
M, Machimoto T, Ikai I. Commitment of bone marrow cells to hepatic class I alcohol dehydrogenase and aldehyde dehydrogenase but not
stellate cells in mouse. J Hepatol 40: 255-260, 2004. cytochrome P4502E1. J Hepatol 28: 40-45, 1998.
13. Bach Kristensen D, Kawada N, Imamura K, Miyamoto Y, Tateno C, 35. Cassiman D, Barlow A, Vander Borght S, Libbrecht L, Pachnis V.
Seki S, Kuroki T, Yoshizato K. Proteome analysis of rat hepatic stellate Hepatic stellate cells do not derive from the neural crest. J Hepatol 44:
cells. Hepatology 32: 268-277, 2000. 1098-1104, 2006.

Volume 3, October 2013 1485


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

36. Cassiman D, Denef C, Desmet VJ, Roskams T. Human and rat hepatic 58. Fallowfield JA, Iredale JP. Reversal of liver fibrosis and cirrhosis–an
stellate cells express neurotrophins and neurotrophin receptors. Hepa- emerging reality. Scott Med J 49: 3-6, 2004.
tology 33: 148-158, 2001. 59. Fibbi G, Pucci M, D’Alessio S, Grappone C, Pellegrini G, Salzano
37. Challa AA, Vukmirovic M, Blackmon J, Stefanovic B. Withaferin-A R, Casini A, Milani S, Del Rosso M. Transforming growth factor
reduces type I collagen expression in vitro and inhibits development of beta-1 stimulates invasivity of hepatic stellate cells by engagement
myocardial fibrosis in vivo. PLoS One 7: e42989, 2012. of the cell-associated fibrinolytic system. Growth Factors 19: 87-100,
38. Chan ES, Montesinos MC, Fernandez P, Desai A, Delano DL, Yee H, 2001.
Reiss AB, Pillinger MH, Chen JF, Schwarzschild MA, Friedman SL, 60. Fischer R, Cariers A, Reinehr R, Haussinger D. Caspase 9-dependent
Cronstein BN. Adenosine A(2A) receptors play a role in the pathogen- killing of hepatic stellate cells by activated Kupffer cells. Gastroen-
esis of hepatic cirrhosis. Br J Pharmacol 148: 1144-1155, 2006. terology 123: 845-861, 2002.
39. Chan KM, Fu YH, Wu TJ, Hsu PY, Lee WC. Hepatic stellate cells 61. Fowell AJ, Collins JE, Duncombe DR, Pickering JA, Rosenberg WM,
promote the differentiation of embryonic stem cell-derived definitive Benyon RC. Silencing tissue inhibitors of metalloproteinases (TIMPs)
endodermal cells into hepatic progenitor cells. Hepatol Res 43: 648- with short interfering RNA reveals a role for TIMP-1 in hepatic stellate
657, 2012. cell proliferation. Biochem Biophys Res Commun 407: 277-282, 2011.
40. Chen L, Zhang W, Zhou QD, Yang HQ, Liang HF, Zhang BX, Long 62. Friedman SL. Seminars in medicine of the Beth Israel Hospital, Boston.
X, Chen XP. HSCs play a distinct role in different phases of oval cell- The cellular basis of hepatic fibrosis. Mechanisms and treatment strate-
mediated liver regeneration. Cell Biochem Funct 30: 588-596, 2012. gies. N Engl J Med 328: 1828-1835, 1993.
41. Chen SL, Zheng MH, Yang T, Song M, Chen YP. Disparate profiles of 63. Friedman SL. Molecular regulation of hepatic fibrosis, an integrated
dys-regulated miRNAs in activated hepatic stellate cells. Hepatology cellular response to tissue injury. J Biol Chem 275: 2247-2250, 2000.
57: 1285-1286, 2013. 64. Friedman SL. Hepatic fibrosis-Overview. Toxicology 254: 120-129,
42. Connolly MK, Bedrosian AS, Mallen-St Clair J, Mitchell AP, Ibrahim J, 2008.
Stroud A, Pachter HL, Bar-Sagi D, Frey AB, Miller G. In liver fibrosis, 65. Friedman SL. Hepatic stellate cells – protean, multifunctional, and
dendritic cells govern hepatic inflammation in mice via TNF-alpha. enigmatic cells of the liver. Physiol Rev 88: 125-172, 2008.
J Clin Invest 119: 3213-3225, 2009. 66. Friedman SL. Mechanisms of hepatic fibrogenesis. Gastroenterology
43. Costentin CE, Roudot-Thoraval F, Zafrani ES, Medkour F, Pawlotsky 134: 1655-1669, 2008.
JM, Mallat A, Hezode C. Association of caffeine intake and histological 67. Friedman SL. Fibrogenic cell reversion underlies fibrosis regression in
features of chronic hepatitis C. J Hepatol 54: 1123-1129, 2011. liver. Proc Natl Acad Sci U S A 109: 9230-9231, 2012.
44. Coulouarn C, Corlu A, Glaise D, Guenon I, Thorgeirsson SS, Clement 68. Friedman SL, Arthur MJ. Reversing hepatic fibrosis. Sci Med 8: 194-
B. Hepatocyte-stellate cell cross-talk in the liver engenders a permissive 205, 2002.
inflammatory microenvironment that drives progression in hepatocel- 69. Friedman SL, Bansal MB. Reversal of hepatic fibrosis – fact or fantasy?
lular carcinoma. Cancer Res 72: 2533-2542, 2012. Hepatology 43: S82-S88, 2006.
45. D’Ambrosio DN, Walewski JL, Clugston RD, Berk PD, Rippe RA, 70. Friedman SL, Roll FJ, Boyles J, Arenson DM, Bissell DM. Mainte-
Blaner WS. Distinct populations of hepatic stellate cells in the mouse nance of differentiated phenotype of cultured rat hepatic lipocytes by
liver have different capacities for retinoid and lipid storage. PLoS One basement membrane matrix. J Biol Chem 264: 10756-10762, 1989.
6: e24993, 2011. 71. Friedman SL, Roll FJ, Boyles J, Bissell DM. Hepatic lipocytes: The
46. Dangi A, Sumpter TL, Kimura S, Stolz DB, Murase N, Raimondi G, principal collagen-producing cells of normal rat liver. Proc Natl Acad
Vodovotz Y, Huang C, Thomson AW, Gandhi CR. Selective expan- Sci U S A 82: 8681-8685, 1985.
sion of allogeneic regulatory T cells by hepatic stellate cells: Role 72. Friedman SL, Wei S, Blaner WS. Retinol release by activated rat hepatic
of endotoxin and implications for allograft tolerance. J Immunol 188: lipocytes: Regulation by Kupffer cell-conditioned medium and PDGF.
