You are on page 1of 14

Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Molecular Aspects of Medicine


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

Organ and tissue fibrosis: Molecular signals, cellular mechanisms and


translational implications
Ralf Weiskirchena, Sabine Weiskirchena, Frank Tackeb,∗
a
Institute of Molecular Pathobiochemistry, Experimental Gene Therapy and Clinical Chemistry, RWTH University Hospital Aachen, Germany
b
Dept. of Medicine III, University Hospital Aachen, Germany

A R T I C LE I N FO A B S T R A C T

Keywords: Fibrosis denotes excessive scarring, which exceeds the normal wound healing response to injury in many tissues.
Organ fibrosis Although the extracellular matrix deposition appears unstructured disrupting the normal tissue architecture and
Matrix subsequently impairing proper organ function, fibrogenesis is a highly orchestrated process determined by de-
Regression fined sequences of molecular signals and cellular response mechanisms. Persistent injury and parenchymal cell
Biomarkers
death provokes tissue inflammation, macrophage activation and immune cell infiltration. The release of biolo-
Imaging
gically highly active soluble mediators (alarmins, cytokines, chemokines) lead to the local activation of collagen
Fibroblasts
Exosome producing mesenchymal cells such as pericytes, myofibroblasts or Gli1 positive mesenchymal stem cell-like cells,
Macrophage to a transition of various cell types into myofibroblasts as well as to the recruitment of fibroblast precursors.
Therapy Clinical observations and experimental models highlighted that fibrosis is not a one-way road. Specific me-
chanistic principles of fibrosis regression involve the resolution of chronic tissue injury, the shift of inflammatory
processes towards recovery, deactivation of myofibroblasts and finally fibrolysis of excess matrix scaffold. The
thorough understanding of common principles of fibrogenic molecular signals and cellular mechanisms in
various organs - such as liver, kidney, lung, heart or skin - is the basis for developing improved diagnostics
including biomarkers or imaging techniques and novel antifibrotic therapeutics.

1. Introduction: principles of physiological wound healing vs. ferroptosis and others) is driven by various injurious agents and me-
pathological fibrosis chanisms. The resulting tissue damage is associated with an in-
flammatory response, in which local immune cells (mainly tissue
Fibrosis is a reparative or reactive process that is majorly char- macrophages) become activated and diverse sets of blood cells enter the
acterised by the formation and deposition of excess fibrous connective affected sites of injury. The local and invading immune cells produce a
tissue resulting in progressive architectural remodeling in nearly all large variety of biologically highly active soluble mediators (cytokines
tissues and organs (Fig. 1). This scarring process is a leading cause of and chemokines) that lead to a local activation of mesenchymal cells,
morbidity and mortality and may account for up to 45% of all causes of which have the capacity to produce ECM and to further increase pro-
death in the United States (Wynn, 2004). Mechanistically, fibrogenic duction of pro-inflammatory cytokines, chemokines, and angiogenic
responses encompass the same fundamental mechanisms that are part factors. A hallmark of these cells is their transition from a resting to an
of the normal wound healing response. However, upon scarring the activated phenotype, by which they become strongly positive for α-
respective mechanisms become exacerbated due to tissue injury. Con- smooth muscle actin (α-SMA) expression and synthesize ECM compo-
sequently, chronic fibrogenesis provokes a shift from a supportive nents. Moreover, the fraction of cells contributing to the formation of
(“good”) fibrotic tissue to a microenvironment, in which extracellular fibrillar collagens and non-structural proteins with regulatory roles in
matrix (ECM) producing cells increase in number or become excessively ECM is significantly increased by other cell populations that can transit
active, thereby forming way too much ECM resulting in substantial into myofibroblasts or matrix producing cells. Such potential progeni-
(“bad”) scar formation and destruction of normal organ architecture tors for myofibroblasts were identified during the last decades. In
(Pakshir and Hinz, 2018). particular, tissue-resident fibroblasts, circulating bone marrow-derived
In most cases, fibrosis is initiated by parenchymal cell destruction, fibrocytes, different epithelial and endothelial cell subsets that acquire
in which cell death (necrosis, apoptosis, necroptosis, pyroptosis, a myofibroblast phenotype in a process termed epithelial-to-


Corresponding author. Department of Medicine III, University Hospital Aachen; Pauwelsstrasse 30, 52074, Aachen, Germany.
E-mail address: frank.tacke@gmx.net (F. Tacke).

https://doi.org/10.1016/j.mam.2018.06.003
Received 22 March 2018; Accepted 25 June 2018
0098-2997/ © 2018 Elsevier Ltd. All rights reserved.

Please cite this article as: Weiskirchen, R., Molecular Aspects of Medicine (2018), https://doi.org/10.1016/j.mam.2018.06.003
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

and even a coordinated systemic mobilization of endocrine and neu-


rological mediators. In the affected organ the process is characterised
by typical cardinal signs of inflammation that include pain, heat, red-
ness, swelling and impairment or loss of function. Most important in the
initiation of this inflammatory response are resident immune cells al-
ready present in the (damaged) organ including macrophages, dendritic
cells, and mast cells. Commonly, these cells of the reticuloendothelial
system possess specialized surface receptors known as pattern re-
cognition receptors playing crucial roles in detecting pathogen-asso-
ciated molecular patterns (PAMPs) such as bacterial toxins and damage-
associated molecular patterns (DAMPs) released by injured cells
(Heymann and Tacke, 2016). After recognition of PAMPs and DAMPs,
respective cells release different inflammatory mediators that in turn
provoke increased blood flow, permeability of blood vessels, over-
heating of the tissue, increase the sensitivity to pain, and exudation of
plasma proteins and fluid into the tissue. These alterations manifest in
the mentioned cardinal signs of inflammation. In addition to the release
of inflammatory mediators, the complement system becomes activated.
This system consists of a number of small blood proteins that after
proteolytic activation attract phagocytes and stimulate them to clear
harmful or damaged material (Danobeitia et al., 2014). Likewise, the
coagulation cascade with capacity to physically trap invading microbes
in blood clots and the fibrinolysis system that reorganise and resorb
blood clots are stimulated. When all these responses are sufficient to
defend against injury, the soluble mediators are degraded and tissue
homeostasis is restored.
However, if this first line of defense is insufficient to remove nox-
ious compounds, the cause of inflammation persists and various im-
mune cells (e.g., macrophages, T-lymphocytes) are triggered to produce
cytokines and enzymes that cause more lasting damage. These pro-
cesses stimulate parenchymal cell death, resulting in loss of cell mem-
brane integrity and uncontrolled release of products of cell death and
pro-fibrogenic mediators into the extracellular space that stimulate
activity of pro-fibrogenic cell populations.
One prototypical factor in this scenario is transforming growth
Fig. 1. Common characteristics of organ fibrosis. Injurious events lead to factor-β (TGF-β). The three members of the TGF-β family (i.e. TGF-β1,
organ damage, inflammation, and fibrosis in liver, kidney, lung, heart and skin. TGF-β2, and TGF-β3) are active as secreted peptides and share similar
Hallmarks of fibrotic processes shared by all of these tissues include the ex- biological activities in vitro, while eliciting more specific biological re-
cessive production of cytokines, chemokines, growth factors, extracellular sponses in vivo (Gressner and Weiskirchen, 2006). During inflammation
matrix (ECM) proteins as well as loss of normal organ architecture and function. and fibrosis, TGF-β1 is the most important member in physiological
repair and collagen accumulation. On one side, it strongly promotes
mesenchymal transition (EMT) as well as other mesenchymal cells such collagen and fibronectin production in both epithelial and mesench-
as pericytes and resident mesenchymal stem/progenitor cells are po- ymal cells leading to the net accumulation of ECM tissue during wound
tential precursors for myofibroblasts (Di Carlo and Peduto, 2018). Be- repair and fibrosis. On the other site, the targeted disruption of the
side their influence on ECM production, these cells have also essential Tgfb1 gene in mice results in tissue necrosis in specific organs, over-
immunomodulatory and regenerative roles during wound healing whelming inflammatory cell infiltration into numerous organs and
(Weiskirchen and Tacke, 2014). Although reparative during normal systemic infiltration suggesting that TGF-β is a general suppressor of
wound healing, the resulting matrix involving the above mentioned excessive inflammation (Shull et al., 1992). Therefore, TGF-β is con-
cells and a plenitude of different soluble factors (chemokines, cyto- sidered as one of the common master switches for the induction of the
kines) loses its reparative capacity and can acquire a different archi- fibrotic program during chronic phases of inflammatory diseases in
tectural structure with elevation of characteristic matrix proteins (col- many organs and tissue. Other important soluble pro-fibrogenic med-
lagen, elastin), structural (basement) glycoproteins, proteoglycans and iators include connective tissue growth factor (CTGF) and members of
carbohydrates such as hyaluronan (Gressner and Weiskirchen, 2006). the platelet-derived growth factor (PDGF) family of growth factors.
This abundant fibrous stroma with increased ECM and growth factor While CTGF is considered to bind to TGF-β and enhance its binding to
production forms a microenvironment that allows the development of membrane-bound receptors (Abreu et al., 2002), members of the PDGF
tumor-associated desmoplasia facilitating tumor initiation. family are potent mitogens and chemoattractants for fibrogenic cells in
most organs driving the recruitment and proliferation of these cells at
sites of tissue injury (Kazlauskas, 2017; Borkham-Kamphorst and
2. Organ-independent mechanisms of fibrosis Weiskirchen, 2016). These common mediators promote activity and
increase total number of myofibroblasts, thereby up-regulating the rate
In the different organs, a wide range of noxious compounds and of matrix production and promoting a multitude of disease that affect a
triggers can lead to initiation and progression of fibrosis (Fig. 2). Tissue variety of organs including skin, lung, liver, kidney, pancreas, heart,
damage initiates an orchestrated repair process aiming to preserve and others.
organ function and tries to eliminate or insulate the initial cause of
injury or underlying harmful stimuli. This inflammatory response in-
volves the local vascular system, components of the immune systems,

2
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Fig. 2. Major causes of organ fibrosis. In the different organs, a wide range of noxious compounds and triggers can lead to initiation and progression of fibrosis. In
the affected organs, different changes and alterations are provoked that may lead to organ dysfunction or failure.

3. Cellular sources of extracellular matrix and fibrogenic signaling adhesive proteins, ground substance, and play fundamental roles in
ECM maintenance and reabsorption. Fibroblasts are dispersed
As outlined above, there is a large variety of cellular subsets and throughout the body and seem to be the least specialized cells in the
mechanisms contributing to the formation of extracellular matrix connective tissue that may differ across sites of the same organ. They
during fibrogenesis (Fig. 3). Potential cellular sources of ECM produ- are chemotactic and can migrate within tissue. This important feature
cing cells promoting fibrosis in liver, kidney, lung, heart and skin are permits to take over pivotal roles in wound healing, inflammation, and
manifold (Table 1). Most strikingly, the resident pro-fibrogenic cells fibrogenesis. Based on their contractile phenotype, they can contract
within the different tissues, their respective progenitors that invade the the matrix to seal an open wound. They further express proteins playing
inflamed tissue or cell fractions that become activated and obtain a a role in blood clotting, produce and are responsive to many in-
matrix-synthesizing phenotype are common in the different organs. flammatory cytokines, and play critical biological roles in angiogenesis.
However, in regards to their biological origin, profibrogenic features, During fibrogenesis, they (or subsets of these cells) can transform into
overall contribution to fibrogenesis, and relevance in different types of myofibroblasts that are hypersensitive to chemical signals such as cy-
fibrogenesis each cell type has its peculiarities. tokines, chemokines and growth factors, thereby provoking ex-
aggerated ECM production (Kendall and Feghali-Bostwick, 2014). In
most organs (e.g. skin, lung), resident fibroblast stromal cells that ex-
3.1. Fibroblasts pand and become activated appear to be the most significant source of
myofibroblasts, while in distinct organs other cell types (e.g. hepatic
Fibroblasts are derived from the embryonic mesoderm tissue re- stellate cells and portal myofibroblasts in liver) are the most significant
presenting the most common cell type of the connective tissue with the contributing source of myofibroblasts.
ability to form morphologically heterogeneous states with diverse ap-
pearances. Active fibroblasts have a branched cytoplasm with abundant
rough endoplasmic reticulum, while in an inactive state they are 3.2. Mesothelial cells
smaller, more spindle-shaped with lower content of rough endoplasmic
reticulum. They produce a large variety of ECM structural proteins, These cells are specialized cells forming monolayers (i.e. the

3
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Fig. 3. Origin and mechanisms leading to myofi-


broblast formation. During fibrogenesis, myofibro-
blasts are key players in extracellular matrix synth-
esis. This cell population can be derived from
resident fibroblast, mesothelial cells, circulating fi-
brocytes, epithelial cells, endothelial cell, pericytes,
vascular smooth muscle cells, Gli1+ perivascular
mesenchymal stem cell-like cells, and other specia-
lized cell types. Different mechanisms including cel-
lular activation, transformation, proliferation, in-
filtration, expansion, epthelial-to-mesenchymal
transition (EMT), mesothelial-to-mesechymal transi-
tion (MMT), and endothelial-to-mesenchymal-tran-
sition (EndoMT) are relevant in increasing the net
pool of myofibroblasts.