3667-3677, 2012. Am J Physiol 264: G947-G952, 1993.
47. Dapito DH, Mencin A, Gwak GY, Pradere JP, Jang MK, Mederacke 73. Fritz D, Stefanovic B. RNA-binding protein RBMS3 is expressed in
I, Caviglia JM, Khiabanian H, Adeyemi A, Bataller R, Lefkowitch JH, activated hepatic stellate cells and liver fibrosis and increases expression
Bower M, Friedman R, Sartor RB, Rabadan R, Schwabe RF. Promotion of transcription factor Prx1. J Mol Biol 371: 585-595, 2007.
of Hepatocellular Carcinoma by the Intestinal Microbiota and TLR4. 74. Frizell E, Liu SL, Abraham A, Ozaki I, Eghbali M, Sage EH, Zern
Cancer Cell 21: 504-516, 2012. MA. Expression of SPARC in normal and fibrotic livers. Hepatology
48. De Bleser PJ, Niki T, Rogiers V, Geerts A. Transforming growth factor- 21: 847-854, 1995.
beta gene expression in normal and fibrotic rat liver. J Hepatol 26: 75. Fujio K, Evarts RP, Hu Z, Marsden ER, Thorgeirsson SS. Expression
886-893, 1997. of stem cell factor and its receptor, c kit, during liver regeneration from
49. de Leeuw AM, McCarthy SP, Geerts A, Knook DL. Purified rat liver putative stem cells in adult rat. Lab Invest 70: 511-516, 1994.
fat-storing cells in culture divide and contain collagen. Hepatology 4: 76. Gao B, Radaeva S. Natural killer and natural killer T cells in liver
392-403, 1984. fibrosis. Biochim Biophys Acta 1832: 1061-1069, 2012.
50. De Minicis S, Brenner DA. NOX in liver fibrosis. Arch Biochem Biophys 77. Gao B, Radaeva S, Park O. Liver natural killer and natural killer T cells:
462: 266-272, 2007. Immunobiology and emerging roles in liver diseases. J Leukoc Biol 86:
51. Dooley S, Delvoux B, Lahme B, Mangasser-Stephan K, Gressner AM. 513-528, 2009.
Modulation of transforming growth factor beta response and signaling 78. Gao R, Brigstock DR. Connective tissue growth factor (CCN2) induces
during transdifferentiation of rat hepatic stellate cells to myofibroblasts. adhesion of rat activated hepatic stellate cells by binding of its C-
Hepatology 31: 1094-1106, 2000. terminal domain to integrin alpha(v)beta(3) and heparan sulfate proteo-
52. Dranoff JA, Ogawa M, Kruglov EA, Gaca MD, Sevigny J, Robson SC, glycan. J Biol Chem 279: 8848-8855, 2004.
Wells RG. Expression of P2Y nucleotide receptors and ectonucleoti- 79. Gascon-Barre M, Demers C, Mirshahi A, Neron S, Zalzal S, Nanci
dases in quiescent and activated rat hepatic stellate cells. Am J Physiol A. The normal liver harbors the vitamin D nuclear receptor in non-
Gastrointest Liver Physiol 287: G417-G424, 2004. parenchymal and biliary epithelial cells. Hepatology 37: 1034-1042,
53. Dranoff JA, Wells RG. Portal fibroblasts: Underappreciated mediators 2003.
of biliary fibrosis. Hepatology 51: 1438-1444, 2010. 80. Geerts A. History, heterogeneity, developmental biology, and functions
54. Ebrahimkhani MR, Oakley F, Murphy LB, Mann J, Moles A, Perugor- of quiescent hepatic stellate cells. Semin Liver Dis 21: 311-335, 2001.
ria MJ, Ellis E, Lakey AF, Burt AD, Douglass A, Wright MC, White 81. Geerts A, Bouwens L, Wisse E. Ultrastructure and function of hepatic
SA, Jaffre F, Maroteaux L, Mann DA. Stimulating healthy tissue regen- fat-storing and pit cells. J Electr Micr Tech 14: 247-256, 1990.
eration by targeting the 5-HT(2B) receptor in chronic liver disease. Nat 82. Geerts A, Lazou JM, De Bleser P, Wisse E. Tissue distribution,
Med 17: 1668-1673, 2011. quantitation and proliferation kinetics of fat-storing cells in carbon
55. Elsharkawy AM, Oakley F, Mann DA. The role and regulation of tetrachloride-injured rat liver. Hepatology 13: 1193-1202, 1991.
hepatic stellate cell apoptosis in reversal of liver fibrosis. Apoptosis 10: 83. Geerts A, Vrijsen R, Rauterberg J, Burt A, Schellinck P, Wisse E. In
927-939, 2005. vitro differentiation of fat-storing cells parallels marked increase of
56. Enzan H, Hayashi Y, Miyazaki E, Naruse K, Tao R, Kuroda N, collagen synthesis and secretion. J Hepatol 9: 59-68, 1989.
Nakayama H, Kiyoku H, Hiroi M, Saibara T. Morphological aspects 84. Georges PC, Hui JJ, Gombos Z, McCormick ME, Wang AY, Uemura
of hepatic fibrosis and Ito cells (hepatic stellate cells), with special M, Mick R, Janmey PA, Furth EE, Wells RG. Increased stiffness of
reference to their myofibroblastic transformation. In: Tanikawa K and the rat liver precedes matrix deposition: Implications for fibrosis. Am J
Ueno T, editors. Liver Diseases and Hepatic Sinusoidal Cells. Tokyo: Physiol Gastrointest Liver Physiol 293: G1147-G1154, 2007.
Springer-Verlag, 1999, p. 219-231. 85. Ghiassi-Nejad Z, Hernandez-Gea V, Woodrell C, Lang UE, Dumic
57. Erkan M, Adler G, Apte MV, Bachem MG, Buchholz M, Detlefsen K, Kwong A, Friedman SL. Reduced hepatic stellate cell expression
S, Esposito I, Friess H, Gress TM, Habisch HJ, Hwang RF, Jaster R, of Kruppel-like factor 6 tumor suppressor isoforms amplifies fibrosis
Kleeff J, Kloppel G, Kordes C, Logsdon CD, Masamune A, Michalski during acute and chronic rodent liver injury. Hepatology 57: 786-796,
CW, Oh J, Phillips PA, Pinzani M, Reiser-Erkan C, Tsukamoto H, 2013.
Wilson J. StellaTUM: Current consensus and discussion on pancreatic 86. Giampieri MP, Jezequel AM, Orlandi F. The lipocytes in normal human
stellate cell research. Gut 61: 172-178, 2012. liver. Digestion 22: 165-169, 1981.

1486 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

87. Gong Y, Hart E, Shchurin A, Hoover-Plow J. Inflammatory macrophage 113. Ikeda K, Wakahara T, Wang YQ, Kadoya H, Kawada N, Kaneda K. In
migration requires MMP-9 activation by plasminogen in mice. J Clin vitro migratory potential of rat quiescent hepatic stellate cells and its
Invest 118: 3012-3024, 2008. augmentation by cell activation. Hepatology 29: 1760-1767, 1999.
88. Gorbig MN, Gines P, Bataller R, Nicolas JM, Garcia-Ramallo E, 114. Ikeda N, Murata S, Maruyama T, Tamura T, Nozaki R, Kawasaki T,
Cejudo P, Sancho-Bru P, Jimenez W, Arroyo V, Rodes J. Human Fukunaga K, Oda T, Sasaki R, Homma M, Ohkohchi N. Platelet-derived
hepatic stellate cells secrete adrenomedullin: Potential autocrine factor adenosine 5 -triphosphate suppresses activation of human hepatic stel-
in the regulation of cell contractility. J Hepatol 34: 222-229, 2001. late cell: In vitro study. Hepatol Res 42: 91-102, 2012.
89. Greenbaum LE, Wells RG. The role of stem cells in liver repair and 115. Imai K, Senoo H. Morphology of sites of adhesion between hepatic
fibrosis. Int J Biochem Cell Biol 43: 222-229, 2011. stellate cells (vitamin A-storing cells) and a three-dimensional extra-
90. Gressner AM. The cell biology of liver fibrogenesis - an imbalance of cellular matrix. Anat Rec 250: 430-437, 1998.
proliferation, growth arrest and apoptosis of myofibroblasts. Cell Tissue 116. Inagaki Y, Kushida M, Higashi K, Itoh J, Higashiyama R, Hong YY,
Res 292: 447-452, 1998. Kawada N, Namikawa K, Kiyama H, Bou-Gharios G, Watanabe T,
91. Gressner AM, Weiskirchen R, Breitkopf K, Dooley S. Roles of tgf-Beta Okazaki I, Ikeda K. Cell type-specific intervention of transforming
in hepatic fibrosis. Front Biosci 7: D793-D807, 2002. growth factor beta/Smad signaling suppresses collagen gene expres-
92. Gressner OA, Gressner AM. Connective tissue growth factor: A fibro- sion and hepatic fibrosis in mice. Gastroenterology 129: 259-268,
genic master switch in fibrotic liver diseases. Liver Int 28: 1065-1079, 2005.
2008. 117. Inagaki Y, Mamura M, Kanamaru Y, Greenwel P, Nemoto T, Takehara
93. Guo CJ, Pan Q, Li DG, Sun H, Liu BW. miR-15b and miR-16 are K, Ten Dijke P, Nakao A. Constitutive phosphorylation and nuclear
implicated in activation of the rat hepatic stellate cell: An essential role localization of Smad3 are correlated with increased collagen gene tran-
for apoptosis. J Hepatol 50: 766-778, 2009. scription in activated hepatic stellate cells. J Cell Physiol 187: 117-123,
94. Guo J, Loke J, Zheng F, Hong F, Yea S, Fukata M, Tarocchi M, Abar 2001.
OT, Huang H, Sninsky JJ, Friedman SL. Functional linkage of cirrhosis- 118. Inagaki Y, Nemoto T, Kushida M, Sheng Y, Higashi K, Ikeda K,
predictive single nucleotide polymorphisms of Toll-like receptor 4 to Kawada N, Shirasaki F, Takehara K, Sugiyama K, Fujii M, Yamauchi
hepatic stellate cell responses. Hepatology 49: 960-968, 2009. H, Nakao A, De Crombrugghe B, Watanabe T, Okazaki I. Interferon
95. Hammel P, Couvelard A, O’Toole D, Ratouis A, Sauvanet A, Flejou alfa down-regulates collagen gene transcription and suppresses experi-
JF, Degott C, Belghiti J, Bernades P, Valla D, Ruszniewski P, Levy mental hepatic fibrosis in mice. Hepatology 38: 890-899, 2003.
P. Regression of liver fibrosis after biliary drainage in patients with 119. Inagaki Y, Truter S, Greenwel P, Rojkind M, Unoura M, Kobayashi K,
chronic pancreatitis and stenosis of the common bile duct. N Engl J Ramirez F. Regulation of the alpha 2(I) collagen gene transcription in
Med 344: 418-423, 2001. fat-storing cells derived from a cirrhotic liver. Hepatology 22: 573-579,
96. Han YP, Yan C, Zhou L, Qin L, Tsukamoto H. A matrix 1995.
metalloproteinase-9 activation cascade by hepatic stellate cells in trans- 120. Iredale JP. Hepatic stellate cell behavior during resolution of liver
differentiation in the three-dimensional extracellular matrix. J Biol injury. Semin Liver Dis 21: 427-436, 2001.
Chem 282: 12928-12939, 2007. 121. Iredale JP. Models of liver fibrosis: Exploring the dynamic nature of
97. Hartland SN, Murphy F, Aucott RL, Abergel A, Zhou X, Waung J, inflammation and repair in a solid organ. J Clin Invest 117: 539-548,
Patel N, Bradshaw C, Collins J, Mann D, Benyon RC, Iredale JP. Active 2007.
matrix metalloproteinase-2 promotes apoptosis of hepatic stellate cells 122. Iredale JP, Benyon RC, Pickering J, McCullen M, Northrop M, Pawley
via the cleavage of cellular N-cadherin. Liver Int 29: 966-978, 2009. S, Hovell C, Arthur MJ. Mechanisms of spontaneous resolution of
98. He H, Mennone A, Boyer JL, Cai SY. Combination of retinoic acid and rat liver fibrosis. Hepatic stellate cell apoptosis and reduced hepatic
ursodeoxycholic acid attenuates liver injury in bile duct-ligated rats and expression of metalloproteinase inhibitors. J Clin Invest 102: 538-549,
human hepatic cells. Hepatology 53: 548-557, 2011. 1998.
99. He Y, Huang C, Zhang SP, Sun X, Long XR, Li J. The potential of 123. Issa R, Williams E, Trim N, Kendall T, Arthur MJ, Reichen J, Benyon
microRNAs in liver fibrosis. Cell Signal 24: 2268-2272, 2012. RC, Iredale JP. Apoptosis of hepatic stellate cells: Involvement in res-
100. Hellemans K, Verbuyst P, Quartier E, Schuit F, Rombouts K, Chan- olution of biliary fibrosis and regulation by soluble growth factors. Gut
draratna RA, Schuppan D, Geerts A. Differential modulation of rat 48: 548-557, 2001.
hepatic stellate phenotype by natural and synthetic retinoids. Hepatol- 124. Issa R, Zhou X, Constandinou CM, Fallowfield J, Millward-Sadler H,
ogy 39: 97-108, 2004. Gaca MD, Sands E, Suliman I, Trim N, Knorr A, Arthur MJ, Benyon
101. Hendriks HF, Verhoofstad WA, Brouwer A, de Leeuw AM, Knook DL. RC, Iredale JP. Spontaneous recovery from micronodular cirrhosis: Evi-
Perisinusoidal fat-storing cells are the main vitamin A storage sites in dence for incomplete resolution associated with matrix cross-linking.
rat liver. Exp Cell Res 160: 138-149, 1985. Gastroenterology 126: 1795-1808, 2004.
102. Hernandez-Gea V, Friedman SL. Pathogenesis of liver fibrosis. Annu 125. Jarnagin WR, Rockey DC, Koteliansky VE, Wang SS, Bissell DM.
Rev Pathol 6: 425-456, 2011. Expression of variant fibronectins in wound healing: Cellular source
103. Hernandez-Gea V, Friedman SL. Autophagy fuels tissue fibrogenesis. and biological activity of the EIIIA segment in rat hepatic fibrogenesis.
Autophagy 8: 849-850, 2012. J Cell Biol 127: 2037-2048, 1994.
104. Hernandez-Gea V, Ghiassi-Nejad Z, Rozenfeld R, Gordon R, Fiel MI, 126. Jezequel AM, Novelli G, Venturini C, Orlandi F. Quantitative analysis
Yue Z, Czaja MJ, Friedman SL. Autophagy releases lipid that promotes of the perisinusoidal cells in human liver; the lipocytes. Front Gastroin-
fibrogenesis by activated hepatic stellate cells in mice and in human testinal Res 8: 85-90, 1984.
tissues. Gastroenterology 142: 938-946, 2012. 127. Ji J, Yu F, Ji Q, Li Z, Wang K, Zhang J, Lu J, Chen L, Qun E, Zeng
105. Hernandez-Gea V, Hilscher M, Rozenfeld R, Lim MP, Nieto N, Werner Y, Ji Y. Comparative proteomic analysis of rat hepatic stellate cell
S, Devi LA, Friedman SL. Endoplasmic reticulum stress induces fibro- activation: A comprehensive view and suppressed immune response.
genic activity in hepatic stellate cells through autophagy. J Hepatol 59: Hepatology 56: 332-349, 2012.
98-104, 2013. 128. Ji J, Zhang J, Huang G, Qian J, Wang X, Mei S. Over-expressed
106. Heymann F, Trautwein C, Tacke F. Monocytes and macrophages as microRNA-27a and 27b influence fat accumulation and cell prolifera-
cellular targets in liver fibrosis. Inflamm Allergy Drug Targets 8: 307- tion during rat hepatic stellate cell activation. FEBS Lett 583: 759-766,
318, 2009. 2009.
107. Holt AP, Haughton EL, Lalor PF, Filer A, Buckley CD, Adams DH. 129. Jiang JX, Chen X, Hsu DK, Baghy K, Serizawa N, Scott F, Takada
Liver myofibroblasts regulate infiltration and positioning of lympho- Y, Fukada H, Chen J, Devaraj S, Adamson R, Liu FT, Torok NJ.
cytes in human liver. Gastroenterology 136: 705-714, 2009. Galectin-3 modulates phagocytosis-induced stellate cell activation and
108. Hong F, Tuyama A, Lee TF, Loke J, Agarwal R, Cheng X, Garg A, liver fibrosis in vivo. Am J Physiol Gastrointest Liver Physiol 302:
Fiel MI, Schwartz M, Walewski J, Branch A, Schecter AD, Bansal MB. G439-G446, 2012.
Hepatic stellate cells express functional CXCR4: Role in stromal cell- 130. Jiang Z, Chen Y, Feng X, Jiang J, Chen T, Xie H, Zhou L, Zheng S.
derived factor-1alpha-mediated stellate cell activation. Hepatology 49: Hepatic stellate cells promote immunotolerance following orthotopic
2055-2067, 2009. liver transplantation in rats via induction of T cell apoptosis and reg-
109. Hsieh CC, Chou HS, Fung JJ, Qian S, Lu L. The role of liver stromal ulation of Th2/Th3-like cell cytokine production. Exp Ther Med 5:
cells in dendritic cells development in mice. Transplant Proc 42: 4279- 165-169, 2013.
4281, 2010. 131. Jiao J, Sastre D, Fiel MI, Lee UE, Ghiassi-Nejad Z, Ginhoux F, Vivier E,
110. Huang G, Brigstock DR. Integrin expression and function in the Friedman SL, Merad M, Aloman C. Dendritic cell regulation of carbon
response of primary culture hepatic stellate cells to connective tissue tetrachloride-induced murine liver fibrosis regression. Hepatology 55:
growth factor (CCN2). J Cell Mol Med 15: 1087-1095, 2011. 244-255, 2012.
111. Huang G, Brigstock DR. Regulation of hepatic stellate cells by connec- 132. Kageyama Y, Ikeda H, Watanabe N, Nagamine M, Kusumoto Y,
tive tissue growth factor. Front Biosci 17: 2495-2507, 2012. Yashiro M, Satoh Y, Shimosawa T, Shinozaki K, Tomiya T, Inoue
112. Iizuka M, Ogawa T, Enomoto M, Motoyama H, Yoshizato K, Ikeda K, Y, Nishikawa T, Ohtomo N, Tanoue Y, Yokota H, Koyama T, Ishimaru
Kawada N. Induction of microRNA-214-5p in human and rodent liver K, Okamoto Y, Takuwa Y, Koike K, Yatomi Y. Antagonism of sph-
fibrosis. Fibrogenesis Tissue Repair 5: 12, 2012. ingosine 1-phosphate receptor 2 causes a selective reduction of portal