mesothelia) covering the pleural, peritoneal and pericardial serosal convey diverse functions in secretion, absorption, protection, and sen-
cavities. They derive from the embryonic mesoderm cell layer dis- sing. In general, epithelial cells are bound by a basal membrane and are
playing a phenotype resembling epithelial cells. Originally, it was closely connected to each other by specialized intercellular connections
thought that the mesothelium is mostly relevant to protect the internal (i.e. gap junctions, tight junctions, adherent junctions). Most strikingly,
organs of the body and to produce a lubricating fluid that is helpful in epithelial cells are most often multipotent having capacity to differ-
protecting the body against infection. There is evidence that these cells entiate into a variety of other cell types. In a process termed epithelial-
upon appropriate stimulus can be genetically reprogrammed. This to-mesenchymal transition (EMT), epithelial cells can undergo multiple
transition process, termed “mesothelial-to-mesenchymal transition” biochemical changes in which they lose their polarity and acquire a
(MMT) is associated with morphological and functional changes and mesenchymal phenotype having enhanced migratory capacity, inva-
lead to cells producing extracellular matrix compounds. In vitro, TGF-β1 siveness, and massively increased production of ECM components
was shown to induce this reformation in differentiated human me- (Kalluri and Weinberg, 2009). As a further characteristic, these trans-
sothelial cells and become positive for α-SMA, highlighting the fact that ited cells become positive for the fibroblast-specific protein (FSP1, also
these cells can act as progentiors for myofibroblasts (Yang et al., 2003). known as S100A4). During the last decade, EMT and the cellular
Using mouse models and conditional cell lineages analysis, it was de- plasticity of epithelial cells allowing them to acquire features of me-
monstrated that MMT give rise to hepatic stellate cells and myofibro- senchymal cells was claimed to be another significant source of myo-
blasts during liver fibrogenesis and that this process can be effectively fibroblasts in injured tissues. Although this concept is still valid in many
blocked in vitro and in vivo by antagonising TGF-β (Li et al., 2013). organs, cell tracing experiments in other organs have questioned the
role of EMT in organ fibrosis (Österreicher et al., 2011). In particular,
3.3. Fibrocytes there are many controversial reports on EMT in hepatic fibrogenesis.
While a large number of independent studies confirm that hepatocytes
This spindle-like shaped cell type is of mesenchymal origin and or biliary epithelial cells (cholangiocytes) undergo EMT upon TGF-β
possesses an inactive phenotype with limited amount of rough en- treatment in culture, in vivo experimentation which are based on ge-
doplasmic reticulum. It was first described in 1994 as a new leukocyte netic fate mapping analysis suggested that EMT is not relevant for he-
subpopulation that appeared concurrently with circulating in- patic fibrogenesis in vivo (Munker et al., 2017).
flammatory cells accounting for approximately 10% of the cells that
infiltrate subcutaneously implanted wound chambers and found in re- 3.5. Endothelial cells
gions of scar formation (Bucala et al., 1994). Typically, these cells ex-
press fibroblasts components including vimentin, collagen, fibronectin This cell population forms a one-cell thick walled layer (i.e. the
and display the leukocyte common antigen CD45 (Bucala et al., 1994). endothelium) that lines the entire interior surface of all blood vessels
Fibrocytes can be induced by TGF-β to express α-SMA, have funda- such as arteries, arterioles, venules, veins and capillaries. These cells
mental roles in adaptive immunity and further have capacity to support contribute to the maintenance of barrier functions, blood flow and are
angiogenesis (Quan et al., 2004). A large variety of studies using bone active cellular participants in inflammatory responses. In 1992, it was
marrow (BM) transplantation in mice have demonstrated that BM is the demonstrated that the bovine aortic endothelial cells express α-SMA
source of circulating fibrocytes that transmigrate with the blood stream when exposed to TGF-β1 suggesting that endothelial cells have capacity
specifically into the side of injury. to transit into a smooth-muscle-like phenotype (Arciniegas et al., 1992).
This phenotypic transition of endothelial cells into mesenchymal cells
3.4. Epithelial cells was later termed “endothelial-to-mesenchymal-transition" (EndoMT).
There are evidence that the process of EndoMT can contribute to car-
This group of cells are located in surfaces of the body such as skin, diac fibrosis, kidney fibrosis, and pulmonary fibrosis (Piera-Velazquez
blood vessels, urinary tract, or organs. Their cell polarity allows them to et al., 2016).

4
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Table 1 In the liver, the sinusoidal pericytes are commonly named hepatic
Potential cellular sources of extracellular matrix producing cells in fibrosis of stellate cells (HSC). These cells are located in the space between par-
liver, kidney, lung, heart and skin. enchymal cells and sinusoidal endothelial cells (i.e. the space of Disse)
Organ Cell type References of the hepatic sinusoid and have several specialized functions. In par-
ticular, these cells are the major vitamin A storage site of the body and
Liver Hepatic stellate cells (lipocytes) Friedman et al., 1985 have hemodynamic and immunoregulatory functions (Tsuchida and
Portal myofibroblasts Tang et al., 1994;
Friedman, 2017). During fibrogenesis, HSC transdifferentiate into
Tuchweber et al., 1996
Resident fibroblast Guyot et al., 2006 myofibroblasts giving rise to 82–96% of the hepatic myofibroblast pool
Vascular smooth muscle cells Guyot et al., 2006 in experimental models of toxic, cholestatic, and fatty liver disease, as
BM-derived myofibroblast precursors Forbes et al., 2004 estimated in fate tracing experiments using a construct in which Cre
(fibrocytes, circulating mesenchymal cells)
expression was driven by a lecithin-retinol acyltransferase (Lrat) gene
Hepatocytes undergoing EMT Zeisberg et al., 2007
Biliary epithelial cells (cholangiocytes) Rygiel et al., 2008
promoter (Mederacke et al., 2013). Interestingly, activation of HSC into
undergoing EMT matrix-producing myofibroblasts in the liver is not unidirectional; ex-
Sinusoidal endothelial cells Bardadin and Desmet, perimental evidence unravelled the deactivation of myofibroblasts into
1985 a “quiescent-like" HSC phenotype during the resolution from injury
Mesothelial cells (MMT or EndoMT) Li et al., 2013
(Troeger et al., 2012; Kisseleva et al., 2012).
Gli1+ perivascular MSC Kramann et al., 2015
Kidney (Tubular) epithelial cells Grgic et al., 2012
Endothelial cells/monocytes Phua et al., 2013 3.7. Vascular smooth muscle cells
Circulating fibrocytes and mesenchymal LeBleu et al., 2013
stem cells of BM origin
These contractile cells are a specialized smooth muscle cell type
Resident fibroblasts LeBleu et al., 2013
Pericytes Lin et al., 2008;
within the normal blood vessel wall meditating contraction and re-
Humphreys et al., 2010 laxation. They typically express α-SMA, transgelin (SM22α), vimentin
Perivasuclar fibroblasts Lin et al., 2008 and desmin. In response to injury, they become activated, proliferate,
Podocytes undergoing EMT Ying and Wu, 2017 migrate and express different extracellular matrix proteins.
Gli1+ perivascular MSC Kramann et al., 2015
Mechanistically, it was demonstrated that bradykinin induces collagen
Lung Resident fibroblasts Low et al., 1984
BM-derived mesenchymal progenitor cells Hashimoto et al., 2004; type I production in vascular smooth muscle cells via autocrine acti-
(fibrocytes) Xu et al., 2009 vation of TGF-β1 and implication of mitogen-activated protein kinase
Pericytes Hung et al., 2013; Rock pathways (Douillet et al., 2000).
et al., 2011
Alveolar epithelial cells undergoing EMT Willis et al., 2005
Endothelial cells (EndoMT) Hashimoto et al., 2010
3.8. Gli1 positive perivascular mesenchymal stem-like cells
Gli1+ perivascular MSC Kramann et al., 2015
Heart Cardiac fibroblasts Kawaguchi et al., 2011 Mesenchymal stem cells (MSC) are multipotent connective tissue
Cardiomyocytes Kawaguchi et al., 2011 cells that reside in the perivascular niche of many organs. Genetic
BM-derived circulating fibrocytes Keeley et al., 2012
lineage tracing analysis demonstrated that tissue-resident, but not cir-
Endothial and vascular smooth muscle cells Douillet et al., 2000; Wu
et al., 2016 culating, glioma-associated oncogene homolog Gli1 positive (Gli1+)
Gli1+ perivascular MSC-like cells Kramann et al., 2015 cells that express typical MSC markers, expand and become myofibro-
Skin Resident fibroblasts (of different lineages) Rinkevich et al., 2015 blasts, thereby contributing to the formation of renal, hepatic, pul-
Pericytes Sundberg et al., 1996; Liu monary and cardiac fibrosis (Kramann et al., 2015). In line with this
et al., 2010
Fibrocytes Iqbal et al., 2012
assumption, the ablation of Gli1+ cells with an inducible diphtheria
Stromal cells Dulauroy et al., 2012 toxin receptor allele resulted in significant amelioration of experi-
mental kidney and heart fibrosis (Kramann et al., 2015).
Abbreviations usedare: BM, bone marrow; EndoMT, Endothelial-mesenchymal
transition; EMT, epithelial-to-mesenchymal transition; Gli1, glioma-associated 3.9. Specialized cells
oncogene family member 1; MSC, mesenchymal stem cell; MMT, mesothelial-
to-mesenchymal transition.
In some organs, specialized resident cells are present that contribute
to the excess of matrix during fibrogenesis. These have several features
3.6. Pericytes in common with the cells discussed above, but are located in defined
areas of a tissue or have somewhat different expression profiles than
Pericytes are multipotent, mesenchym-derived, and heterogeneous other myofibroblasts. For example, beside the above mentioned hepatic
cells with a variety of origins, functions, morphologies and surface
stellate cells, the liver contains so called portal myofibroblasts. These
markers. They contain contractile elements and are wrapped around the perivascular mesenchymal cells have been defined as liver myofibro-
endothelial cells that line the capillaries and venules or are embedded
blasts derived from cells that are distinct from HSC and located in the
in basement membranes in proximity to endothelial cells. They are portal tract (Lemoinne et al., 2013). It is well accepted that these cells
involved in regulation of capillary blood flow, phagocytic clearance of
become activated and obtain a myofibroblastic phenotype in cholestatic
cell debris, and support endothelial cells in forming the blood-brain liver fibrosis and can be distinguished from HSC on the basis of mor-
barrier. Importantly, although pericytes endogenously express α-SMA, phological criteria and growth behaviour (El Mourabit et al., 2016). In
they are not terminally differentiated and can transit into fibroblasts, the kidney, the podocytes found in the lining of the Bowman's capsules
osteoblasts, and smooth muscle cells (Kendall and Feghali-Bostwick, in the nephrons possess a well-developed endoplasmic reticulum and a
2014). The precise contribution of this cell population to the formation large Golgi apparatus. They are derived from the mesenchyme of the
of fibrogenesis in different organs is still controversially discussed. mesoderm and can undergo EMT in pathological states. During podo-
Particularly, conflicting data were reported for pericytes in the patho-
cyte transition, the expression of nephrin, podcin, P-cadherin and ZO-1
genesis of lung fibrosis. While one report excluded pericytes that ex- is downregulated, while the expression of α-SMA and N-cadherin are
press α-SMA, PDGFRβ and NG2 as an origin of myofibroblasts in the
increased in a process that is majorly dependent on TGF-β (Ying and
lung, another report suggested that Foxd1 progenitor-derived pericytes Wu, 2017). Similarly, a historical landmark paper that has provided
significantly contribute to the formation of lung fibrosis (Rock et al.,
first evidence for EMT in kidney demonstrated that tubular epithelial
2011; Hung et al., 2013). cells might also acquire migratory properties and transit into a

5
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

myofibroblast-like phenotype characterised by de novo expression of 4.2. Cytokines