Volume 3, October 2013 1487


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

vein pressure in bile duct-ligated rodents. Hepatology 56: 1427-1438, 154. Kordes C, Sawitza I, Muller-Marbach A, Ale-Agha N, Keitel V,
2012. Klonowski-Stumpe H, Haussinger D. CD133+ hepatic stellate cells
133. Kalinichenko VV, Bhattacharyya D, Zhou Y, Gusarova GA, Kim W, are progenitor cells. Biochem Biophys Res Commun 352: 410-417,
Shin B, Costa RH. Foxf1+/− mice exhibit defective stellate cell activa- 2007.
tion and abnormal liver regeneration following CCl4 injury. Hepatology 155. Krizhanovsky V, Yon M, Dickins RA, Hearn S, Simon J, Miething C,
37: 107-117, 2003. Yee H, Zender L, Lowe SW. Senescence of activated stellate cells limits
134. Kang N, Gores GJ, Shah VH. Hepatic stellate cells: Partners in crime liver fibrosis. Cell 134: 657-667, 2008.
for liver metastases? Hepatology 54: 707-713, 2011. 156. Kruglov EA, Nathanson RA, Nguyen T, Dranoff JA. Secretion of MCP-
135. Kaur S, Tripathi D, Dongre K, Garg V, Rooge S, Mukopadhyay A, 1/CCL2 by bile duct epithelia induces myofibroblastic transdifferentia-
Sakhuja P, Sarin SK. Increased number and function of endothelial tion of portal fibroblasts. Am J Physiol Gastrointest Liver Physiol 290:
progenitor cells stimulate angiogenesis by resident liver sinusoidal G765-G771, 2006.
endothelial cells (SECs) in cirrhosis through paracrine factors. J Hep- 157. Kubota H, Yao HL, Reid LM. Identification and characterization of
atol 57: 1193-1198, 2012. vitamin A-storing cells in fetal liver: Implications for functional impor-
136. Kawashima R, Mochida S, Matsui A, YouLuTu ZY, Ishikawa K, tance of hepatic stellate cells in liver development and hematopoiesis.
Toshima K, Yamanobe F, Inao M, Ikeda H, Ohno A, Nagoshi S, Uede Stem Cells 25: 2339-2349, 2007.
T, Fujiwara K. Expression of osteopontin in Kupffer cells and hepatic 158. Laleman W, Van Landeghem L, Severi T, Vander Elst I, Zeegers M,
macrophages and Stellate cells in rat liver after carbon tetrachloride Bisschops R, Van Pelt J, Roskams T, Cassiman D, Fevery J, Nevens F.
intoxication: A possible factor for macrophage migration into hepatic Both Ca2+ -dependent and -independent pathways are involved in rat
necrotic areas. Biochem Biophys Res Commun 256: 527-531, 1999. hepatic stellate cell contraction and intrahepatic hyperresponsiveness
137. Kendall TJ, Hennedige S, Aucott RL, Hartland SN, Vernon MA, to methoxamine. Am J Physiol Gastrointest Liver Physiol 292: G556-
Benyon RC, Iredale JP. p75 Neurotrophin receptor signaling regu- G564, 2007.
lates hepatic myofibroblast proliferation and apoptosis in recovery from 159. Lee JS, Semela D, Iredale J, Shah VH. Sinusoidal remodeling and
rodent liver fibrosis. Hepatology 49: 901-910, 2009. angiogenesis: A new function for the liver-specific pericyte? Hepatol-
138. Khimji AK, Rockey DC. Endothelin and hepatic wound healing. Phar- ogy 45: 817-825, 2007.
macol Res 63: 512-518, 2011. 160. Lee SJ, Friedman SL, Whalen R, Boyer TD. Cellular sources of glu-
139. Kida Y, Xia Z, Zheng S, Mordwinkin NM, Louie SG, Zheng SG, Feng tathione S-transferase P in primary cultured rat hepatocytes: Localiza-
M, Shi H, Duan Z, Han YP. Interleukin-1 as an injury signal mobilizes tion by in situ hybridization. Biochem J 299: 79-83, 1994.
retinyl esters in hepatic stellate cells through down regulation of lecithin 161. Lee TF, Mak KM, Rackovsky O, Lin YL, Kwong AJ, Loke JC, Fried-
retinol acyltransferase. PLoS One 6: e26644, 2011. man SL. Downregulation of hepatic stellate cell activation by retinol and
140. Kikuchi S, Griffin CT, Wang SS, Bissell DM. Role of CD44 in epithelial palmitate mediated by adipose differentiation-related protein (ADRP).
wound repair: Migration of rat hepatic stellate cells utilizes hyaluronic J Cell Physiol 223: 648-657, 2010.
acid and CD44v6. J Biol Chem 280: 15398-15404, 2005. 162. Lemoinne S, Cadoret A, Mourabit HE, Thabut D, Housset C. Origins
141. Kinnman N, Francoz C, Barbu V, Wendum D, Rey C, Hultcrantz R, and functions of liver myofibroblasts. Biochim Biophys Acta 1832:
Poupon R, Housset C. The myofibroblastic conversion of peribiliary 948-954, 2013.
fibrogenic cells distinct from hepatic stellate cells is stimulated by 163. Li J, Ghazwani M, Zhang Y, Lu J, Fan J, Gandhi CR, Li S. miR-122
platelet-derived growth factor during liver fibrogenesis. Lab Invest 83: regulates collagen production via targeting hepatic stellate cells and
163-173, 2003. suppressing P4HA1 expression. J Hepatol 58: 522-528, 2013.
142. Kinnman N, Hultcrantz R, Barbu V, Rey C, Wendum D, Poupon R, 164. Li T, Shi Z, Rockey DC. Preproendothelin-1 expression is negatively
Housset C. PDGF-mediated chemoattraction of hepatic stellate cells by regulated by IFNgamma during hepatic stellate cell activation. Am J
bile duct segments in cholestatic liver injury. Lab Invest 80: 697-707, Physiol Gastrointest Liver Physiol 302: G948-G957, 2012.
2000. 165. Li Y, Wang J, Asahina K. Mesothelial cells give rise to hepatic stellate
143. Kisseleva T, Cong M, Paik Y, Scholten D, Jiang C, Benner C, Iwaisako cells and myofibroblasts via mesothelial-mesenchymal transition in
K, Moore-Morris T, Scott B, Tsukamoto H, Evans SM, Dillmann W, liver injury. Proc Natl Acad Sci U S A 110: 2324-2329, 2013.
Glass CK, Brenner DA. Myofibroblasts revert to an inactive phenotype 166. Libbrecht L, Cassiman D, Desmet V, Roskams T. The correlation
during regression of liver fibrosis. Proc Natl Acad Sci U S A 109: between portal myofibroblasts and development of intrahepatic bile
9448-9453, 2012. ducts and arterial branches in human liver. Liver 22: 252-258, 2002.
144. Kluwe J, Wongsiriroj N, Troeger JS, Gwak GY, Dapito DH, Pradere JP, 167. Lindquist J, Stefanovic B, Brenner D. Regulation of collagen alpha1(I)
Jiang H, Siddiqi M, Piantedosi R, O’Byrne SM, Blaner WS, Schwabe expression in hepatic stellate cells. J Gastroenterol 35(Suppl 12): 80-83,
RF. Absence of hepatic stellate cell retinoid lipid droplets does not 2000.
enhance hepatic fibrosis but decreases hepatic carcinogenesis. Gut 60: 168. Liu Z, Rossen EV, Timmermans JP, Geerts A, van Grunsven LA,
1260-1268, 2011. Reynaert H. Distinct roles for non-muscle myosin II isoforms in mouse
145. Knittel T, Mehde M, Kobold D, Saile B, Dinter C, Ramadori G. hepatic stellate cells. J Hepatol 54: 132-141, 2011.
Expression patterns of matrix metalloproteinases and their inhibitors 169. Lujambio A, Akkari L, Simon J, Grace D, Tschaharganeh DF, Bolden
in parenchymal and non-parenchymal cells of rat liver: Regulation by JE, Zhao Z, Thapar V, Joyce JA, Krizhanovsky V, Lowe SW. Non-Cell-
TNF-alpha and TGF-beta1. J Hepatol 30: 48-60, 1999. Autonomous Tumor Suppression by p53. Cell 153: 449-460, 2013.
146. Knolle PA, Gerken G. Local control of the immune response in the 170. Magness ST, Bataller R, Yang L, Brenner DA. A dual reporter gene
liver. Immunol Rev 174: 21-34, 2000. transgenic mouse demonstrates heterogeneity in hepatic fibrogenic cell
147. Kodama T, Takehara T, Hikita H, Shimizu S, Li W, Miyagi T, Hosui populations. Hepatology 40: 1151-1159, 2004.
A, Tatsumi T, Ishida H, Tadokoro S, Ido A, Tsubouchi H, Hayashi 171. Maher JJ, McGuire RF. Extracellular matrix gene expression increases
N. Thrombocytopenia exacerbates cholestasis-induced liver fibrosis in preferentially in rat lipocytes and sinusoidal endothelial cells during
mice. Gastroenterology 138: 2487-2498, 2498 e2481-2487, 2010. hepatic fibrosis in vivo. J Clin Invest 86: 1641-1648, 1990.
148. Kojima N, Sato M, Imai K, Miura M, Matano Y, Senoo H. Hepatic 172. Mann J, Chu DC, Maxwell A, Oakley F, Zhu NL, Tsukamoto H, Mann
stellate cells (vitamin A-storing cells) change their cytoskeleton struc- DA. MeCP2 controls an epigenetic pathway that promotes myofibrob-
ture by extracellular matrix components through a signal transduction last transdifferentiation and fibrosis. Gastroenterology 138: 705-714,
system. Histochem Cell Biol 110: 121-128, 1998. 714 e701-704, 2010.
149. Kojima S, Hayashi S, Shimokado K, Suzuki Y, Shimada J, Crippa MP, 173. Mann J, Mann DA. Transcriptional regulation of hepatic stellate cells.
Friedman SL. Transcriptional activation of urokinase by the Kruppel- Adv Drug Deliv Rev 61: 497-512, 2009.
like factor Zf9/COPEB activates latent TGF-beta1 in vascular endothe- 174. Mann J, Oakley F, Akiboye F, Elsharkawy A, Thorne AW, Mann DA.
lial cells. Blood 95: 1309-1316, 2000. Regulation of myofibroblast transdifferentiation by DNA methylation
150. Komatsu M, Waguri S, Chiba T, Murata S, Iwata J, Tanida I, Ueno T, and MeCP2: Implications for wound healing and fibrogenesis. Cell
Koike M, Uchiyama Y, Kominami E, Tanaka K. Loss of autophagy in Death Differ 14: 275-285, 2007.
the central nervous system causes neurodegeneration in mice. Nature 175. Manojlovic Z, Stefanovic B. A novel role of RNA helicase A in reg-
441: 880-884, 2006. ulation of translation of type I collagen mRNAs. RNA 18: 321-334,
151. Komatsu M, Wang QJ, Holstein GR, Friedrich VL, Jr., Iwata J, Komi- 2012.
nami E, Chait BT, Tanaka K, Yue Z. Essential role for autophagy protein 176. Marra F, Romanelli RG, Giannini C, Failli P, Pastacaldi S, Arrighi
Atg7 in the maintenance of axonal homeostasis and the prevention of MC, Pinzani M, Laffi G, Montalto P, Gentilini P. Monocyte chemo-
axonal degeneration. Proc Natl Acad Sci U S A 104: 14489-14494, tactic protein-1 as a chemoattractant for human hepatic stellate cells.
2007. Hepatology 29: 140-148, 1999.
152. Kong X, Horiguchi N, Mori M, Gao B. Cytokines and STATs in liver 177. Marsden ER, Hu Z, Fujio K, Nakatsukasa H, Thorgeirsson SS, Evarts
fibrosis. Front Physiol 3: 69, 2012. RP. Expression of acidic fibroblast growth factor in regenerating liver
153. Kordes C, Sawitza I, Gotze S, Haussinger D. Stellate cells from rat and during hepatic differentiation. Lab Invest 67: 427-433, 1992.
pancreas are stem cells and can contribute to liver regeneration. PLoS 178. Matsumoto K, Miki R, Nakayama M, Tatsumi N, Yokouchi Y.
One 7: e51878, 2012. Wnt9a secreted from the walls of hepatic sinusoids is essential for