collagen, vimentin and FSP1 (Strutz et al., 1995).
Cytokines are small proteins produced by a broad range of cells that
can bind on specific cell-surface exposed receptors. They can act in an
4. Key mediators of inflammation and fibrogenesis
autocrine (i.e. binding to receptors on to the cell that produces it) or
paracrine (i.e. released from one cell and binding to receptors on a
There are numerous pro-inflammatory and pro-fibrogenic mediators
nearby cell) manner. After binding, cytokines initiate specific in-
that can act in a hierarchical manner. During the last decades, the
tracellular signaling pathways impacting protein activity and mod-
emerging role of chemokine production as a critical factor regulating
ulating downstream gene expression. Some of these cytokines that are
immune cell entry into the inflamed tissue at the onset and during
critically involved in the pathogenesis of fibrosis are briefly discussed in
progression became evident. Similarly, several cytokines (TGF-β, PDGF,
the following.
IL-1β, IL-6, IL-13, IL-33, and TNF-α) hold important functions in the
inflammatory process and in the progression to fibrosis. Work from
4.2.1. TGF-β
recent years have shown that also some non-peptidic mediators (e.g.
The three different TGF-βs belong to a large family of at least 35
reactive oxygen species, ROS; lipid mediators, acetaldehyde) induce
structurally related members forming the TGF-β superfamily. The in-
expression or activity of mediators relevant in fibrogenesis. In the fol-
dividual members of this family are encoded by genes dispersed
lowing, we will briefly highlight the biological activities and impact of
throughout the genome (Weiskirchen and Meurer, 2013). Typically,
these molecules on inflammation and fibrogenesis.
they form similar, symmetric homodimers secreted as biologically in-
active mediators that are transformed into their biologically active form
4.1. Chemokines by proteolytic processing. Once activated, they bind to two kinds of
membrane receptors with serine/threonine kinase activity, which
As stated above, the concerted recruitment of immune cells from the mediate their signals through Smad-dependent and Smad-independent
circulation into persistently injured tissue is a key mechanism during pathways (Weiskirchen and Meurer, 2013). There is strong evidence
fibrogenesis in liver (Heymann and Tacke, 2016), kidney (Tecklenborg that the blockade of TGF-β alone is sufficient to completely block ex-
et al., 2018), lung (Kolahian et al., 2016), heart (Frantz et al., 2018) or perimental fibrogenesis in liver, kidney, lung, heart and skin (Gressner
other tissues affected by fibrosis. It has long been appreciated that et al., 2002; Weiskirchen, 2016; Natase et al., 2017; Aschner and
chemokines, chemoattractant cytokines that bind to specific chemokine Downey, 2016; Thanigaimani et al., 2017; Andrews et al., 2016). In all
receptors on leukocytes, orchestrate the recruitment and partially also organs, TGF-β is a key cytokine driving collagen gene expression. This
the activation of immune cells in inflammation (Charo and Ransohoff, was most obviously demonstrated in experimental liver fibrosis. The
2006). Altogether, the chemokine system includes about 50 different suppression of TGF-β synthesis (Arias et al., 2003) or the expression of a
chemokine ligands and 20 cognate chemokine receptors, of which many soluble TGF-β receptor sequestering biological active TGF-β resulted in
have been linked to the progression or regression of organ fibrosis attenuation of experimentally induced hepatic fibrosis. Similar results
(Zlontnik and Yoshie, 2000). The exact functions of chemokines in were obtained when BMP-7 acting as a physiological opposing factor of
organ injury and fibrosis have been reviewed elsewhere in detail for the TGF-β was overexpressed in a model of hepatic fibrogenesis (Kinoshita
liver (Marra and Tacke, 2014), the kidney (Kurts et al., 2013), the lung et al., 2007). Most importantly, also antioxidants that interfere with
(Tomankova et al., 2015) and the heart (Jones et al., 2017; Swirski and distinct molecular steps in TGF-β singaling were shown to mediate
Nahrendorf, 2013). beneficial effects on experimental liver fibrosis (Meurer et al., 2005;
One prime example for an organ-independent chemokine signal in Weiskirchen, 2016). As discussed above, TGF-β not only induces ex-
fibrosis is the C-C motif chemokine 2 (CCL2) or monocyte chemoat- tracellular matrix formation in pro-fibrogenic cells, it also increases the
tractant protein-1 (MCP-1), which is released by parenchymal cells, pool of myofibroblasts by promoting or triggering the processes of EMT,
activated tissue macrophages and fibroblasts in organ injury. In human MMT, and EndoMT.
disease as well as in experimental animal models of fibrosis, CCL2
provokes the accumulation of inflammatory, monocyte-derived mac- 4.2.2. PDGF
rophages in injured tissue, which provoke activation of fibroblasts. This The PDGF family is comprised of five members (PDGF-AA, PDGF-
principle has been recapitulated in liver (Karlmark et al., 2009), kidney AB, PDGF-BB, PDGF-CC, and PDGF-DD). They mediate their biological
(Kitagawa et al., 2004), cardiac (Frangogiannis et al., 2007) and pul- effects via the receptor tyrosine kinases PDGFRα and PDGFRβ that can
monary fibrosis (Okuma et al., 2004). Moreover, in several animal form pure or mixed dimers. Typically, PDGFs are potent mitogens
models of fibrosis, the CCR2-CCL2 axis was an excellent target for driving cell proliferation and differentiation. With regards to in-
pharmacological interventions (Tacke, 2017; Baeck et al., 2012; flammatory liver insult, the expression of the different PDGF ligand and
Krenkel and Tacke, 2017; Ninichuk et al., 2008). In an exploratory early receptors is differentially regulated (Borkham-Kamphorst et al., 2008;
phase clinical trial, inhibition of CCL2 by the pegylated aptamer NOX- Martin et al., 2013). In diverse models of hepatic fibrogenesis is was
E36 improved disease biomarkers in patients with diabetic nephropathy suggested that the targeting of PDGF-B, PDGF-D, and PDGFRβ should
(Menne et al., 2017). In a multicenter, randomized, placebo-controlled be therapeutically more effective than targeting ligands that act via the
phase 2 clinical trial involving 289 patients with non-alcoholic steato- PDGFRα receptor branch that seem to be more relevant in the patho-
hepatitis and liver fibrosis, treatment with the dual CCR2/CCR5 in- genesis of tumor formation (Weiskirchen, 2016). Similarly, PDGF-DD
hibitor cenicriviroc demonstrated a higher rate of patients with histo- and PDGFRβ were markedly upregulated in both human and murine
logical improvement of hepatic fibrosis after 1 year of therapy kidneys on activated mesenchymal cells demonstrating their im-
(Friedman et al., 2018). This led to the implementation of a large phase portance for kidney disease (Buhl et al., 2016). Besides these simila-
3 clinical trial on this substance with the primary end-point to treat rities there are also marked differences between hepatic and renal fi-
liver fibrosis (Tacke, 2018). As there is a high redundancy within the brosis. While PDGF-C deficiency or antagonism failed to protect from
chemokine system as well as the involvement of multiple chemokines liver fibrosis, the lack of PDGF-C prevented kidney fibrogenesis in the
and chemokine receptors during fibrogenesis, it can be anticipated that unilateral ureteral obstruction model (Martin et al., 2013). With regards
targeting other (or even multiple) chemokine pathways will offer to heart, lung, and skin, there is also clear evidence that PDGF sig-
therapeutic potential for organ fibrosis as well (Lee et al., 2015; nificantly contributes to the formation of cardiac and pulmonary fi-
Cavalera and Frangogiannis., 2014; Tomankova et al., 2015; Marra and brosis by stimulating the proliferation and migration of collagen-se-
Tacke, 2014). creting fibroblasts as well as activating mesenchymal stem and/or

6
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

progenitor cells contributing to the formation of novel myofibroblasts prototypes of these non-peptidic mediators will be briefly discussed.
during disease progression (Kong et al., 2014; Nishioka et al., 2013;
Iwayama and Olson, 2013). 4.3.1. ROS
Reactive oxygen species (ROS) formation is a critical driver of in-
4.2.3. Interleukins flammation and fibrosis. Under healthy conditions, NADPH oxidase
IL-1β, IL-6, IL-13, and IL-33 belong to the interleukin family of (NOX)-derived ROS are essential modulators of signal transductions
cytokines that contain over 40 different members. Most interleukins pathways controlling cell growth, proliferation, migration, differentia-
possess pleiotropic activities in the innate and adaptive immune re- tion and apoptosis (Manea et al., 2015). However, uncontrolled ROS
sponse and in nearly all types of inflammatory responses. IL-1β and IL-6 generation and oxidative stress induce cell damage. In liver damage,
are critical proinflammatory contributors to the pathogenesis of hepatic elevated ROS concentrations play an important role in promoting liver
inflammation, steatosis, and fibrosis (Del Campo et al., 2018). How- damage and initiating fibrosis by disrupting integrity of lipids, proteins,
ever, IL-6 can also exert tissue-protective and pro-regenerative func- DNA, inducing parenchymal cell necrosis and apoptosis, and stimu-
tions in the liver, thereby attenuating liver fibrosis (Kovalovich et al., lating the production and release of profibrogenic mediators from
2000; Hammerich and Tacke, 2014). IL-6 is also fundamentally in- Kupffer cells and circulating inflammatory cells that subsequently ac-
volved in renal fibrosis. It is produced by renal resident cells, including tivate hepatic stellate cells (Sánchez-Valle et al., 2012).
podocytes, mesengial cells, endothelial cells and tubular epithelial cells Mechanistically, it is also now well accepted that ROS and TGF-β
that are highly responsive towards IL-6 via classic signaling or via trans- are interlinked (Richter und Kietzmann, 2016). On one side, ROS for-
signaling pathways that are induced by binding of a ligand-bound so- mation has been shown to be enhanced in response to TGF-β1, most
luble IL-6 receptor to gp130 on cells that do not express endogenous IL- likely by suppressed expression of several antioxidant enzymes. On the
6 receptors (Su et., 2017). other hand, increased production of ROS leads to proteolytic activation
Recently, it was demonstrated that IL-33 can promote the process of and elevated expression of TGF-β1 in different organs (De Bleser et al.,
pulmonary fibrosis by inducing the imbalance between MMP-9 and 1999; Barnes and Gorin, 2011). During the last year it was recognized
TIMP-1 (Wu et al., 2018). On the other side, IL-33 signaling has been that one of the key enzymes linking ROS to TGF-β activity are the
found to have protective effects in cardiovascular disease and to at- different NOXs. Different NOX subtypes critically contribute to pro-
tenuate cardiac fibrosis by modulating fibroblast function and gene duction of intracellular ROS (Manea et al., 2015). In line with this as-
expression (Zhu and Carver, 2012). However, it this study it was de- sumption, NOX1-or NOX-4 deficient mice are protected against liver
monstrated that the major effects of IL-33 were induced by modulation inflammation and fibrosis (Lan et al., 2015). Similar results were found
of cell migration and activation of cytokine/chemokine expression, in the kidney, in which the depletion of the regulatory p47phox subunit
rather than modulating expression of extracellular matrix components of NOX1 and NOX2 required for NOX-mediated ROS production re-
or impacting proliferation. Moreover, in liver fibrosis, IL-33 was de- sulted in amelioration of renal fibrosis following injury (Wang et al.,
scribed to activate fibrosis-promoting innate lymphoid cells type 2 as 2015). In human gingival and dermal fibroblasts, the inhibition of
well as collagen-producing stellate cells (McHedlidze et al., 2013; NOX1/4 by selective inhibitors resulted in an impairment of the ability
Weiskirchen and Tacke, 2017). These examples show that the effect of of TGF-β to induce profibrotic gene expression, further underpinning
individual interleukins might vary between organs and in different the notion that ROS and TGF-β are connected in the pathogenesis of
disease settings. In addition, there is a wealth of information showing fibrosis (Murphy-Marshman et al., 2017).
that individual members of the interleukin family have overlapping but
also distinct biological activities and may evolve pro- or anti-in- 4.3.2. Lipid mediators and extracellular vesicles
flammatory activities (as well as both activities). Therefore, it is most Another mechanism contributing to the production of enhanced
likely that strategies aiming to antagonize a specific interleukin could TGF-β within a tissue was recently identified in lung fibrosis (Romero
have serious side effects and might be beneficial only in selected disease et al., 2015). In the respective study, it was demonstrated that lipid-
conditions. A good example for a potential therapeutic interleukin laden macrophages (i.e. foam cells) that are regularly observed in fi-
target is Schistosomiasis, in which IL-13 appears to be the dominant brotic lung tissue release oxidized phosphatidylcholine, which induces
pro-fibrotic cytokine associated with liver fibrogenesis by directly in- the production of TGF-β1 and promotes M2 polarization of macro-
ducing expression of fibrosis-associated genes including collagens and phages. This finding provides first evidence of a paracrine lipid sig-
CTGF in profibrogenic hepatic stellate cells (Liu et al., 2012). naling axis linking epithelial injury to macrophage activation, accu-
mulation of oxidized lipids, expression of TGF-β, and initiation of
4.2.4. TNF-α fibrotic cascades. Moreover, stressed parenchymal cells can release
Tumor necrosis factor-α (TNF-α) belongs to the TNF family con- extracellular vesicles, termed exosomes, that can transmit stress signals
sisting of 19 different proteins. It is mainly secreted by macrophage to neighbouring immune cells, mainly macrophages; this mechanism
subsets and binds to two different receptor types (TNF-R1 and TNF-R2) has been well described in many fibrotic disease settings including
that initiate a plenitude of signaling cascades leading to activation of heart, kidney, lung and liver (Hirsova and Gores, 2015; Zhang et al.,
NF-κB and triggering cellular activation, differentiation, cytokine pro- 2016; Szabo and Momen-Heravi, 2017; Vanhaverbeke et al., 2017;
duction and cell apoptosis (Weiskirchen, 2016). As such it is an im- Kubo, 2018).
portant inflammatory signaling molecule during fibrogenesis of liver,
kidney, heart, lung, and skin (Borkham-Kamphorst et al., 2013; Del 4.3.3. Acetaldehyde
Campo et al., 2018; Lv et al., 2018; Sun et al., 2007; Murdaca et al., In the liver, acetaldehyde induces expression of both type I collagen
2016). genes via acetaldehyde-responsive elements binding different tran-
scriptions factors and by a mechanism dependent on the generation of
4.3. Non-peptide mediators H2O2 (Greenwel et al., 2000; Svegliati-Baroni et al., 2005). Although
acetaldehyde and TGF-β1 both stimulate the phosphorylation and ex-
Excessive ROS generation, accelerated lipid accumulation, and for- pression of the intracellular mediators Smad3 and Smad4 and their
mation of acetaldehyde were identified as crucial factors in the pa- binding to the collagen gene promoter, only TGF-β can enhance phos-
thogenesis of fibrosis. Some of their biological activities are directly phorylation of Smad2 in hepatic stellate cells (Reyes-Gordillo et al.,
associated with the induction of TGF-β gene expression or activity, 2014). In respective cells, neither acetaldehyde nor TGF-β1 alone had
while others seemed to be mediated in a TGF-β-independent manner. significant stimulatory effect on a collagen reporter construct, while
The biological significance of ROS, lipids, and acetaldehyde as cells treated with the combination of both showed a highly significant

7
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

increase in reporter activity (Reyes-Gordillo et al., 2014). This finding inflammatory microenvironment: this implies major phenotypic ad-
suggests that the fibrogenic signaling pathway triggered by acet- justments of the immune cells, especially the induction of a “re-
aldehyde and TGF-β are additive and distinctly different. Since myo- storative phenotype” in macrophages and modulation of type 2
cardial fibrosis is regularly seen in patients with chronic alcohol in- immunity (Krenkel et al., 2018; Pakshir and Hinz, 2018; Gieseck
gestion (Ren and Wold, 2008), it might be possible that the activity of et al., 2018)
acetaldehyde in promoting fibrogenesis is not restricted to the liver. (3) deactivation and elimination of myofibroblasts: activated fibrogenic
myofibroblasts can undergo apoptosis, become senescent (irrever-
5. Mechanisms of fibrosis regression and resolution sible cell cycle arrest) or revert to an inactive/quiescent phenotype
(Jun and Lau et al., 2018)
Organ fibrosis with an excessive accumulation of extracellular ma- (4) degradation of extracellular matrix: this includes the activation of
trix in response to chronic tissue injury may disintegrate the regular matrix metalloproteinases (MMPs) that digest collagen and other
architecture of the organ and subsequently promote organ failure. This ECM proteins, the contribution of macrophages that phagocytize
is well known in chronic liver diseases, where cirrhosis is the end-stage matrix fragments as well as the reduction of MMP-inhibitory pro-
with impaired liver function and portal hypertension, chronic kidney teins (e.g., tissue inhibitors of MMPs) (Karsdal et al., 2017; Jun and
diseases with a destroyed glomerular filtration function in case of se- Lau et al., 2018).
vere glomerulosclerosis as well as in pulmonary diseases, in which fi-
brosis reduces the vital capacity of the lung (Rosenbloom et al., 2017). 6. Diagnosis and staging of fibrosis
While fibrosis development has long been viewed as a unidirectional
process consisting of non-reversible sequela from chronic injury to fi- The presence and extent of organ fibrosis determines disease pro-
brosis to tissue destruction, there is clear evidence that fibrogenesis is to gression and outcome, as evidenced for chronic kidney diseases
a large extent reversible, even at later stages (Jun and Lau et al., 2018). (Berchtold et al., 2017), non-alcoholic fatty liver (Dulai et al., 2017)
Already two decades ago, Fioretto and colleagues reported that patients and pulmonary fibrosis (Kaur et al., 2017). Thus, it is important to di-
with type 1 diabetes (n = 8), who underwent pancreas transplantation, agnose fibrosis – ideally at early stages – and to accurately assess the
reverted fibrotic lesions in their kidneys by histological analysis five stage of fibrosis. The gold standard for diagnosis and staging of fibrosis
and ten years after pancreas transplantation, demonstrating that (early) is histology, which is usually performed on fine-needle biopsies or
kidney fibrosis is in principle reversible (Fioretto et al., 1998). Around surgically obtained samples. Different histological scores exist for the
the same time, Hammel and colleagues noted that patients with severe various organs that usually also take into account the different pattern
liver fibrosis due to chronic obstruction of the common bile duct of fibrosis dependent on the disease etiology. Nonetheless, histological
(n = 11) regressed with their hepatic fibrosis by histological analysis assessment of fibrosis has serious limitations: obtaining histology, such
2.5 years after biliary drainage (Hammel et al., 2001). In a prospective as by percutaneous biopsy of the liver or the kidney, has procedure-
clinical trial on the use of tenofovir for chronic hepatitis B virus in- associated risks including bleeding (Kumar et al., 2018); the biopsy
fection, 75% of patients with cirrhosis at baseline were able to regress represents only a fraction of the organ that might not be representative
to a non-cirrhotic stage in a follow-up liver biopsy five years after in- for the distribution of fibrosis (Patel et al., 2015); and the interpretation
itiation of an effective antiviral therapy (Marcellin et al., 2013). Given of the results shows a remarkable intra- and interobserver variability
the exceptional clinical relevance of fibrosis for predicting morbidity regarding the exact fibrosis stage (Mäkelä et al., 2018; Manning and
and mortality in patients with liver diseases (Dulai et al., 2017), trials in Afdahl, 2008). Therefore, many additional methods have been pro-
the area of non-alcoholic steatohepatitis commonly use “histological posed and partially already implemented in clinical algorithms to non-
improvement of fibrosis” as an end-point, and some of the compounds invasively assess the stage of fibrosis. These techniques include imaging
under investigation yielded positive signals in early clinical trials methods, mechanical and functional tests, serum biomarkers and ge-
(Loomba et al., 2018; Friedman et al., 2018; Neuschwander-Tetri et al., netic risk factors for fibrosis (Table 2).
2015). For cardiac or pulmonary fibrosis, there are less histological data Imaging techniques are widely used in patients with chronic dis-
available from human patients that would support structural fibrosis eases, in which fibrosis is suspected. However, ultrasound, for instance,
regression in these organs. Nonetheless, functional recovery of the can only show indirect signs of fibrosis, such as small organ size with
organ, e.g., improved left ventricular ejection function or improved impaired parenchyma in fibrotic kidneys or coarsened echotexture,
forced vital capacity in idiopathic pulmonary fibrosis, has been clearly blunting of liver edges and signs of portal hypertension in fibrotic/
documented with presumably anti-fibrotic interventions (Murtha et al., cirrhotic livers (Berchtold et al., 2017; Baues et al., 2017). Magnetic
2017; Karimi-Shah and Chowdhury, 2015). resonance imaging (MRI) might be more suitable in these organs, as T1/
The clinical findings on reversibility of fibrosis prompted extensive T2 mapping, specific contrast agents and diffusion-weighted imaging
studies using animal models to understand the mechanisms of fibrosis emerge as tools for fibrosis evaluation (Berchtold et al., 2017; Baues
regression and fibrosis resolution. In principle, fibrosis regression could et al., 2017). In the liver, the MRI-derived proton density fat fraction
be recapitulated in experimental fibrosis models of the liver (Duffield (MRI-PDFF) is already broadly applied to quantify hepatic fat content,
et al., 2005), kidney (Aldigier et al., 2005), lung (Reddy et al., 2014) which has been used as an endpoint in early clinical trials with anti-
and heart (Jeong et al., 2016). In case that the underlying chronic in- fibrotic agents in non-alcoholic steatohepatitis (Caussy et al., 2018).
jury can be removed, the degree of full recovery seems to depend on the Cardiac MRI is firmly implemented in clinical algorithms, because it
capacity of the tissue to regenerate, likely explaining a higher chance allows the assessment of cardiac wall thickness, myocardial mass, late
for fibrosis resolution in the liver as compared to lung, kidney or heart gadolinium enhancement and myocardial extracellular volume frac-
(Cordero-Espinoza and Huch et al., 2018; Murtha et al., 2017). tion, which is related to myocardial fibrosis (Amano et al., 2018). Re-
From these various experimental animal models of organ fibrosis garding pulmonary fibrosis, high-resolution chest computed tomo-
and recovery, general principles of fibrosis regression have been iden- graphy (CT) provides diagnostic and prognostic information in patients
tified (Tacke and Trautwein, 2015): with lung fibrosis, and different radiological parameters are established
to stage disease severity so that lung biopsy is only seldom needed
(1) cessation of chronic tissue injury: the removal or termination of the (Ohkubo et al., 2018; Martinez et al., 2017). The particular advantages
underlying cause of tissue damage promotes regenerative pathways and disadvantages of imaging modalities for organ fibrosis are reviewed
in parenchymal cells and avoids further activation of myofibro- elsewhere (Baues et al., 2017).
blasts (Cordero-Espinoza and Huch et al., 2018) Progressive fibrosis changes the “mechanical properties” of an
(2) deactivation of inflammatory pathways and establishment of an anti- organ. This has been successfully translated into elastography

8
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Table 2
Current methods for diagnosis and staging of fibrosis in selected organs.
Liver Kidney Lung Heart

Gold standard histology (biopsy)

Imaging ultrasound, CT scan, MRI ultrasound, MRI high-resolution chest CT echocardiography, cardiac MRI (e.g.,
extracellular volume)
Mechanical or elastography elastography (?) body plethysmography echocardiography, cardiac MRI
functional test
Serum biomarkers Composite scores (e.g., FIB-4, ELF, Not validated yet; TGF-β, Not validated yet; IL-8, CCL18, MMP- Not validated yet; NT-proBNP (?)
(examples) Fibrotest, NFS), hyaluronic acid, MCP-1, MMP-2, urinary 1, MMP-7, IGBPs
PIIINP, pro-C3 PIIINP
Genetic markers PNPLA3, TM6SF2, MBOAT7, GCKR, ACE, TGFβ1, PARN (?) MUC5B, TOLLIP, TERT, TERC, MYH7, MYBPC3
(examples) A1AT, HFE FAM13A, DSP, OBFC1, ATP11A, DPP9

Abbreviations usedare: A1AT, α1-Antitrypsin; ACE, angiotensin I-converting enzyme; ATP11A, ATPase phospholipid transporting 11A; CCL18, (C-C motif) ligand 18;
CT, computed tomography; DPP9, dipeptidyl peptidase IX; DSP, desmoplakin; ELF, enhanced liver fibrosis; FAM13A, family with sequence similarity 13 member A;
FIB-4, fibrosis-4; GCKR, glucokinase regulatory protein; HFE, hereditary hemochromatosis; IGBPs, immunoglobulin G binding proteins; IL-8, interleukin 8; MBOAT7,
membrane bound O-acyltransferase domain containing 7; MCP-1, monocyte chemoattractant protein-1; MMP-1/2/7, matrix metalloproteinase-1/2/7; MRI, magnetic
resonance imaging; MUC5B, mucin 5 subtype B; MYBPC3, myosin-binding protein C cardiac; MYH7, myosin heavy chain 7; NFS, nonalcoholic fatty liver disease
(NAFLD) fibrosis score; NT-proBNP, N-terminal pro-brain natriuretic peptide; OBFC1, oligonucleotide/oligosaccharide binding fold-containing protein 1; PIIINP, N-
terminal propeptide of type III collagen; PARN, polyadenylate-specific ribonuclease; PNPLA3, Patatin-like phospholipase domain-containing protein 3; pro-C3, N-
terminal type III collagen propeptide; TERC, telomerase RNA component; TERT, telomerase reverse transcriptase; TGF-β, transforming growth factor-β; TM6SF2,
transmembrane 6 superfamily 2; TOLLIP, toll interacting protein.

techniques for assessing fibrosis of the liver. These methods comprise ELF or FibroTest, and they overall provide valid diagnostic and prog-
ultrasound-based transient elastography (TE, also termed fibroscan), nostic information in hepatic fibrosis (European Association for Study
acoustic radiation force impulse imaging (ARFI) and shear wave elas- of Liver et al., 2015; Boursier et al., 2016).
tography (SWE) as well as MRI-based elastography (MRE). The under- As many gene loci have been identified as risk factors for fibrosis
lying principle is that the propagation of a shear wave or ultrasound progression by genome-wide association studies, it can be expected that
pulse is followed by ultrasound or MRI and directly reflects the liver the analysis of gene polymorphisms might be useful in the diagnosis of
stiffness as a surrogate for liver fibrosis (Friedrich-Rust et al., 2016). fibrosis and/or risk stratification. For instance, a variant (rs35705950)
This methodology has been widely evaluated in patients with different in the MUC5B gene promoter region, alongside other genetic risk loci,
chronic liver diseases, so that newer guidelines recommend in- has a strong association with idiopathic pulmonary fibrosis (IPF) and
corporated this technique in the diagnosis of hepatic fibrosis (European familial interstitial pneumonia across multiple different cohorts
Association for Study of Liver, 2015). Similar ultrasound and MR (Mathai et al., 2016). In cardiology, many genetic risk factors have been
elastography approaches are currently adapted for renal fibrosis identified for different disease entities, of which some are likely related
(Berchtold et al., 2017; Morrell et al., 2017). On the contrary, lung and to fibrosis. For instance, sarcomere-associated genetic variations in the
heart fibrosis are oftentimes diagnostically evaluated by functional tests β-myosin heavy chain (MYH7) and myosin-binding protein C
(e.g., vital capacity of the lung or left ventricular ejection fraction of the (MYBPC3) are linked to hypertrophic cardiomyopathy and interstitial
heart), using traditional methods like body plethysmography, echo- fibrosis (Marian and Braunwald et al., 2017). Genetic risk factors for
cardiography or cardiac MRI (Table 2). kidney fibrosis appear less well defined, but polymorphisms in the
In contrast to histology, imaging or mechanical assessment, circu- angiotensin converting enzyme or transforming growth factor (TGF)-β1
lating biomarkers of fibrosis are simple, reproducible and inexpensive. have been associated with increased renal scarring after urinary tract
Their biggest disadvantage, however, is their general lack of specificity infections (Zaffanello et al., 2011). Large whole-exome sequencing
for fibrosis, the large overlap between fibrosis stages (which hampers studies are ongoing that are expected to identify new risk factors for
accurate staging) and the many confounding factors for single markers. renal fibrosis (Lata et al., 2018). Regarding liver fibrosis, genetic sus-
In pulmonary fibrosis, many mediators involved in alveolar epithelial ceptibility has long been recognized (Zimmer and Lammert, 2011). A
cell injury, inflammation, and matrix remodeling have been proposed particular focus of current research are genetic risk factors for fibrosis
like chemokines (IL-8, CCL18), proteases (MMP-1, MMP-7) or growth progression in non-alcoholic fatty liver disease. The I148M variant of
factors, but their utility in clinical routine is not yet established Patatin-like phospholipase domain-containing protein 3 (PNPLA3,
(Drakopanagiotakis et al., 2018; Guiot et al., 2017). For kidney fibrosis, “adiponutrin”) is a major risk genetic risk factor for disease progression,
biomarkers are routinely used to monitor glomerular filtration rate and significant contributions have been also found for other genes such
(e.g., creatinine or cystatin C) or tubular damage (e.g., KIM-1, IL-18, as TM6SF2, MBOAT7 and GCKR (Eslam et al., 2018).
urinary NGAL), while biomarkers detecting renal fibrosis are not widely Novel targets in fibrosis biomarker research include circulating
accepted. Promising potential biomarkers for renal fibrosis include micro-RNA (miRNA), long non-coding RNA, epigenetic changes, mi-
TGF-β, monocyte chemoattractant protein-1 (MCP-1) and MMP-2 or the tochondrial DNA or microbiome signatures in the stool
urinary excretion of collagen III and its associated protein like N- (Drakopnagiotakis et al., 2018; Tan et al., 2014; Hardy et al., 2017;
terminal propeptide of type III collagen (Mansour et al., 2017; Ghoul Loomba et al., 2017; Chen et al., 2017). It appears biologically plausible
et al., 2010). In the setting of suspected liver fibrosis, manifold bio- to search for biomarkers that directly reflect pathogenic processes of
markers have been proposed and are to some extent routinely used fibrogenesis or fibrolysis. The different collagens and collagen frag-
(e.g., hyaluronic acid or N-terminal propeptide of type III collagen, ments that are being released into the circulating might therefore have
PIIINP). Best diagnostic and prognostic accuracy is achieved, when the exceptional potential as biomarkers (Karsdal et al., 2017). The cleaved
biomarkers are implemented in predictive models that account distinct N-terminal propeptide of type III collagen (Pro-C3) might represent
risk factors and integrate information from multiple biomarkers (Vilar- such a biomarker that mirrors collagen deposition (Nielsen et al.,
Gomez and Chalasani, 2018). Such composite scores are the NAFLD 2015). In a similar direction, novel imaging methods combine visuali-
fibrosis scores, FIB-4 index, APRI score or the commercially available zation with molecular targeting. For instance, collagen- or elastin-