1488 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

morphogenesis, proliferation, and glycogen accumulation of chick hep- 199. Novo E, Cannito S, Zamara E, Valfre di Bonzo L, Caligiuri A, Cra-
atic epithelium. Dev Biol 319: 234-247, 2008. vanzola C, Compagnone A, Colombatto S, Marra F, Pinzani M, Parola
179. Mehal WZ, Friedman SL. The role of inflammation and immunity in the M. Proangiogenic cytokines as hypoxia-dependent factors stimulating
pathogenesis of liver fibrosis. In: Gershwin ME, Veirling JM, Manns migration of human hepatic stellate cells. Am J Pathol 170: 1942-1953,
MP, editors. Liver Immunology. Totowa, New Jersey: Humana Press, 2007.
2007, p. 99-109. 200. Novo E, Povero D, Busletta C, Paternostro C, di Bonzo LV, Cannito S,
180. Melton AC, Yee HF. Hepatic stellate cell protrusions couple platelet- Compagnone A, Bandino A, Marra F, Colombatto S, David E, Pinzani
derived growth factor-BB to chemotaxis. Hepatology 45: 1446-1453, M, Parola M. The biphasic nature of hypoxia-induced directional migra-
2007. tion of activated human hepatic stellate cells. J Pathol 226: 588-597,
181. Meng F, Wang K, Aoyama T, Grivennikov SI, Paik Y, Scholten D, 2012.
Cong M, Iwaisako K, Liu X, Zhang M, Osterreicher CH, Stickel F, Ley 201. Novobrantseva TI, Majeau GR, Amatucci A, Kogan S, Brenner I, Casola
K, Brenner DA, Kisseleva T. Interleukin-17 signaling in inflammatory, S, Shlomchik MJ, Koteliansky V, Hochman PS, Ibraghimov A. Attenu-
Kupffer cells, and hepatic stellate cells exacerbates liver fibrosis in ated liver fibrosis in the absence of B cells. J Clin Invest 115: 3072-3082,
mice. Gastroenterology 143: 765-776 e761-763, 2012. 2005.
182. Meyer DH, Bachem MG, Gressner AM. Modulation of hepatic lipocyte 202. Oakley F, Meso M, Iredale JP, Green K, Marek CJ, Zhou X, May MJ,
proteoglycan synthesis and proliferation by Kupffer cell-derived trans- Millward-Sadler H, Wright MC, Mann DA. Inhibition of inhibitor of
forming growth factors type beta 1 and type alpha. Biochem Biophys kappaB kinases stimulates hepatic stellate cell apoptosis and acceler-
Res Commun 171: 1122-1129, 1990. ated recovery from rat liver fibrosis. Gastroenterology 128: 108-120,
183. Milani S, Herbst H, Schuppan D, Grappone C, Pellegrini G, Pinzani 2005.
M, Casini A, Calabró A, Ciancio G, Stefanini F, Ciancio AK, Surrenti 203. Oakley F, Trim N, Constandinou CM, Ye W, Gray AM, Frantz G,
C. Differential expression of matrix-metalloproteinase-1 and -2 genes Hillan K, Kendall T, Benyon RC, Mann DA, Iredale JP. Hepatocytes
in normal and fibrotic human liver. Am J Pathol 144: 528-537, 1994. express nerve growth factor during liver injury: Evidence for paracrine
184. Miura K, Nagai H, Ueno Y, Goto T, Mikami K, Nakane K, Yoneyama regulation of hepatic stellate cell apoptosis. Am J Pathol 163: 1849-
K, Watanabe D, Terada K, Sugiyama T, Imai K, Senoo H, Watanabe S. 1858, 2003.
Epimorphin is involved in differentiation of rat hepatic stem-like cells 204. Odena G, Bataller R. Actin-binding proteins as molecular targets to
through cell-cell contact. Biochem Biophys Res Commun 311: 415-423, modulate hepatic stellate cell proliferation. Focus on “MARCKS actin-
2003. binding capacity mediates actin filament assembly during mitosis in
185. Miura K, Yang L, van Rooijen N, Ohnishi H, Seki E. Hepatic recruit- human hepatic stellate cells”. Am J Physiol Cell Physiol 303: C355-
ment of macrophages promotes nonalcoholic steatohepatitis through C356, 2012.
CCR2. Am J Physiol Gastrointest Liver Physiol 302: G1310-G1321, 205. Ogawa T, Enomoto M, Fujii H, Sekiya Y, Yoshizato K, Ikeda K,
2012. Kawada N. MicroRNA-221/222 upregulation indicates the activation
186. Miyata E, Masuya M, Yoshida S, Nakamura S, Kato K, Sugimoto Y, of stellate cells and the progression of liver fibrosis. Gut 61: 1600-1609,
Shibasaki T, Yamamura K, Ohishi K, Nishii K, Ishikawa F, Shiku H, 2012.
Katayama N. Hematopoietic origin of hepatic stellate cells in the adult 206. Ohata M, Lin M, Satre M, Tsukamoto H. Diminished retinoic acid
liver. Blood 111: 2427-2435, 2008. signaling in hepatic stellate cells in cholestatic liver fibrosis. Am J
187. Modi AA, Feld JJ, Park Y, Kleiner DE, Everhart JE, Liang TJ, Hoof- Physiol 272: G589-G596, 1997.
nagle JH. Increased caffeine consumption is associated with reduced 207. Okuno M, Moriwaki H, Imai S, Muto Y, Kawada N, Suzuki Y, Kojima
hepatic fibrosis. Hepatology 51: 201-209, 2010. S. Retinoids exacerbate rat liver fibrosis by inducing the activation of
188. Moriwaki H, Blaner WS, Piantedosi R, Goodman DS. Effects of dietary latent TGF-beta in liver stellate cells. Hepatology 26: 913-921, 1997.
retinoid and triglyceride on the lipid composition of rat liver stellate 208. Olaso E, Arteta B, Benedicto A, Crende O, Friedman SL. Loss of dis-
cells and stellate cell lipid droplets. J Lipid Res 29: 1523-1534, 1988. coidin domain receptor 2 promotes hepatic fibrosis after chronic car-
189. Muhanna N, Abu Tair L, Doron S, Amer J, Azzeh M, Mahamid M, bon tetrachloride through altered paracrine interactions between hep-
Friedman S, Safadi R. Amelioration of hepatic fibrosis by NK cell atic stellate cells and liver-associated macrophages. Am J Pathol 179:
activation. Gut 60: 90-98, 2011. 2894-2904, 2011.
190. Muhanna N, Doron S, Wald O, Horani A, Eid A, Pappo O, Friedman 209. Olaso E, Salado C, Egilegor E, Gutierrez V, Santisteban A, Sancho-Bru
SL, Safadi R. Activation of hepatic stellate cells after phagocytosis of P, Friedman SL, Vidal-Vanaclocha F. Proangiogenic role of tumor-
lymphocytes: A novel pathway of fibrogenesis. Hepatology 48: 963- activated hepatic stellate cells in experimental melanoma metastasis.
977, 2008. Hepatology 37: 674-685, 2003.
191. Mullhaupt B, Feren A, Fodor E, Jones A. Liver expression of epidermal 210. Olsen AL, Sackey BK, Marcinkiewicz C, Boettiger D, Wells RG.
growth factor RNA. Rapid increases in immediate-early phase of liver Fibronectin extra domain-A promotes hepatic stellate cell motility but
regeneration. J Biol Chem 269: 19667-19670, 1994. not differentiation into myofibroblasts. Gastroenterology 142: 928-937
192. Murphy FR, Issa R, Zhou X, Ratnarajah S, Nagase H, Arthur MJ, e923, 2012.
Benyon C, Iredale JP. Inhibition of apoptosis of activated hep- 211. Ong A, Wong VW, Wong GL, Chan HL. The effect of caffeine and
atic stellate cells by tissue inhibitor of metalloproteinase-1 is medi- alcohol consumption on liver fibrosis - a study of 1045 Asian hepati-
ated via effects on matrix metalloproteinase inhibition: Implications tis B patients using transient elastography. Liver Int 31: 1047-1053,
for reversibility of liver fibrosis. J Biol Chem 277: 11069-11076, 2011.
2002. 212. Paik YH, Iwaisako K, Seki E, Inokuchi S, Schnabl B, Osterreicher
193. Nieto N, Friedman SL, Cederbaum AI. Cytochrome P450 2E1-derived CH, Kisseleva T, Brenner DA. The nicotinamide adenine dinucleotide
reactive oxygen species mediate paracrine stimulation of collagen I phosphate oxidase (NOX) homologues NOX1 and NOX2/gp91(phox)
protein synthesis by hepatic stellate cells. J Biol Chem 277: 9853-9864, mediate hepatic fibrosis in mice. Hepatology 53: 1730-1741, 2011.
2002. 213. Paik YH, Schwabe RF, Bataller R, Russo MP, Jobin C, Brenner
194. Niiya M, Uemura M, Zheng XW, Pollak ES, Dockal M, Scheiflinger DA. Toll-like receptor 4 mediates inflammatory signaling by bacte-
F, Wells RG, Zheng XL. Increased ADAMTS-13 proteolytic activity in rial lipopolysaccharide in human hepatic stellate cells. Hepatology 37:
rat hepatic stellate cells upon activation in vitro and in vivo. J Thromb 1043-1055, 2003.
Haemost 4: 1063-1070, 2006. 214. Parsons CJ, Stefanovic B, Seki E, Aoyama T, Latour AM, Marzluff WF,
195. Niki T, De Bleser PJ, Xu G, Van Den Berg K, Wisse E, Geerts A. Rippe RA, Brenner DA. Mutation of the 5 -untranslated region stem-
Comparison of glial fibrillary acidic protein and desmin staining in loop structure inhibits alpha1(I) collagen expression in vivo. J Biol
normal and CCl4-induced fibrotic rat livers. Hepatology 23: 1538-1545, Chem 286: 8609-8619, 2011.
1996. 215. Passino MA, Adams RA, Sikorski SL, Akassoglou K. Regulation
196. Noetel A, Elfimova N, Altmuller J, Becker C, Becker D, Lahr W, Nurn- of hepatic stellate cell differentiation by the neurotrophin receptor
berg P, Wasmuth H, Teufel A, Buttner R, Dienes HP, Odenthal M. Next p75NTR. Science 315: 1853-1856, 2007.
generation sequencing of the Ago2 interacting transcriptome identified 216. Patsenker E, Stickel F. Role of integrins in fibrosing liver diseases. Am J
chemokine family members as novel targets of neuronal microRNAs in Physiol Gastrointest Liver Physiol 301: G425-G434, 2011.
hepatic stellate cells. J Hepatol 58: 335-341, 2013. 217. Perez-Tamayo R. Cirrhosis of the liver: A reversible disease? Pathol
197. Noetel A, Kwiecinski M, Elfimova N, Huang J, Odenthal M. microRNA Annu 14(Pt 2): 183-213, 1979.
are central players in anti- and profibrotic gene regulation during liver 218. Perugorria MJ, Wilson CL, Zeybel M, Walsh M, Amin S, Robinson S,
fibrosis. Front Physiol 3: 49, 2012. White SA, Burt AD, Oakley F, Tsukamoto H, Mann DA, Mann J. His-
198. Novo E, Busletta C, Bonzo LV, Povero D, Paternostro C, Mareschi K, tone methyltransferase ASH1 orchestrates fibrogenic gene transcription
Ferrero I, David E, Bertolani C, Caligiuri A, Cannito S, Tamagno E, during myofibroblast transdifferentiation. Hepatology 56: 1129-1139,
Compagnone A, Colombatto S, Marra F, Fagioli F, Pinzani M, Parola 2012.
M. Intracellular reactive oxygen species are required for directional 219. Pintilie DG, Shupe TD, Oh SH, Salganik SV, Darwiche H, Petersen
migration of resident and bone marrow-derived hepatic pro-fibrogenic BE. Hepatic stellate cells’ involvement in progenitor-mediated liver
cells. J Hepatol 54: 964-974, 2011. regeneration. Lab Invest 90: 1199-1208, 2010.

Volume 3, October 2013 1489


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

220. Pinzani M. PDGF and signal transduction in hepatic stellate cells. Front 243. Roderburg C, Luedde M, Cardenas DV, Vucur M, Mollnow T, Zim-
Biosci 7: d1720-d1726, 2002. mermann HW, Koch A, Hellerbrand C, Weiskirchen R, Frey N, Tacke
221. Pinzani M, Milani S, De FR, Grappone C, Caligiuri A, Gentilini A, F, Trautwein C, Luedde T. miR-133a mediates TGF-beta-dependent
Tosti GC, Maggi M, Failli P, Ruocco C, Gentilini P. Endothelin 1 is de-repression of collagen-synthesis in hepatic stellate cells during liver
overexpressed in human cirrhotic liver and exerts multiple effects on fibrosis. J Hepatol 58: 736-742, 2012.
activated hepatic stellate cells. Gastroenterology 110: 534-548, 1996. 244. Roderburg C, Urban GW, Bettermann K, Vucur M, Zimmermann
222. Popov Y, Patsenker E, Stickel F, Zaks J, Bhaskar KR, Niedobitek G, H, Schmidt S, Janssen J, Koppe C, Knolle P, Castoldi M, Tacke F,
Kolb A, Friess H, Schuppan D. Integrin alphavbeta6 is a marker of the Trautwein C, Luedde T. Micro-RNA profiling reveals a role for miR-
progression of biliary and portal liver fibrosis and a novel target for 29 in human and murine liver fibrosis. Hepatology 53: 209-218, 2011.
antifibrotic therapies. J Hepatol 48: 453-464, 2008. 245. Rojkind M, Giambrone MA, Biempica L. Collagen types in normal and
223. Popper H, Schaffner F. Hepatic cirrhosis. A problem in communication. cirrhotic liver. Gastroenterology 76: 710-719, 1979.
Isr J Med Sci 4: 1-7, 1968. 246. Rombouts K, Mello T, Liotta F, Galli A, Caligiuri A, Annunziato F,
224. Popper H, Udenfriend S. Hepatic fibrosis. Correlation of biologic and Pinzani M. MARCKS actin-binding capacity mediates actin filament
morphologic investigations. Am J Med 49: 707-721, 1970. assembly during mitosis in human hepatic stellate cells. Am J Physiol
225. Putz-Bankuti C, Pilz S, Stojakovic T, Scharnagl H, Pieber TR, Cell Physiol 303: C357-C367, 2012.
Trauner M, Obermayer-Pietsch B, Stauber RE. Association of 25- 247. Rosenbaum J, Blazejewski S, Preaux AM, Mallat A, Dhumeaux D,
hydroxyvitamin D levels with liver dysfunction and mortality in chronic Mavier P. Fibroblast growth factor 2 and transforming growth factor
liver disease. Liver Int 32: 845-851, 2012. beta 1 interactions in human liver myofibroblasts. Gastroenterology
226. Qin L, Han YP. Epigenetic repression of matrix metalloproteinases in 109: 1986-1996, 1995.
myofibroblastic hepatic stellate cells through histone deacetylases 4: 248. Roskams T. Different types of liver progenitor cells and their niches.
Implication in tissue fibrosis. Am J Pathol 177: 1915-1928, 2010. J Hepatol 45: 1-4, 2006.
227. Racine-Samson L, Rockey DC, Bissell DM. Expression of the collagen- 249. Roskams T, Cassiman D, De Vos R, Libbrecht L. Neuroregulation of
binding integrins alpha1 beta1 and alpha 2 beta1 during activation of the neuroendocrine compartment of the liver. Anat Rec A Discov Mol
rat hepatic stellate cells in vivo. J Biol Chem 272(49): 30911-30917, Cell Evol Biol 280: 910-923, 2004.
1997. 250. Sahin H, Trautwein C, Wasmuth HE. Functional role of chemokines in
228. Radaeva S, Sun R, Jaruga B, Nguyen VT, Tian Z, Gao B. Natural liver disease models. Nat Rev Gastroenterol Hepatol 7: 682-690, 2010.
killer cells ameliorate liver fibrosis by killing activated stellate cells 251. Saile B, Matthes N, Knittel T, Ramadori G. Transforming growth factor
in NKG2D-dependent and tumor necrosis factor-related apoptosis- beta and tumor necrosis factor alpha inhibit both apoptosis and prolif-
inducing ligand-dependent manners. Gastroenterology 130: 435-452, eration of activated rat hepatic stellate cells. Hepatology 30: 196-202,
2006. 1999.
229. Radbill BD, Gupta R, Ramirez MC, DiFeo A, Martignetti JA, Alvarez 252. Sato M, Kojima N, Miura M, Imai K, Senoo H. Induction of cellular
CE, Friedman SL, Narla G, Vrabie R, Bowles R, Saiman Y, Bansal MB. processes containing collagenase and retinoid by integrin-binding to
Loss of matrix metalloproteinase-2 amplifies murine toxin-induced interstitial collagen in hepatic stellate cell culture. Cell Biol Int 22:
liver fibrosis by upregulating collagen I expression. Dig Dis Sci 56: 115-125, 1998.
406-416, 2011. 253. Sauvant P, Cansell M, Atgie C. Vitamin A and lipid metabolism: Rela-
230. Ramachandran P, Iredale JP. Macrophages: Central regulators of hep- tionship between hepatic stellate cells (HSCs) and adipocytes. J Physiol
atic fibrogenesis and fibrosis resolution. J Hepatol 56: 1417-1419, Biochem 67: 487-496, 2011.
2012. 254. Sawitza I, Kordes C, Reister S, Haussinger D. The niche of stellate cells
231. Ramachandran P, Pellicoro A, Vernon MA, Boulter L, Aucott RL, within rat liver. Hepatology 50: 1617-1624, 2009.
Ali A, Hartland SN, Snowdon VK, Cappon A, Gordon-Walker TT, 255. Schildberg FA, Wojtalla A, Siegmund SV, Endl E, Diehl L, Abdullah
Williams MJ, Dunbar DR, Manning JR, van Rooijen N, Fallowfield Z, Kurts C, Knolle PA. Murine hepatic stellate cells veto CD8 T cell
JA, Forbes SJ, Iredale JP. Differential Ly-6C expression identifies the activation by a CD54-dependent mechanism. Hepatology 54: 262-272,
recruited macrophage phenotype, which orchestrates the regression of 2011.
murine liver fibrosis. Proc Natl Acad Sci U S A 109: E3186-E3195, 256. Schirmacher P, Geerts A, Pietrangelo A, Dienes HP, Rogler CE. Hep-
2012. atocyte growth factor/hepatopoietin A is expressed in fat-storing cells
232. Ramadori G, Neubauer K, Odenthal M, Nakamura T, Knittel T, from rat liver but not myofibroblast-like cells derived from fat-storing
Schwogler S, Meyer zum Buschenfelde KH. The gene of hepatocyte cells. Hepatology 15: 5-11, 1992.
growth factor is expressed in fat-storing cells of rat liver and is down- 257. Schmitt-Graff A, Desmouliere A, Gabbiani G. Heterogeneity of myofi-
regulated during cell growth and by transforming growth factor-beta. broblast phenotypic features: An example of fibroblastic cell plasticity.
BBRC 183: 739-742, 1992. Virchows Archiv 425: 3-24, 1994.
233. Ramadori G, Rieder H, Theiss F, Meyer zum Buschenfelde KH. Fat- 258. Schnabl B, Purbeck CA, Choi YH, Hagedorn CH, Brenner D. Replica-
storing (Ito) cells of rat liver synthesize and secrete apolipoproteins: tive senescence of activated human hepatic stellate cells is accompanied
Comparison with hepatocytes. Gastroenterology 97: 163-172, 1989. by a pronounced inflammatory but less fibrogenic phenotype. Hepatol-
234. Ramadori G, Veit T, Schwogler S, Dienes HP, Knittel T, Rieder H, ogy 37: 653-664, 2003.
Meyer zum Buschenfelde KH. Expression of the gene of the alpha 259. Schuppan D. Structure of the extracellular matrix in normal and fibrotic
smooth muscle alpha actin isoform in rat liver and in rat fat-storing liver: Collagens and glycoproteins. Semin Liver Dis 10: 1-10, 1990.
(ITO) cells. Virchows Arch B Cell Pathol Incl Mol Pathol 59: 349-357, 260. Schwabe RF, Bataller R, Brenner DA. Human hepatic stellate cells
1990. express CCR5 and RANTES to induce proliferation and migration.
235. Ramm GA, Britton RS, O’Neill R, Blaner WS, Bacon BR. Vitamin A- Am J Physiol Gastrointest Liver Physiol 285: G949-G958, 2003.
poor lipocytes: A novel desmin-negative lipocyte subpopulation, which 261. Schwettmann L, Wehmeier M, Jokovic D, Aleksandrova K, Brand K,
can be activated to myofibroblasts. Am J Physiol 269: G532-G541, Manns MP, Lichtinghagen R, Bahr MJ. Hepatic expression of A disinte-
1995. grin and metalloproteinase (ADAM) and ADAMs with thrombospondin
236. Rashid ST, Humphries JD, Byron A, Dhar A, Askari JA, Selley JN, motives (ADAM-TS) enzymes in patients with chronic liver diseases.
Knight D, Goldin RD, Thursz M, Humphries MJ. Proteomic analysis J Hepatol 49: 243-250, 2008.
of extracellular matrix from the hepatic stellate cell line LX-2 identifies 262. Seki E, De Minicis S, Gwak GY, Kluwe J, Inokuchi S, Bursill CA,
CYR61 and Wnt-5a as novel constituents of fibrotic liver. J Proteome Llovet JM, Brenner DA, Schwabe RF. CCR1 and CCR5 promote hep-
Res 11: 4052-4064, 2012. atic fibrosis in mice. J Clin Invest 119: 1858-1870, 2009.
237. Reynaert H, Thompson MG, Thomas T, Geerts A. Hepatic stellate cells: 263. Seki E, de Minicis S, Inokuchi S, Taura K, Miyai K, van Rooijen N,
Role in microcirculation and pathophysiology of portal hypertension. Schwabe RF, Brenner DA. CCR2 promotes hepatic fibrosis in mice.
Gut 50: 571-581, 2002. Hepatology 50: 185-197, 2009.
238. Rippe RA. Role of transcriptional factors in stellate cell activation. 264. Seki E, De Minicis S, Osterreicher CH, Kluwe J, Osawa Y, Brenner DA,
Alcohol Clin Exp Res 23: 926-929, 1999. Schwabe RF. TLR4 enhances TGF-beta signaling and hepatic fibrosis.
239. Rockey DC. Hepatic blood flow regulation by stellate cells in normal Nat Med 13: 1324-1332, 2007.
and injured liver. Semin Liver Dis 21: 337-350, 2001. 265. Sekiya Y, Ogawa T, Iizuka M, Yoshizato K, Ikeda K, Kawada N. Down-
240. Rockey DC. Vascular mediators in the injured liver. Hepatology 37: regulation of cyclin E1 expression by microRNA-195 accounts for
4-12, 2003. interferon-beta-induced inhibition of hepatic stellate cell proliferation.
241. Rockey DC, Boyles JK, Gabbiani G, Friedman SL. Rat hepatic lipocytes J Cell Physiol 226: 2535-2542, 2011.
express smooth muscle actin upon activation in vivo and in culture. 266. Sekiya Y, Ogawa T, Yoshizato K, Ikeda K, Kawada N. Suppression
J Submicrosc Cytol Pathol 24: 193-203, 1992. of hepatic stellate cell activation by microRNA-29b. Biochem Biophys
242. Rockey DC, Fouassier L, Chung JJ, Carayon A, Vallee P, Rey C, Hous- Res Commun 412: 74-79, 2011.
set C. Cellular localization of endothelin-1 and increased production in 267. Senoo H, Yoshikawa K, Morii M, Miura M, Imai K, Mezaki Y. Hepatic
liver injury in the rat: Potential for autocrine and paracrine effects on stellate cell (vitamin A-storing cell) and its relative – past, present and
stellate cells. Hepatology 27: 472-480, 1998. future. Cell Biol Int 34: 1247-1272, 2010.