9
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

specific probes yielded promising results in molecular MRI of animal injury (Tacke, 2018). The landscape is similarly complex with many
models of liver, renal, cardiac and pulmonary fibrosis (Baues et al., compounds under clinical investigation for renal fibrosis (Klinkhammer
2017). et al., 2017) and pulmonary fibrosis (Raghu, 2017).
Surprisingly, not many of the currently proposed pharmacological
7. Perspective: general and organ-specific targets of antifibrotic therapies therapies target the myofibroblast directly, but rather modulate their
activating signals or intracellular cascades (Schon et al., 2016). Many
The identification of the manifold mediators and pathways acti- new research directions therefore aim at improving strategies for drug
vated in tissue fibrogenesis offers a multitude of potential targets for targeting to myofibroblasts in the liver, kidney or lung (Yazdani et al.,
antifibrotic drugs. However, up to now, only two antifibrotic drugs 2017). Potential targeting systems include modified albumin-, hy-
have been approved for idiopathic pulmonary fibrosis, while no specific drogel-, peptide-, nanoparticle-, aptamer- or antibody-based systems,
antifibrotic pharmacological therapy exists for liver, kidney or heart which demonstrated promising myofibroblast targeting in preclinical
fibrosis. And even in the case of pulmonary fibrosis, the two inhibitors models (Yazdani et al., 2017; Bartneck et al., 2014; Nastase et al.,
of profibrotic signaling, pirfenidone and nintedanib, improved forced 2017). In addition, cell-based strategies are being tested, in which al-
vital capacity and survival, but did not show histological resolution of logeneic or autologous mesenchymal stem cells or bone marrow-de-
pulmonary fibrosis (Martinez et al., 2017). Building on these experi- rived cells are systemically infused to induce restorative processes in
ences, early clinical data indicate that the monoclonal antibody FG- fibrotic organs like heart, lung, kidney or liver (Lim et al., 2017; Pandey
3019 that interferes with connective tissue growth factor (CTGF) sig- et al., 2017). The better understanding of the processes and mechan-
naling might convey antifibrotic activity in pulmonary fibrosis patients isms that contribute to organ fibrosis should open new avenues for
(Raghu et al., 2016). therapeutic interventions, ultimately helping to decrease the enormous
Despite the paucity of anti-fibrotic drugs approved to date, the morbidity and mortality linked to fibrotic diseases.
tremendous progress in deciphering the mechanisms of organ fibrosis
fueled enthusiasm in the field, giving rise to the expectation that anti- Grant support
fibrotic therapies will become a reality in the near future (Lee et al.,
2015; Lee et al., 2015; Klinkhammer et al., 2017; Murtha et al., 2017). This work was supported by the German Research Foundation (DFG;
A particular interest are so-called “core mechanisms of fibrosis” that are SFB/TRR57, TA434/3-1) and grants from the Interdisciplinary Centre
preserved between different organs; the expectation is that targeting for Clinical Research within the Faculty of Medicine at the RWTH
such pathways would provide compounds with anti-fibrotic activities Aachen University.
across different types of diseases and organs (Mehal et al., 2011). Such
pathways could be inflammation-linked mechanisms (e.g., in- Conflicts of interest
flammatory macrophage accumulation), receptor-ligand interactions
with subsequent fibrogenic intracellular signaling, myofibroblast acti- Work in the laboratory of Frank Tacke has been supported by
vation, or matrix synthesis, deposition and stabilization (Weiskirchen funding from Tobira Therapeutics and Galapagos. Ralf Weiskirchen
and Tacke, 2016). One of these core mechanisms is the stabilization of cooperates with Silence Therapeutics.
extracellular matrix mediated by crosslinking activity via the enzyme
lysyl oxidase (LOX) and its homologs, LOX-like enzymes 1–4 Acknowledgements
(LOXL1–LOXL4). Inhibition of LOXL2 by a neutralizing antibody de-
monstrated potent antifibrotic efficacy in animal models of heart, lung We cordially thank all members of our labs and collaborating sci-
and liver fibrosis (Barry-Hamilton et al., 2010; Yang et al., 2016; entists for helpful discussions.
Ikenaga et al., 2017). Unfortunately, the humanized monoclonal anti-
LOXL2 antibody simtuzumab failed to show antifibrotic activity in References
patients with pulmonary or hepatic fibrosis (Raghu et al., 2017;
Loomba et al., 2018). Other universal profibrogenic signaling cascades Abreu, J.G., Ketpura, N.I., Reversade, B., De Robertis, E.M., 2002. Connective-tissue
and cytokines, such as TGF-β, may be more difficult to neutralize in growth factor (CTGF) modulates cell signalling by BMP and TGF-β. Nat. Cell Biol. 4,
599–604. http://dx.doi.org/10.1038/ncb826.
fibrotic tissue (Rosenbloom et al., 2017). Aldigier, J.C., Kanjanbuch, T., Ma, L.J., Brown, N.J., Fogo, A.B., 2005. Regression of
Therefore, there is currently a trend towards organ-specific anti-fi- existing glomerulosclerosis by inhibition of aldosterone. J. Am. Soc. Nephrol. 16,
brotic drug development. This strategy carries the advantages of pre- 3306–3314. http://dx.doi.org/10.1681/ASN.2004090804.
Amano, Y., Kitamura, M., Takano, H., Yanagisawa, F., Tachi, M., Suzuki, Y., Kumita, S.,
clinical testing in multiple, translationally relevant experimental Takayama, M., 2018 Jan 18. Cardiac MR imaging of hypertrophic cardiomyopathy:
models, reducing systemic side effects by an organ-specific targeting techniques, findings, and clinical relevance. Magn. Reson. Med. Sci. http://dx.doi.
and a more tailored patient selection considering the proposed mode of org/10.2463/mrms.rev.2017-0145.
Andrews, J.P., Marttala, J., Macarak, E., Rosenbloom, J., Uitto, J., 2016. Keloids: the
action (Lee et al., 2015; Klinkhammer et al., 2017; Murtha et al., 2017). paradigm of skin fibrosis - pathomechanisms and treatment. Matrix Biol. 51, 37–46.
Many clinical studies are currently ongoing in the different fields of http://dx.doi.org/10.1016/j.matbio.2016.01.013.
organ fibrosis. For liver fibrosis, different compounds under investiga- Arciniegas, E., Sutton, A.B., Allen, T.D., Schor, A.M., 1992. Transforming growth factor
beta 1 promotes the differentiation of endothelial cells into smooth muscle-like cells
tion target the underlying disease etiology, mostly non-alcoholic stea-
in vitro. J. Cell Sci. 103, 521–529.
tohepatitis, by reducing hepatic fat accumulation, alteration of meta- Arias, M., Sauer-Lehnen, S., Treptau, J., Janoschek, N., Theuerkauf, I., Buettner, R.,
bolic pathways or resultant hepatocyte cell death. Examples for these Gressner, A.M., Weiskirchen, R., 2003. Adenoviral expression of a transforming
compounds include farnesoid X receptor (FXR) agonists (obeticholic growth factor-β1 antisense mRNA is effective in preventing liver fibrosis in bile-duct
ligated rats. BMC Gastroenterol. 3, 29. http://dx.doi.org/10.1186/1471-230X-3-29.
acid or non-steroidal FXR agonists), peroxisome proliferator-activator Aschner, Y., Downey, G.P., 2016. Transforming growth factor-β: master regulator of the
receptor agonists (e.g., elafibranor, lanifibranor, saroglitazar), acetyl- respiratory system in health and disease. Am. J. Respir. Cell Mol. Biol. 54, 647–655.
CoA carboxylase inhibitors (e.g., GS-0976), caspase inhibitors (e.g., http://dx.doi.org/10.1165/rcmb.2015-0391TR.
Baeck, C., Wehr, A., Karlmark, K.R., Heymann, F., Vucur, M., Gassler, N., Huss, S.,
emricasan) and fibroblast growth factor (FGF)-21 or FGF-19 analogues Klussmann, S., Eulberg, D., Luedde, T., Trautwein, C., Tacke, F., 2012.
(Rotman and Sanyal, 2017). More direct anti-fibrotic compounds under Pharmacological inhibition of the chemokine CCL2 (MCP-1) diminishes liver mac-
clinical investigation are selonsertib, an inhibitor of apoptosis signal- rophage infiltration and steatohepatitis in chronic hepatic injury. Gut 61, 416–426.
http://dx.doi.org/10.1136/gutjnl-2011-300304.
regulating kinase 1 (ASK1) that is activated in myofibroblasts and Bardadin, K.A., Desmet, V.J., 1985. Ultrastructural observations on sinusoidal endothelial
macrophages during fibrogenesis (Loomba et al., 2018), and cen- cells in chronic active hepatitis. Histopathology 9, 171–181. http://dx.doi.org/10.
icriviroc (CVC), an inhibitor of the chemokine receptors CCR2 and 1111/j.1365-2559.1985.tb02433.x.
Barnes, J.L., Gorin, Y., 2011. Myofibroblast differentiation during fibrosis: role of NAD(P)
CCR5 that mediate inflammatory monocyte recruitment upon liver

10
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

H oxidases. Kidney Int. 79, 944–956. http://dx.doi.org/10.1038/ki.2010.516. 1038/nm.2848.