1490 Volume 3, October 2013


Comprehensive Physiology Hepatic Stellate Cells and Liver Fibrosis

268. Seok J, Warren HS, Cuenca AG, Mindrinos MN, Baker HV, Xu W, 291. Tarrats N, Moles A, Morales A, Garcia-Ruiz C, Fernandez-Checa JC,
Richards DR, McDonald-Smith GP, Gao H, Hennessy L, Finnerty CC, Mari M. Critical role of tumor necrosis factor receptor 1, but not 2, in
Lopez CM, Honari S, Moore EE, Minei JP, Cuschieri J, Bankey PE, hepatic stellate cell proliferation, extracellular matrix remodeling, and
Johnson JL, Sperry J, Nathens AB, Billiar TR, West MA, Jeschke MG, liver fibrogenesis. Hepatology 54: 319-327, 2011.
Klein MB, Gamelli RL, Gibran NS, Brownstein BH, Miller-Graziano 292. Taura K, De Minicis S, Seki E, Hatano E, Iwaisako K, Osterreicher CH,
C, Calvano SE, Mason PH, Cobb JP, Rahme LG, Lowry SF, Maier Kodama Y, Miura K, Ikai I, Uemoto S, Brenner DA. Hepatic stellate
RV, Moldawer LL, Herndon DN, Davis RW, Xiao W, Tompkins RG. cells secrete angiopoietin 1 that induces angiogenesis in liver fibrosis.
Genomic responses in mouse models poorly mimic human inflamma- Gastroenterology 135: 1729-1738, 2008.
tory diseases. Proc Natl Acad Sci U S A 110: 3507-3512, 2013. 293. Thabut D, Routray C, Lomberk G, Shergill U, Glaser K, Huebert R,
269. Shafiei MS, Rockey DC. The function of integrin-linked kinase in Patel L, Masyuk T, Blechacz B, Vercnocke A, Ritman E, Ehman R,
normal and activated stellate cells: Implications for fibrogenesis in Urrutia R, Shah V. Complementary vascular and matrix regulatory
wound healing. Lab Invest 92: 305-316, 2012. pathways underlie the beneficial mechanism of action of sorafenib in
270. Shao R, Yan W, Rockey DC. Regulation of endothelin-1 synthesis by liver fibrosis. Hepatology 54: 573-585, 2011.
endothelin-converting enzyme-1 during wound healing. J Biol Chem 294. Troeger JS, Mederacke I, Gwak GY, Dapito DH, Mu X, Hsu CC,
274: 3228-3234, 1999. Pradere JP, Friedman RA, Schwabe RF. Deactivation of hepatic stellate
271. Shi M, Zhu J, Wang R, Chen X, Mi L, Walz T, Springer TA. Latent cells during liver fibrosis resolution in mice. Gastroenterology 143:
TGF-beta structure and activation. Nature 474: 343-349, 2011. 1073-1083 e1022, 2012.
272. Shirakami Y, Gottesman ME, Blaner WS. Diethylnitrosamine-induced 295. Tsukada S, Parsons CJ, Rippe RA. Mechanisms of liver fibrosis. Clin
hepatocarcinogenesis is suppressed in lecithin:retinol acyltransferase- Chim Acta 364: 33-60, 2006.
deficient mice primarily through retinoid actions immediately after 296. Tsukada S, Westwick JK, Ikejima K, Sato N, Rippe RA. SMAD and p38
carcinogen administration. Carcinogenesis 33: 268-274, 2012. MAPK signaling pathways independently regulate alpha1(I) collagen
273. Shomron N, Hamasaki-Katagiri N, Hunt R, Hershko K, Pommier E, gene expression in unstimulated and transforming growth factor-beta-
Geetha S, Blaisdell A, Dobkin A, Marple A, Roma I, Newell J, Allen stimulated hepatic stellate cells. J Biol Chem 280: 10055-10064, 2005.
C, Friedman S, Kimchi-Sarfaty C. A splice variant of ADAMTS13 is 297. Tsukamoto H. Epigenetic mechanism of stellate cell trans-
expressed in human hepatic stellate cells and cancerous tissues. Thromb differentiation. J Hepatol 46: 352-353, 2007.
Haemost 104: 531-535, 2010. 298. Tsukamoto H, Zhu NL, Wang J, Asahina K, Machida K. Morphogens
274. Singh R, Kaushik S, Wang Y, Xiang Y, Novak I, Komatsu M, Tanaka K, and hepatic stellate cell fate regulation in chronic liver disease. J Gas-
Cuervo AM, Czaja MJ. Autophagy regulates lipid metabolism. Nature troenterol Hepatol 27 Suppl 2: 94-98, 2012.
458: 1131-1135, 2009. 299. Ueno T, Sata M, Sakata R, Torimura T, Sakamoto M, Sugawara H,
275. Smedsrod B, Le Couteur D, Ikejima K, Jaeschke H, Kawada N, Naito Tanikawa K. Hepatic stellate cells and intralobular innervation in human
M, Knolle P, Nagy L, Senoo H, Vidal-Vanaclocha F, Yamaguchi N. liver cirrhosis. Hum Pathol 28: 953-959, 1997.
Hepatic sinusoidal cells in health and disease: Update from the 14th 300. Urtasun R, Lopategi A, George J, Leung TM, Lu Y, Wang X, Ge X,
International Symposium. Liver Int 29: 490-501, 2009. Fiel MI, Nieto N. Osteopontin, an oxidant stress sensitive cytokine,
276. Sohail MA, Hashmi AZ, Hakim W, Watanabe A, Zipprich A, Grosz- up-regulates collagen-I via integrin alpha(V)beta(3) engagement and
mann RJ, Dranoff JA, Torok NJ, Mehal WZ. Adenosine induces loss PI3K/pAkt/NFkappaB signaling. Hepatology 55: 594-608, 2012.
of actin stress fibers and inhibits contraction in hepatic stellate cells via 301. Vinas O, Bataller R, Sancho-Bru P, Gines P, Berenguer C, Enrich C,
Rho inhibition. Hepatology 49: 185-194, 2009. Nicolas JM, Ercilla G, Gallart T, Vives J, Arroyo V, Rodes J. Human
277. Sprenger H, Kaufmann A, Garn H, Lahme B, Gemsa D, Gressner AM. hepatic stellate cells show features of antigen-presenting cells and stim-
Differential expression of monocyte chemotactic protein-1 (MCP-1) in ulate lymphocyte proliferation. Hepatology 38: 919-929, 2003.
transforming rat hepatic stellate cells. J Hepatol 30: 88-94, 1999. 302. Vyas SK, Leyland H, Gentry J, Arthur MJ. Rat hepatic lipocytes syn-
278. Starkel P, Sempoux C, Leclercq I, Herin M, Deby C, Desager JP, Hors- thesize and secrete transin (stromelysin) in early primary culture. Gas-
mans Y. Oxidative stress, KLF6 and transforming growth factor-beta troenterology 109: 889-898, 1995.
up-regulation differentiate non-alcoholic steatohepatitis progressing to 303. Wake K. “Sternzellen” in the liver: Perisinuosoidal cells with special
fibrosis from uncomplicated steatosis in rats. J Hepatol 39: 538-546, reference to storage of vitamin A. Am J Anat 132: 429-462, 1971.
2003. 304. Wake K. Perisinusoidal stellate cells (Fat-storing cells, interstitial cells,
279. Stefanovic B, Hellerbrand C, Brenner DA. Regulatory role of the con- lipocytes) their related structure in and around liver sinusoids, and
served stem-loop structure at the 5 end of collagen alpha1(I) mRNA. vitamin A storing cells in extrahepatic organs. Int Rev Cytology 66:
Mol Cell Biol 19: 4334-4342, 1999. 303-353, 1980.
280. Stefanovic B, Hellerbrand C, Holcik M, Briendl M, Aliebhaber S, Bren- 305. Wake K, Sato T. Intralobular heterogeneity of perisinusoidal stellate
ner DA. Posttranscriptional regulation of collagen alpha1(I) mRNA in cells in porcine liver. Cell Tissue Res 273: 227-237, 1993.
hepatic stellate cells. Mol Cell Biol 17: 5201-5209, 1997. 306. Wandzioch E, Kolterud A, Jacobsson M, Friedman SL, Carlsson L.
281. Stefanovic L, Brenner DA, Stefanovic B. Direct hepatotoxic effect of Lhx2-/- mice develop liver fibrosis. Proc Natl Acad Sci U S A 101:
KC chemokine in the liver without infiltration of neutrophils. Exp Biol 16549-16554, 2004.
Med (Maywood) 230: 573-586, 2005. 307. Wang B, Li W, Guo K, Xiao Y, Wang Y, Fan J. miR-181b Promotes
282. Steiling H, Muhlbauer M, Bataille F, Scholmerich J, Werner S, Heller- hepatic stellate cells proliferation by targeting p27 and is elevated in the
brand C. Activated hepatic stellate cells express keratinocyte growth serum of cirrhosis patients. Biochem Biophys Res Commun 421: 4-8,
factor in chronic liver disease. Am J Pathol 165: 1233-1241, 2004. 2012.
283. Su YH, Shu KH, Hu C, Cheng CH, Wu MJ, Yu TM, Chuang YW, Huang 308. Wang J, Leclercq I, Brymora JM, Xu N, Ramezani-Moghadam M,
ST, Chen CH. Hepatic stellate cells attenuate the immune response in London RM, Brigstock D, George J. Kupffer cells mediate leptin-
renal transplant recipients with chronic hepatitis. Transplant Proc 44: induced liver fibrosis. Gastroenterology 137: 713-723, 2009.
725-729, 2012. 309. Wang L, Wang X, Chiu JD, van de Ven G, Gaarde WA, Deleve LD.
284. Suzuki K, Tanaka M, Watanabe N, Saito S, Nonaka H, Miyajima A. p75 Hepatic vascular endothelial growth factor regulates recruitment of rat
Neurotrophin receptor is a marker for precursors of stellate cells and liver sinusoidal endothelial cell progenitor cells. Gastroenterology 143:
portal fibroblasts in mouse fetal liver. Gastroenterology 135: 270-281 1555-1563 e1552, 2012.
e273, 2008. 310. Wang L, Wang X, Xie G, Hill CK, DeLeve LD. Liver sinusoidal
285. Syal G, Fausther M, Dranoff JA. Advances in cholangiocyte immuno- endothelial cell progenitor cells promote liver regeneration in rats.
biology. Am J Physiol Gastrointest Liver Physiol 303: G1077-G1086, J Clin Invest 122: 1567-1573, 2012.
2012. 311. Wang Y, Yao HL, Cui CB, Wauthier E, Barbier C, Costello MJ, Moss N,
286. Szabo G, Mandrekar P, Dolganiuc A. Innate immune response and Yamauchi M, Sricholpech M, Gerber D, Loboa EG, Reid LM. Paracrine
hepatic inflammation. Semin Liver Dis 27: 339-350, 2007. signals from mesenchymal cell populations govern the expansion and
287. Tacke F. Monocytes and Macrophages as Cellular Targets in Liver differentiation of human hepatic stem cells to adult liver fates. Hepa-
Fibrosis. Inflamm Allergy Drug Targets 8: 307-318, 2009. tology 52: 1443-1454, 2010.
288. Tacke F, Weiskirchen R. Update on hepatic stellate cells: Pathogenic 312. Wang Y, Zhang JS, Qian J, Huang GC, Chen Q. Adrenomedullin
role in liver fibrosis and novel isolation techniques. Expert Rev Gas- regulates expressions of transforming growth factor-beta1 and beta1-
troenterol Hepatol 6: 67-80, 2012. induced matrix metalloproteinase-2 in hepatic stellate cells. Int J Exp
289. Takahara T, Kojima T, Miyabayashi C, Inoue K, Sasaki H, Muragaki Y, Pathol 87: 177-184, 2006.
Ooshima A. Collagen production in fat-storing cells after carbon tetra- 313. Watanabe A, Sohail MA, Gomes DA, Hashmi A, Nagata J, Sutter-
chloride intoxication in the rat. Immunoelectron microscopic observa- wala FS, Mahmood S, Jhandier MN, Shi Y, Flavell RA, Mehal WZ.
tion of type I, type III collagens, and prolyl hydroxylase. Lab Invest 59: Inflammasome-mediated regulation of hepatic stellate cells. Am J Phys-
509-521, 1988. iol Gastrointest Liver Physiol 296: G1248-G1257, 2009.
290. Tanaka H, Leung PS, Kenny TP, Gershwin ME, Bowlus CL. Immuno- 314. Watanabe N, Ikeda H, Kume Y, Satoh Y, Kaneko M, Takai D, Tejima K,
logical orchestration of liver fibrosis. Clin Rev Allergy Immunol 43: Nagamine M, Mashima H, Tomiya T, Noiri E, Omata M, Matsumoto M,
220-229, 2012. Fujimura Y, Yatomi Y. Increased production of ADAMTS13 in hepatic

Volume 3, October 2013 1491


Hepatic Stellate Cells and Liver Fibrosis Comprehensive Physiology

stellate cells contributes to enhanced plasma ADAMTS13 activity in 326. Yin L, Lynch D, Ilic Z, Sell S. Proliferation and differentiation of
rat models of cholestasis and steatohepatitis. Thromb Haemost 102: ductular progenitor cells and littoral cells during the regeneration of the
389-396, 2009. rat liver to CCl4/2-AAF injury. Histol Histopathol 17: 65-81, 2002.
315. Wells RG. The role of matrix stiffness in hepatic stellate cell activation 327. Yoshiji H, Kuriyama S, Yoshii J, Ikenaka Y, Noguchi R, Hicklin DJ,
and liver fibrosis. J Clin Gastroenterol 39: S158-S161, 2005. Wu Y, Yanase K, Namisaki T, Yamazaki M, Tsujinoue H, Imazu H,
316. wIshikawa K, Mochida S, Mashiba S, Inao M, Matsui A, Ikeda H, Ohno Masaki T, Fukui H. Vascular endothelial growth factor and receptor
A, Shibuya M, Fujiwara K. Expressions of vascular endothelial growth interaction is a prerequisite for murine hepatic fibrogenesis. Gut 52:
factor in nonparenchymal as well as parenchymal cells in rat liver after 1347-1354, 2003.
necrosis. Biochem Biophys Res Commun 254: 587-593, 1999. 328. Yu C, Wang F, Jin C, Huang X, Miller DL, Basilico C, McKeehan WL.
317. Wong L, Yamasaki G, Johnson RJ, Friedman SL. Induction of beta- Role of fibroblast growth factor type 1 and 2 in carbon tetrachloride-
platelet-derived growth factor receptor in rat hepatic lipocytes during induced hepatic injury and fibrogenesis. Am J Pathol 163: 1653-1662,
cellular activation in vivo and in culture. J Clin Invest 94: 1563-1569, 2003.
1994. 329. Zhao Y, Wang Y, Wang Q, Liu Z, Liu Q, Deng X. Hepatic stellate
318. Wu TJ, Wang YC, Wu TH, Lee CF, Chan KM, Lee WC. Inhibition cells produce vascular endothelial growth factor via phospho-p44/42
of allogenic T-cell cytotoxicity by hepatic stellate cell via CD4(+) mitogen-activated protein kinase/cyclooxygenase-2 pathway. Mol Cell
CD25(+) Foxp3(+) regulatory T cells in vitro. Transplant Proc 44: Biochem 359: 217-223, 2012.
1055-1059, 2012. 330. Zhu NL, Asahina K, Wang J, Ueno A, Lazaro R, Miyaoka Y, Miyajima
319. Xie G, Wang X, Wang L, Atkinson RD, Kanel GC, Gaarde WA, Deleve A, Tsukamoto H. Hepatic stellate cell-derived delta-like homolog 1
LD. Role of differentiation of liver sinusoidal endothelial cells in pro- (DLK1) protein in liver regeneration. J Biol Chem 287: 10355-10367,
gression and regression of hepatic fibrosis in rats. Gastroenterology 2012.
142: 918-927 e916, 2012. 331. Zhu Q, Zou L, Jagavelu K, Simonetto DA, Huebert RC, Jiang ZD,
320. Yamada M, Blaner WS, Soprano DR, Dixon JL, Kjeldbye HM, Good- DuPont HL, Shah VH. Intestinal decontamination inhibits TLR4 depen-
man DS. Biochemical characteristics of isolated rat liver stellate cells. dent fibronectin-mediated cross-talk between stellate cells and endothe-
Hepatology 7: 1224-1229, 1987. lial cells in liver fibrosis in mice. J Hepatol 56: 893-899, 2012.
321. Yamada T, Imaoka S, Kawada N, Seki S, Kuroki T, Kobayashi K, 332. Zindy F, Lamas E, Schmidt S, Kirn A, Brechot C. Expression of insulin-
Monna T, Funae Y. Expression of cytochrome P450 isoforms in rat like growth factor II (IGF-II) and IGF-II, IGF-I and insulin receptors
hepatic stellate cells. Life Sci 61: 171-179, 1997. mRNAs in isolated non-parenchymal rat liver cells. J Hepatol 14: 30-
322. Yanagisawa M, Kurihara H, Kimura S, Tomobe Y, Kobayashi M, Mitsui 34, 1992.
Y, Yazaki Y, Goto K, Masaki T. A novel potent vasoconstrictor peptide 333. Zvibel I, Atias D, Phillips A, Halpern Z, Oren R. Thyroid hormones
produced by vascular endothelial cells. Nature 332: 411-415, 1988. induce activation of rat hepatic stellate cells through increased expres-
323. Yang C, Zeisberg M, Mosterman B, Sudhakar A, Yerramalla U, sion of p75 neurotrophin receptor and direct activation of Rho. Lab
Holthaus K, Xu L, Eng F, Afdhal N, Kalluri R. Liver fibrosis: Insights Invest 90: 674-684, 2010.
into migration of hepatic stellate cells in response to extracellular matrix
and growth factors. Gastroenterology 124: 147-159, 2003.
324. Yang L, Jung Y, Omenetti A, Witek RP, Choi S, Vandongen HM, Huang
J, Alpini GD, Diehl AM. Fate-mapping evidence that hepatic stellate
cells are epithelial progenitors in adult mouse livers. Stem Cells 26:
Further Reading
2104-2113, 2008.
325. Yin C, Evason KJ, Maher JJ, Stainier DY. The basic helix-loop-helix Friedman SL, Sheppard D, Duffield J and Violette S. Therapy for fibrotic
transcription factor, heart and neural crest derivatives expressed tran- diseases: Nearing the starting line. Sci Transl Med, 5:167sr1, 2013.
script 2, marks hepatic stellate cells in zebrafish: Analysis of stellate Friedman SL, guest editor. Special issue on fibrosis. Biochim Biophys Acta.
cell entry into the developing liver. Hepatology 56: 1958-1970, 2012. 2013 Mar 16. pii: S0925-4439(13)00077-X. doi: 10.1016/j.bbadis.

1492 Volume 3, October 2013

You might also like