Barry-Hamilton, V., Spangler, R., Marshall, D., McCauley, S., Rodriguez, H.M., Oyasu, M., El Mourabit, H., Loeuillard, E., Lemoinne, S., Cadoret, A., Housset, C., 2016. Culture
Mikels, A., Vaysberg, M., Ghermazien, H., Wai, C., Garcia, C.A., Velayo, A.C., model of rat portal myofibroblasts. Front. Physiol. 7, 120. http://dx.doi.org/10.
Jorgensen, B., Biermann, D., Tsai, D., Green, J., Zaffryar-Eilot, S., Holzer, A., Ogg, S., 3389/fphys.2016.00120.
Thai, D., Neufeld, G., Van Vlasselaer, P., Smith, V., 2010. Allosteric inhibition of lysyl Eslam, M., Valenti, L., Romeo, S., 2018. Genetics and epigenetics of NAFLD and NASH:
oxidase-like-2 impedes the development of a pathologic microenvironment. Nat. clinical impact. J. Hepatol. 68, 268–279. http://dx.doi.org/10.1016/j.jhep.2017.09.
Med. 16, 1009–1017. http://dx.doi.org/10.1038/nm.2208. 003.
Bartneck, M., Warzecha, K.T., Tacke, F., 2014. Therapeutic targeting of liver inflamma- European Association for Study of Liver, 2015. Asociacion Latinoamericana para el
tion and fibrosis by nanomedicine. Hepatobiliary Surg. Nutr. 3, 364–376. http://dx. Estudio del Higado. EASL-ALEH Clinical Practice Guidelines: non-invasive tests for
doi.org/10.3978/j.issn.2304-3881.2014.11.02. evaluation of liver disease severity and prognosis. J. Hepatol. 63, 237–264. http://dx.
Baues, M., Dasgupta, A., Ehling, J., Prakash, J., Boor, P., Tacke, F., Kiessling, F., Lammers, doi.org/10.1016/j.jhep.2015.04.006.
T., 2017. Fibrosis imaging: current concepts and future directions. Adv. Drug Deliv. Fioretto, P., Steffes, M.W., Sutherland, D.E., Goetz, F.C., Mauer, M., 1998. Reversal of
Rev. 121, 9–26. http://dx.doi.org/10.1016/j.addr.2017.10.013. lesions of diabetic nephropathy after pancreas transplantation. N. Engl. J. Med. 339,
Berchtold, L., Friedli, I., Vallée, J.P., Moll, S., Martin, P.Y., de Seigneux, S., 2017. 69–75. http://dx.doi.org/10.1056/NEJM199807093390202.
Diagnosis and assessment of renal fibrosis: the state of the art. Swiss Med. Wkly. Forbes, S.J., Russo, F.P., Rey, V., Burra, P., Rugge, M., Wright, N.A., Alison, M.R., 2004. A
147http://dx.doi.org/10.4414/smw.2017.14442. w14442. significant proportion of myofibroblasts are of bone marrow origin in human liver
Borkham-Kamphorst, E., Kovalenko, E., van Roeyen, C.R., Gassler, N., Bomble, M., fibrosis. Gastroenterology 126, 955–963. http://dx.doi.org/10.1053/j.gastro.2004.
Ostendorf, T., Floege, J., Gressner, A.M., Weiskirchen, R., 2008. Platelet-derived 02.025.
growth factor isoform expression in carbon tetrachloride-induced chronic liver in- Frangogiannis, N.G., Dewald, O., Xia, Y., Ren, G., Haudek, S., Leucker, T., Kraemer, D.,
jury. Lab. Invest. 88, 1090–1100. http://dx.doi.org/10.1038/labinvest.2008.71. Taffet, G., Rollins, B.J., Entman, M.L., 2007. Critical role of monocyte chemoat-
Borkham-Kamphorst, E., van de Leur, E., Zimmermann, H.W., Karlmark, K.R., Tihaa, L., tractant protein-1/CC chemokine ligand 2 in the pathogenesis of ischemic cardio-
Haas, U., Tacke, F., Berger, T., Mak, T.W., Weiskirchen, R., 2013. Protective effects of myopathy. Circulation 115, 584–592. http://dx.doi.org/10.1161/
lipocalin-2 (LCN2) in acute liver injury suggest a novel function in liver homeostasis. CIRCULATIONAHA.106.646091.
Biochim. Biophys. Acta 1832, 660–673. http://dx.doi.org/10.1016/j.bbadis.2013.01. Frantz, S., Falcao-Pires, I., Balligand, J.L., Bauersachs, J., Brutsaert, D., Ciccarelli, M.,
014. Dawson, D., de Windt, L.J., Giacca, M., Hamdani, N., Hilfiker-Kleiner, D., Hirsch, E.,
Borkham-Kamphorst, E., Weiskirchen, R., 2016. The PDGF system and its antagonists in Leite-Moreira, A., Mayr, M., Thum, T., Tocchetti, C.G., van der Velden, J., Varricchi,
liver fibrosis. Cytokine Growth Factor Rev. 28, 53–61. http://dx.doi.org/10.1016/j. G., Heymans, S., 2018. The innate immune system in chronic cardiomyopathy: a
cytogfr.2015.10.002. European Society of Cardiology (ESC) scientific statement from the Working Group
Boursier, J., Vergniol, J., Guillet, A., Hiriart, J.B., Lannes, A., Le Bail, B., Michalak, S., on Myocardial Function of the ESC. Eur. J. Heart Fail. http://dx.doi.org/10.1002/
Chermak, F., Bertrais, S., Foucher, J., Oberti, F., Charbonnier, M., Fouchard-Hubert, ejhf.1138. Jan 15.
I., Rousselet, M.C., Calès, P., de Lédinghen, V., 2016. Diagnostic accuracy and Friedman, S.L., Roll, F.J., Boyles, J., Bissell, D.M., 1985. Hepatic lipocytes: the principal
prognostic significance of blood fibrosis tests and liver stiffness measurement by collagen-producing cells of normal rat liver. Proc. Natl. Acad. Sci. U. S. A. 82,
FibroScan in non-alcoholic fatty liver disease. J. Hepatol. 65, 570–578. http://dx.doi. 8681–8685. http://dx.doi.org/10.1073/pnas.82.24.8681.
org/10.1016/j.jhep.2016.04.023. Friedman, S.L., Ratziu, V., Harrison, S.A., Abdelmalek, M.F., Aithal, G.P., Caballeria, J.,
Bucala, R., Spiegel, L.A., Chesney, J., Hogan, M., Cerami, A., 1994. Circulating fibrocytes Francque, S., Farrell, G., Kowdley, K.V., Craxi, A., Simon, K., Fischer, L., Melchor-
define a new leukocyte subpopulation that mediates tissue repair. Mol. Med. (Camb.) Khan, L., Vest, J., Wiens, B.L., Vig, P., Seyedkazemi, S., Goodman, Z., Wong, V.W.,
1, 71–81. Loomba, R., Tacke, F., Sanyal, A., Lefebvre, E., 2018. A randomized, placebo-con-
Buhl, E.M., Djudjaj, S., Babickova, J., Klinkhammer, B.M., Folestad, E., Borkham- trolled trial of cenicriviroc for treatment of nonalcoholic steatohepatitis with fibrosis.
Kamphorst, E., Weiskirchen, R., Hudkins, K., Alpers, C.E., Eriksson, U., Floege, J., Hepatology 67, 1754–1767. https://doi.org/10.1002/hep.29477.
Boor, P., 2016. The role of PDGF-D in healthy and fibrotic kidneys. Kidney Int. 89, Friedrich-Rust, M., Poynard, T., Castera, L., 2016. Critical comparison of elastography
848–861. http://dx.doi.org/10.1016/j.kint.2015.12.037. methods to assess chronic liver disease. Nat. Rev. Gastroenterol. Hepatol. 13,
Caussy, C., Reeder, S.B., Sirlin, C.B., Loomba, R., 2018. Non-invasive, quantitative as- 402–411. http://dx.doi.org/10.1038/nrgastro.2016.86.
sessment of liver fat by MRI-PDFF as an endpoint in NASH trials. Hepatology. http:// Ghoul, B.E., Squalli, T., Servais, A., Elie, C., Meas-Yedid, V., Trivint, C., Vanmassenhove,
dx.doi.org/10.1002/hep.29797. Jan 21. J., Grünfeld, J.P., Olivo-Marin, J.C., Thervet, E., Noël, L.H., Prié, D., Fakhouri, F.,
Cavalera, M., Frangogiannis, N.G., 2014. Targeting the chemokines in cardiac repair. 2010. Urinary procollagen III aminoterminal propeptide (PIIINP): a fibrotest for the
Curr. Pharmaceut. Des. 20, 1971–1979. http://dx.doi.org/10.2174/ nephrologist. Clin. J. Am. Soc. Nephrol. 5, 205–210. http://dx.doi.org/10.2215/CJN.
13816128113199990449. 06610909.
Charo, I.F., Ransohoff, R.M., 2006. The many roles of chemokines and chemokine re- Gieseck 3rd, R.L., Wilson, M.S., Wynn, T.A., 2018. Type 2 immunity in tissue repair and
ceptors in inflammation. N. Engl. J. Med. 354, 610–621. http://dx.doi.org/10.1056/ fibrosis. Nat. Rev. Immunol. 18, 62–76. http://dx.doi.org/10.1038/nri.2017.90.
NEJMra052723. Greenwel, P., Domínguez-Rosales, J.A., Mavi, G., Rivas-Estilla, A.M., Rojkind, M., 2000.
Chen, Z., Li, C., Lin, K., Cai, H., Ruan, W., Han, J., Rao, L., 2017. Non-coding RNAs in Hydrogen peroxide: a link between acetaldehyde-elicited alpha1(I) collagen gene up-
cardiac fibrosis: emerging biomarkers and therapeutic targets. Cardiol. J. http://dx. regulation and oxidative stress in mouse hepatic stellate cells. Hepatology 31,
doi.org/10.5603/CJ.a2017.0153. Dec 14. 109–116. http://dx.doi.org/10.1002/hep.510310118.
Cordero-Espinoza, L., Huch, M., 2018. The balancing act of the liver: tissue regeneration Gressner, A.M., Weiskirchen, R., Breitkopf, K., Dooley, S., 2002. Roles of TGF-β in hepatic
versus fibrosis. J. Clin. Invest. 128, 85–96. http://dx.doi.org/10.1172/JCI93562. fibrosis. Front. Biosci. 7, d793–807. http://dx.doi.org/10.2741/A812.
Danobeitia, J.S., Djamali, A., Fernandez, L.A., 2014. The role of complement in the pa- Gressner, A.M., Weiskirchen, R., 2006. Modern pathogenetic concepts of liver fibrosis
thogenesis of renal ischemia-reperfusion injury and fibrosis. Fibrogenesis Tissue suggest stellate cells and TGF-β as major players and therapeutic targets. J. Cell Mol.
Repair 7, 16. http://dx.doi.org/10.1186/1755-1536-7-16. Med. 10, 76–99. http://dx.doi.org/10.1111/j.1582-4934.2006.tb00292.x.
De Bleser, P.J., Xu, G., Rombouts, K., Rogiers, V., Geerts, A., 1999. Glutathione levels Grgic, I., Campanholle, G., Bijol, V., Wang, C., Sabbisetti, V.S., Ichimura, T., Humphreys,
discriminate between oxidative stress and transforming growth factor-β signaling in B.D., Bonventre, J.V., 2012. Targeted proximal tubule injury triggers interstitial fi-
activated rat hepatic stellate cells. J. Biol. Chem. 274, 33881–33887. http://dx.doi. brosis and glomerulosclerosis. Kidney Int. 82, 172–183. http://dx.doi.org/10.1038/
org/10.1074/jbc.274.48.33881. ki.2012.20.
Del Campo, J.A., Gallego, P., Grande, L., 2018. Role of inflammatory response in liver Guiot, J., Moermans, C., Henket, M., Corhay, J.L., Louis, R., 2017. Blood biomarkers in
diseases: therapeutic strategies. World J. Hepatol. 10, 1–7. http://dx.doi.org/10. idiopathic pulmonary fibrosis. Lung 195, 273–280. http://dx.doi.org/10.1007/
4254/wjh.v10.i1.1. s00408-017-9993-5.
Di Carlo, S.E., Peduto, L., 2018. The perivascular origin of pathological fibroblasts. J. Guyot, C., Lepreux, S., Combe, C., Doudnikoff, E., Bioulac-Sage, P., Balabaud, C.,
Clin. Invest. 128, 54–63. http://dx.doi.org/10.1172/JCI93558. Desmoulière, A., 2006. Hepatic fibrosis and cirrhosis: the (myo)fibroblastic cell
Douillet, C.D., Velarde, V., Christopher, J.T., Mayfield, R.K., Trojanowska, M.E., Jaffa, subpopulations involved. Int. J. Biochem. Cell Biol. 38, 135–151.
A.A., 2000. Mechanisms by which bradykinin promotes fibrosis in vascular smooth Hammel, P., Couvelard, A., O'Toole, D., Ratouis, A., Sauvanet, A., Fléjou, J.F., Degott, C.,
muscle cells: role of TGF-beta and MAPK. Am. J. Physiol. Heart Circ. Physiol. 279, Belghiti, J., Bernades, P., Valla, D., Ruszniewski, P., Lévy, P., 2001. Regression of
H2829–H2837. http://dx.doi.org/10.1152/ajpheart.2000.279.6.H2829. liver fibrosis after biliary drainage in patients with chronic pancreatitis and stenosis
Drakopanagiotakis, F., Wujak, L., Wygrecka, M., Markart, P., 2018. Biomarkers in idio- of the common bile duct. N. Engl. J. Med. 344, 418–423. http://dx.doi.org/10.1056/
pathic pulmonary fibrosis. Matrix Biol. http://dx.doi.org/10.1016/j.matbio.2018.01. NEJM200102083440604.
023. Jan 30. Hammerich, L., Tacke, F., 2014. Interleukins in chronic liver disease: lessons learned from
Duffield, J.S., Forbes, S.J., Constandinou, C.M., Clay, S., Partolina, M., Vuthoori, S., Wu, experimental mouse models. Clin. Exp. Gastroenterol. 7, 297–306. http://dx.doi.org/
S., Lang, R., Iredale, J.P., 2005. Selective depletion of macrophages reveals distinct, 10.2147/CEG.S43737.
opposing roles during liver injury and repair. J. Clin. Invest. 115, 56–65. http://dx. Hardy, T., Zeybel, M., Day, C.P., Dipper, C., Masson, S., McPherson, S., Henderson, E.,
doi.org/10.1172/JCI200522675. Tiniakos, D., White, S., French, J., Mann, D.A., Anstee, Q.M., Mann, J., 2017. Plasma
Dulai, P.S., Singh, S., Patel, J., Soni, M., Prokop, L.J., Younossi, Z., Sebastiani, G., Ekstedt, DNA methylation: a potential biomarker for stratification of liver fibrosis in non-
M., Hagstrom, H., Nasr, P., Stal, P., Wong, V.W., Kechagias, S., Hultcrantz, R., alcoholic fatty liver disease. Gut 66, 1321–1328. http://dx.doi.org/10.1136/gutjnl-
Loomba, R., 2017. Increased risk of mortality by fibrosis stage in nonalcoholic fatty 2016-311526.
liver disease: systematic review and meta-analysis. Hepatology 65, 1557–1565. Hashimoto, N., Jin, H., Liu, T., Chensue, S.W., Phan, S.H., 2004. Bone marrow-derived
http://dx.doi.org/10.1002/hep.29085. progenitor cells in pulmonary fibrosis. J. Clin. Invest. 113, 243–252. http://dx.doi.
Dulauroy, S., Di Carlo, S.E., Langa, F., Eberl, G., Peduto, L., 2012. Lineage tracing and org/10.1172/JCI200418847.
genetic ablation of ADAM12+ perivascular cells identify a major source of profibrotic Hashimoto, N., Phan, S.H., Imaizumi, K., Matsuo, M., Nakashima, H., Kawabe, T.,
cells during acute tissue injury. Nat. Med. 18, 1262–1270. http://dx.doi.org/10. Shimokata, K., Hasegawa, Y., 2010. Endothelial-mesenchymal transition in

11
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

bleomycin-induced pulmonary fibrosis. Am. J. Respir. Cell Mol. Biol. 43, 161–172. Kong, P., Christia, P., Frangogiannis, N.G., 2014. The pathogenesis of cardiac fibrosis.
http://dx.doi.org/10.1165/rcmb.2009-0031OC. Cell. Mol. Life Sci. 71, 549–574. http://dx.doi.org/10.1007/s00018-013-1349-6.
Heymann, F., Tacke, F., 2016. Immunology in the liver–from homeostasis to disease. Nat. Kovalovich, K., DeAngelis, R.A., Li, W., Furth, E.E., Ciliberto, G., Taub, R., 2000.
Rev. Gastroenterol. Hepatol. 13, 88–110. http://dx.doi.org/10.1038/nrgastro.2015. Increased toxin-induced liver injury and fibrosis in interleukin-6-deficient mice.
200. Hepatology 31, 149–159. http://dx.doi.org/10.1002/hep.510310123.
Hirsova, P., Gores, G.J., 2015. Death receptor-mediated cell death and proinflammatory Kramann, R., Schneider, R.K., DiRocco, D.P., Machado, F., Fleig, S., Bondzie, P.A.,
signaling in nonalcoholic steatohepatitis. Cell Mol Gastroenterol Hepatol 1, 17–27. Henderson, J.M., Ebert, B.L., Humphreys, B.D., 2015. Perivascular Gli1+ progenitors
http://dx.doi.org/10.1016/j.jcmgh.2014.11.005. are key contributors to injury-induced organ fibrosis. Cell Stem Cell. 16, 51–66.
Humphreys, B.D., Lin, S.L., Kobayashi, A., Hudson, T.E., Nowlin, B.T., Bonventre, J.V., http://dx.doi.org/10.1016/j.stem.2014.11.004.
Valerius, M.T., McMahon, A.P., Duffield, J.S., 2010. Fate tracing reveals the pericyte Krenkel, O., Tacke, F., 2017. Liver macrophages in tissue homeostasis and disease. Nat.
and not epithelial origin of myofibroblasts in kidney fibrosis. Am. J. Pathol. 176, Rev. Immunol. 17, 306–321. http://dx.doi.org/10.1038/nri.2017.11.
85–97. http://dx.doi.org/10.2353/ajpath.2010.090517. Krenkel, O., Puengel, T., Govaere, O., Abdallah, A.T., Mossanen, J.C., Kohlhepp, M.,
Hung, C., Linn, G., Chow, Y.H., Kobayashi, A., Mittelsteadt, K., Altemeier, W.A., Gharib, Liepelt, A., Lefebvre, E., Luedde, T., Hellerbrand, C., Weiskirchen, R., Longerich, T.,
S.A., Schnapp, L.M., Duffield, J.S., 2013. Role of lung pericytes and resident fibro- Costa, I.G., Anstee, Q.M., Trautwein, C., Tacke, F., 2018. Therapeutic inhibition of
blasts in the pathogenesis of pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 188, inflammatory monocyte recruitment reduces steatohepatitis and liver fibrosis.
820–830. http://dx.doi.org/10.1164/rccm.201212-2297OC. Hepatology 67, 1270–1283. http://dx.doi.org/10.1002/hep.29544.
Ikenaga, N., Peng, Z.W., Vaid, K.A., Liu, S.B., Yoshida, S., Sverdlov, D.Y., Mikels-Vigdal, Kubo, H., 2018. Extracellular vesicles in lung disease. Chest 153, 210–216. http://dx.doi.
A., Smith, V., Schuppan, D., Popov, Y.V., 2017. Selective targeting of lysyl oxidase- org/10.1016/j.chest.2017.06.026.
like 2 (LOXL2) suppresses hepatic fibrosis progression and accelerates its reversal. Kumar, V., Mitchell, M.D., Umscheid, C.A., Berns, J.S., Hogan, J.J., 2018. Risk of com-
Gut 66, 1697–1708. http://dx.doi.org/10.1136/gutjnl-2016-312473. plications with use of aspirin during renal biopsy: a systematic review. Clin. Nephrol.
Iqbal, S.A., Sidgwick, G.P., Bayat, A., 2012. Identification of fibrocytes from mesenchymal 89, 67–76. http://dx.doi.org/10.5414/CN109274.
stem cells in keloid tissue: a potential source of abnormal fibroblasts in keloid scar- Kurts, C., Panzer, U., Anders, H.J., Rees, A.J., 2013. The immune system and kidney
ring. Arch. Dermatol. Res. 304, 665–671. http://dx.doi.org/10.1007/s00403-012- disease: basic concepts and clinical implications. Nat. Rev. Immunol. 13, 738–753.
1225-5. http://dx.doi.org/10.1038/nri3523.
Iwayama, T., Olson, L.E., 2013. Involvement of PDGF in fibrosis and scleroderma: recent Lan, T., Kisseleva, T., Brenner, D.A., 2015. Deficiency of NOX1 or NOX4 prevents liver
insights from animal models and potential therapeutic opportunities. Curr. inflammation and fibrosis in mice through inhibition of hepatic stellate cell activa-
Rheumatol. Rep. 15, 304. http://dx.doi.org/10.1007/s11926-012-0304-0. tion. PLoS One (7), 10. http://dx.doi.org/10.1371/journal.pone.0129743. e0129743.
Jeong, D., Lee, M.A., Li, Y., Yang, D.K., Kho, C., Oh, J.G., Hong, G., Lee, A., Song, M.H., Lata, S., Marasa, M., Li, Y., Fasel, D.A., Groopman, E., Jobanputra, V., Rasouly, H.,
LaRocca, T.J., Chen, J., Liang, L., Mitsuyama, S., D'Escamard, V., Kovacic, J.C., Kwak, Mitrotti, A., Westland, R., Verbitsky, M., Nestor, J., Slater, L.M., D'Agati, V., Zaniew,
T.H., Hajjar, R.J., Park, W.J., 2016. Matricellular protein CCN5 reverses established M., Materna-Kiryluk, A., Lugani, F., Caridi, G., Rampoldi, L., Mattoo, A., Newton,
cardiac fibrosis. J. Am. Coll. Cardiol. 67, 1556–1568. http://dx.doi.org/10.1016/j. C.A., Rao, M.K., Radhakrishnan, J., Ahn, W., Canetta, P.A., Bomback, A.S., Appel,
jacc.2016.01.030. G.B., Antignac, C., Markowitz, G.S., Garcia, C.K., Kiryluk, K., Sanna-Cherchi, S.,
Jones, D.P., True, H.D., Patel, J., 2017. Leukocyte trafficking in cardiovascular disease: Gharavi, A.G., 2018. Whole-exome sequencing in adults with chronic kidney disease:
insights from experimental models. Mediat. Inflamm. 2017, 9746169. http://dx.doi. a pilot study. Ann. Intern. Med. 168, 100–109. http://dx.doi.org/10.7326/M17-
org/10.1155/2017/9746169. 1319.
Jun, J.I., Lau, L.F., 2018. Resolution of organ fibrosis. J. Clin. Invest. 128, 97–107. http:// LeBleu, V.S., Taduri, G., O'Connell, J., Teng, Y., Cooke, V.G., Woda, C., Sugimoto, H.,
dx.doi.org/10.1172/JCI93563. Kalluri, R., 2013. Origin and function of myofibroblasts in kidney fibrosis. Nat. Med.
Kalluri, R., Weinberg, R.A., 2009. The basics of epithelial-mesenchymal transition. J. Clin. 19, 1047–1053. http://dx.doi.org/10.1038/nm.3218.
Invest. 119, 1420–1428. http://dx.doi.org/10.1172/JCI39104. Lee, S.Y., Kim, S.I., Choi, M.E., 2015. Therapeutic targets for treating fibrotic kidney
Karimi-Shah, B.A., Chowdhury, B.A., 2015. Forced vital capacity in idiopathic pulmonary diseases. Transl. Res. 165, 512–530. http://dx.doi.org/10.1016/j.trsl.2014.07.010.
fibrosis–FDA review of pirfenidone and nintedanib. N. Engl. J. Med. 372, 1189–1191. Lee, Y.A., Wallace, M.C., Friedman, S.L., 2015. Pathobiology of liver fibrosis: a transla-
http://dx.doi.org/10.1056/NEJMp1500526. tional success story. Gut 64, 830–841. http://dx.doi.org/10.1136/gutjnl-2014-
Karlmark, K.R., Weiskirchen, R., Zimmermann, H.W., Gassler, N., Ginhoux, F., Weber, C., 306842. Erratum in: Gut 2015;64:1337.
Merad, M., Luedde, T., Trautwein, C., Tacke, F., 2009. Hepatic recruitment of the Lemoinne, S., Cadoret, A., El Mourabit, H., Thabut, D., Housset, C., 2013. Origins and
inflammatory Gr1+ monocyte subset upon liver injury promotes hepatic fibrosis. functions of liver myofibroblasts. Biochim. Biophys. Acta 1832, 948–954. http://dx.
Hepatology 50, 261–274. http://dx.doi.org/10.1002/hep.22950. doi.org/10.1016/j.bbadis.2013.02.019.
Karsdal, M.A., Nielsen, S.H., Leeming, D.J., Langholm, L.L., Nielsen, M.J., Manon-Jensen, Li, Y., Wang, J., Asahina, K., 2013. Mesothelial cells give rise to hepatic stellate cells and
T., Siebuhr, A., Gudmann, N.S., Rønnow, S., Sand, J.M., Daniels, S.J., Mortensen, myofibroblasts via mesothelial-mesenchymal transition in liver injury. Proc. Natl.
J.H., Schuppan, D., 2017. The good and the bad collagens of fibrosis - their role in Acad. Sci. U. S. A. 110, 2324–2329. http://dx.doi.org/10.1073/pnas.1214136110.
signaling and organ function. Adv. Drug Deliv. Rev. 121, 43–56. http://dx.doi.org/ Lim, R., Ricardo, S.D., Sievert, W., 2017. Cell-based therapies for tissue fibrosis. Front.
10.1016/j.addr.2017.07.014. Pharmacol. 8, 633. http://dx.doi.org/10.3389/fphar.2017.00633.
Kaur, A., Mathai, S.K., Schwartz, D.A., 2017. Genetics in idiopathic pulmonary fibrosis Lin, S.L., Kisseleva, T., Brenner, D.A., Duffield, J.S., 2008. Pericytes and perivascular fi-
pathogenesis, prognosis, and treatment. Front. Med. 4, 154. http://dx.doi.org/10. broblasts are the primary source of collagen-producing cells in obstructive fibrosis of
3389/fmed.2017.00154. the kidney. Am. J. Pathol. 173, 1617–1627. http://dx.doi.org/10.2353/ajpath.2008.
Kawaguchi, M., Takahashi, M., Hata, T., Kashima, Y., Usui, F., Morimoto, H., Izawa, A., 080433.
Takahashi, Y., Masumoto, J., Koyama, J., Hongo, M., Noda, T., Nakayama, J., Sagara, Liu, S., Taghavi, R., Leask, A., 2010. Connective tissue growth factor is induced in
J., Taniguchi, S., Ikeda, U., 2011. Inflammasome activation of cardiac fibroblasts is bleomycin-induced skin scleroderma. J Cell Commun Signal 4, 25–30. http://dx.doi.
essential for myocardial ischemia/reperfusion injury. Circulation 123, 594–604. org/10.1007/s12079-009-0081-3.
http://dx.doi.org/10.1161/CIRCULATIONAHA.110.982777. Liu, Y., Munker, S., Müllenbach, R., Weng, H.L., 2012. IL-13 signaling in liver fibrogen-
Kazlauskas, A., 2017. PDGFs and their receptors. Gene 614, 1–7. http://dx.doi.org/10. esis. Front. Immunol. 3, 116. http://dx.doi.org/10.3389/fimmu.2012.00116.
1016/j.gene.2017.03.003. Loomba, R., Seguritan, V., Li, W., Long, T., Klitgord, N., Bhatt, A., Dulai, P.S., Caussy, C.,
Keeley, E.C., Mehrad, B., Janardhanan, R., Salerno, M., Hunter, J.R., Burdick, M.M., Field, Bettencourt, R., Highlander, S.K., Jones, M.B., Sirlin, C.B., Schnabl, B., Brinkac, L.,
J.J., Strieter, R.M., Kramer, C.M., 2012. Elevated circulating fibrocyte levels in pa- Schork, N., Chen, C.H., Brenner, D.A., Biggs, W., Yooseph, S., Venter, J.C., Nelson,
tients with hypertensive heart disease. J. Hypertens. 30, 1856–1861. http://dx.doi. K.E., 2017. Gut microbiome-based metagenomic signature for non-invasive detection
org/10.1097/HJH.0b013e32835639bb. of advanced fibrosis in human nonalcoholic fatty liver disease. Cell Metabol. 25,
Kendall, R.T., Feghali-Bostwick, C.A., 2014. Fibroblasts in fibrosis: novel roles and 1054–1062. http://dx.doi.org/10.1016/j.cmet.2017.04.001. e5.
mediators. Front. Pharmacol. 5, 123. http://dx.doi.org/10.3389/fphar.2014.00123. Loomba, R., Lawitz, E., Mantry, P.S., Jayakumar, S., Caldwell, S.H., Arnold, H., Diehl,
Kisseleva, T., Cong, M., Paik, Y., Scholten, D., Jiang, C., Benner, C., Iwaisako, K., Moore- A.M., Djedjos, C.S., Han, L., Myers, R.P., Subramanian, G.M., McHutchison, J.G.,
Morris, T., Scott, B., Tsukamoto, H., Evans, S.M., Dillmann, W., Glass, C.K., Brenner, Goodman, Z.D., Afdhal, N.H., Charlton, M.R., 2018. GS-US-384-1497 Investigators.
D.A., 2012. Myofibroblasts revert to an inactive phenotype during regression of liver The ASK1 inhibitor selonsertib in patients with nonalcoholic steatohepatitis: a ran-
fibrosis. Proc. Natl. Acad. Sci. U. S. A. 109, 9448–9453. http://dx.doi.org/10.1073/ domized, phase 2 trial. Hepatology 67, 549–559. http://dx.doi.org/10.1002/hep.
pnas.1201840109. 29514.
Kinoshita, K., Iimuro, Y., Otogawa, K., Saika, S., Inagaki, Y., Nakajima, Y., Kawada, N., Low, R.B., Woodcock-Mitchell, J., Evans, J.N., Adler, K.B., 1984. Actin content of normal
Fujimoto, J., Friedman, S.L., Ikeda, K., 2007. Adenovirus-mediated expression of and of bleomycin-fibrotic rat lung. Am. Rev. Respir. Dis. 129, 311–316.
BMP-7 suppresses the development of liver fibrosis in rats. Gut 56, 706–714. http:// Lv, W., Booz, G.W., Wang, Y., Fan, F., Roman, R.J., 2018. Inflammation and renal fibrosis:
dx.doi.org/10.1136/gut.2006.092460. recent developments on key signaling molecules as potential therapeutic targets. Eur.
Kitagawa, K., Wada, T., Furuichi, K., Hashimoto, H., Ishiwata, Y., Asano, M., Takeya, M., J. Pharmacol. 820, 65–76. http://dx.doi.org/10.1016/j.ejphar.2017.12.016.
Kuziel, W.A., Matsushima, K., Mukaida, N., Yokoyama, H., 2004. Blockade of CCR2 Mäkelä, K., Hodgson, U., Piilonen, A., Kelloniemi, K., Bloigu, R., Sutinen, E., Salmenkivi,
ameliorates progressive fibrosis in kidney. Am. J. Pathol. 165, 237–246. http://dx. K., Rönty, M., Lappi-Blanco, E., Myllärniemi, M., Kaarteenaho, R., 2018. Analysis of
doi.org/10.1016/S0002-9440(10)63292-0. the histologic features associated with interobserver variation in idiopathic pul-
Klinkhammer, B.M., Goldschmeding, R., Floege, J., Boor, P., 2017. Treatment of renal monary fibrosis. Am. J. Surg. Pathol. http://dx.doi.org/10.1097/PAS.
fibrosis-turning challenges into opportunities. Adv. Chron. Kidney Dis. 24, 117–129. 0000000000001031. Feb 12.
http://dx.doi.org/10.1053/j.ackd.2016.11.002. Manea, S.A., Constantin, A., Manda, G., Sasson, S., Manea, A., 2015. Regulation of Nox
Kolahian, S., Fernandez, I.E., Eickelberg, O., Hartl, D., 2016. Immune mechanisms in enzymes expression in vascular pathophysiology: focusing on transcription factors
pulmonary fibrosis. Am. J. Respir. Cell Mol. Biol. 55, 309–322. http://dx.doi.org/10. and epigenetic mechanisms. Redox Biol. 5, 358–366. http://dx.doi.org/10.1016/j.
1165/rcmb.2016-0121TR. redox.2015.06.012.

12
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Manning, D.S., Afdhal, N.H., 2008. Diagnosis and quantitation of fibrosis. tomography image analysis in idiopathic pulmonary fibrosis: a mini review. Respir
Gastroenterology 134, 1670–1681. http://dx.doi.org/10.1053/j.gastro.2008.03.001. Investig 56, 5–13. http://dx.doi.org/10.1016/j.resinv.2017.10.003.
Mansour, S.G., Puthumana, J., Coca, S.G., Gentry, M., Parikh, C.R., 2017. Biomarkers for Okuma, T., Terasaki, Y., Kaikita, K., Kobayashi, H., Kuziel, W.A., Kawasuji, M., Takeya,
the detection of renal fibrosis and prediction of renal outcomes: a systematic review. M., 2004. C-C chemokine receptor 2 (CCR2) deficiency improves bleomycin-induced
BMC Nephrol. 18 (1), 72. http://dx.doi.org/10.1186/s12882-017-0490-0. pulmonary fibrosis by attenuation of both macrophage infiltration and production of
Marcellin, P., Gane, E., Buti, M., Afdhal, N., Sievert, W., Jacobson, I.M., Washington, macrophage-derived matrix metalloproteinases. J. Pathol. 204, 594–604. http://dx.
M.K., Germanidis, G., Flaherty, J.F., Aguilar Schall, R., Bornstein, J.D., Kitrinos, K.M., doi.org/10.1002/path.1667.
Subramanian, G.M., McHutchison, J.G., Heathcote, E.J., 2013. Regression of cirrhosis Österreicher, C.H., Penz-Österreicher, M., Grivennikov, S.I., Guma, M., Koltsova, E.K.,
during treatment with tenofovir disoproxil fumarate for chronic hepatitis B: a 5-year Datz, C., Sasik, R., Hardiman, G., Karin, M., Brenner, D.A., 2011. Fibroblast-specific
open-label follow-up study. Lancet 381, 468–475. http://dx.doi.org/10.1016/S0140- protein 1 identifies an inflammatory subpopulation of macrophages in the liver. Proc.
6736(12)61425-1. Natl. Acad. Sci. U. S. A. 108, 308–313. http://dx.doi.org/10.1073/pnas.1017547108.
Marian, A.J., Braunwald, E., 2017. Hypertrophic cardiomyopathy: genetics, pathogenesis, Pakshir, P., Hinz, B., 2018. The big five in fibrosis: macrophages, myofibroblasts, matrix,
clinical manifestations, diagnosis, and therapy. Circ. Res. 121, 749–770. http://dx. mechanics, and miscommunication. Matrix Biol. http://dx.doi.org/10.1016/j.
doi.org/10.1161/CIRCRESAHA.117.311059. matbio.2018.01.019. Jan 30. pii: S0945–053X(18)30034-9.
Marra, F., Tacke, F., 2014. Roles for chemokines in liver disease. Gastroenterology 147, Pandey, A.C., Lancaster, J.J., Harris, D.T., Goldman, S., Juneman, E., 2017. Cellular
577–594. http://dx.doi.org/10.1053/j.gastro.2014.06.043. e1. therapeutics for heart failure: focus on mesenchymal stem cells. Stem Cell. Int. 2017,
Martin, I.V., Borkham-Kamphorst, E., Zok, S., van Roeyen, C.R., Eriksson, U., Boor, P., 9640108. http://dx.doi.org/10.1155/2017/9640108.
Hittatiya, K., Fischer, H.P., Wasmuth, H.E., Weiskirchen, R., Eitner, F., Floege, J., Patel, K., Bedossa, P., Castera, L., 2015. Diagnosis of liver fibrosis: present and future.
Ostendorf, T., 2013. Platelet-derived growth factor (PDGF)-C neutralization reveals Semin. Liver Dis. 35, 166–183. http://dx.doi.org/10.1055/s-0035-1550059.
differential roles of PDGF receptors in liver and kidney fibrosis. Am. J. Pathol. 182, Phua, Y.L., Martel, N., Pennisi, D.J., Little, M.H., Wilkinson, L., 2013. Distinct sites of
107–117. http://dx.doi.org/10.1016/j.ajpath.2012.09.006. renal fibrosis in Crim1 mutant mice arise from multiple cellular origins. J. Pathol.
Martinez, F.J., Collard, H.R., Pardo, A., Raghu, G., Richeldi, L., Selman, M., Swigris, J.J., 229, 685–696. http://dx.doi.org/10.1002/path.4155.
Taniguchi, H., Wells, A.U., 2017. Idiopathic pulmonary fibrosis. Nat Rev Dis Primers Piera-Velazquez, S., Mendoza, F.A., Jimenez, S.A., 2016. Endothelial to mesenchymal
3, 17074. http://dx.doi.org/10.1038/nrdp.2017.74. transition (EndoMT) in the pathogenesis of human fibrotic diseases. J. Clin. Med. (4),
Mathai, S.K., Newton, C.A., Schwartz, D.A., Garcia, C.K., 2016. Pulmonary fibrosis in the 5. http://dx.doi.org/10.3390/jcm5040045. pii: E45.
era of stratified medicine. Thorax 71, 1154–1160. http://dx.doi.org/10.1136/ Quan, T.E., Cowper, S., Wu, S.P., Bockenstedt, L.K., Bucala, R., 2004. Circulating fi-
thoraxjnl-2016-209172. brocytes: collagen-secreting cells of the peripheral blood. Int. J. Biochem. Cell Biol.
McHedlidze, T., Waldner, M., Zopf, S., Walker, J., Rankin, A.L., Schuchmann, M., 36, 598–606. http://dx.doi.org/10.1016/j.biocel.2003.10.005.
Voehringer, D., McKenzie, A.N., Neurath, M.F., Pflanz, S., Wirtz, S., 2013. Raghu, G., Scholand, M.B., de Andrade, J., Lancaster, L., Mageto, Y., Goldin, J., Brown,
Interleukin-33-dependent innate lymphoid cells mediate hepatic fibrosis. Immunity K.K., Flaherty, K.R., Wencel, M., Wanger, J., Neff, T., Valone, F., Stauffer, J., Porter,
39, 357–371. http://dx.doi.org/10.1016/j.immuni.2013.07.018. S., 2016. FG-3019 anti-connective tissue growth factor monoclonal antibody: results
Mederacke, I., Hsu, C.C., Troeger, J.S., Huebener, P., Mu, X., Dapito, D.H., Pradere, J.P., of an open-label clinical trial in idiopathic pulmonary fibrosis. Eur. Respir. J. 47,
Schwabe, R.F., 2013. Fate tracing reveals hepatic stellate cells as dominant con- 1481–1491. http://dx.doi.org/10.1183/13993003.01030-2015.
tributors to liver fibrosis independent of its aetiology. Nat. Commun. 4, 2823. http:// Raghu, G., 2017. Pharmacotherapy for idiopathic pulmonary fibrosis: current landscape
dx.doi.org/10.1038/ncomms3823. and future potential. Eur. Respir. Rev. (145), 26. http://dx.doi.org/10.1183/
Mehal, W.Z., Iredale, J., Friedman, S.L., 2011. Scraping fibrosis: expressway to the core of 16000617.0071-2017.
fibrosis. Nat. Med. 17, 552–553. http://dx.doi.org/10.1038/nm0511-552. Raghu, G., Brown, K.K., Collard, H.R., Cottin, V., Gibson, K.F., Kaner, R.J., Lederer, D.J.,
Menne, J., Eulberg, D., Beyer, D., Baumann, M., Saudek, F., Valkusz, Z., Więcek, A., Martinez, F.J., Noble, P.W., Song, J.W., Wells, A.U., Whelan, T.P., Wuyts, W.,
Haller, H., 2017. Emapticap Study Group. C-C motif-ligand 2 inhibition with emap- Moreau, E., Patterson, S.D., Smith, V., Bayly, S., Chien, J.W., Gong, Q., Zhang, J.J.,
ticap pegol (NOX-E36) in type 2 diabetic patients with albuminuria. Nephrol. Dial. O'Riordan, T.G., 2017. Efficacy of simtuzumab versus placebo in patients with idio-
Transplant. 32, 307–315. http://dx.doi.org/10.1093/ndt/gfv459. pathic pulmonary fibrosis: a randomised, double-blind, controlled, phase 2 trial.
Meurer, S.K., Lahme, B., Tihaa, L., Weiskirchen, R., Gressner, A.M., 2005. N-acetyl-L- Lancet Respir Med 5, 22–32. http://dx.doi.org/10.1016/S2213-2600(16)30421-0.
cysteine suppresses TGF-β signaling at distinct molecular steps: the biochemical and Reddy, A.T., Lakshmi, S.P., Zhang, Y., Reddy, R.C., 2014. Nitrated fatty acids reverse
biological efficacy of a multifunctional, antifibrotic drug. Biochem. Pharmacol. 70, pulmonary fibrosis by dedifferentiating myofibroblasts and promoting collagen up-
1026–1034. http://dx.doi.org/10.1016/j.bcp.2005.07.001. take by alveolar macrophages. Faseb. J. 28, 5299–5310. http://dx.doi.org/10.1096/
Morrell, G.R., Zhang, J.L., Lee, V.S., 2017. Magnetic resonance imaging of the fibrotic fj.14-256263.
kidney. J. Am. Soc. Nephrol. 28, 2564–2570. http://dx.doi.org/10.1681/ASN. Ren, J., Wold, L.E., 2008. Mechanisms of alcoholic heart disease. Ther Adv Cardiovasc Dis
2016101089. 2, 497–506. http://dx.doi.org/10.1177/1753944708095137.
Munker, S., Wu, Y.L., Ding, H.G., Liebe, R., Weng, H.L., 2017. Can a fibrotic liver afford Reyes-Gordillo, K., Shah, R., Arellanes-Robledo, J., Hernández-Nazara, Z., Rincón-
epithelial-mesenchymal transition? World J. Gastroenterol. 23, 4661–4668. http:// Sánchez, A.R., Inagaki, Y., Rojkind, M., Lakshman, M.R., 2014. Mechanisms of action
dx.doi.org/10.3748/wjg.v23.i26.4661. v23.i26.4661" > . of acetaldehyde in the up-regulation of the human α2(I) collagen gene in hepatic
Murdaca, G., Contatore, M., Gulli, R., Mandich, P., Puppo, F., 2016. Genetic factors and stellate cells: key roles of Ski, SMAD3, SMAD4, and SMAD7. Am. J. Pathol. 184,
systemic sclerosis. Autoimmun. Rev. 15, 427–432. http://dx.doi.org/10.1016/j. 1458–1467. http://dx.doi.org/10.1016/j.ajpath.2014.01.020.
autrev.2016.01.016. Richter, K., Kietzmann, T., 2016. Reactive oxygen species and fibrosis: further evidence of
Murtha, L.A., Schuliga, M.J., Mabotuwana, N.S., Hardy, S.A., Waters, D.W., Burgess, J.K., a significant liaison. Cell Tissue Res. 365, 591–605. http://dx.doi.org/10.1007/
Knight, D.A., Boyle, A.J., 2017. The processes and mechanisms of cardiac and pul- s00441-016-2445-3.
monary fibrosis. Front. Physiol. 8, 777. http://dx.doi.org/10.3389/fphys.2017. Rinkevich, Y., Walmsley, G.G., Hu, M.S., Maan, Z.N., Newman, A.M., Drukker, M.,
00777. Januszyk, M., Krampitz, G.W., Gurtner, G.C., Lorenz, H.P., Weissman, I.L., Longaker,
Murphy-Marshman, H., Quensel, K., Shi-Wen, X., Barnfield, R., Kelly, J., Peidl, A., M.T., 2015. Skin fibrosis. Identification and isolation of a dermal lineage with in-
Stratton, R.J., Leask, A., 2017. Antioxidants and NOX1/NOX4 inhibition blocks trinsic fibrogenic potential. Science 348http://dx.doi.org/10.1126/science.aaa2151.
TGFβ1-induced CCN2 and α-SMA expression in dermal and gingival fibroblasts. PLoS aaa2151.
One (10), 12. http://dx.doi.org/10.1371/journal.pone.0186740. e0186740. Rock, J.R., Barkauskas, C.E., Cronce, M.J., Xue, Y., Harris, J.R., Liang, J., Noble, P.W.,
Nastase, M.V., Zeng-Brouwers, J., Wygrecka, M., Schaefer, L., 2017. Targeting renal fi- Hogan, B.L., 2011. Multiple stromal populations contribute to pulmonary fibrosis
brosis: mechanisms and drug delivery systems. Adv. Drug Deliv. Rev. http://dx.doi. without evidence for epithelial to mesenchymal transition. Proc. Natl. Acad. Sci. U. S.
org/10.1016/j.addr.2017.12.019. Dec 27. A. 108, E1475–E1483. http://dx.doi.org/10.1073/pnas.1117988108.
Neuschwander-Tetri, B.A., Loomba, R., Sanyal, A.J., Lavine, J.E., Van Natta, M.L., Romero, F., Shah, D., Duong, M., Penn, R.B., Fessler, M.B., Madenspacher, J., Stafstrom,
Abdelmalek, M.F., Chalasani, N., Dasarathy, S., Diehl, A.M., Hameed, B., Kowdley, W., Kavuru, M., Lu, B., Kallen, C.B., Walsh, K., Summer, R., 2015. A pneumocyte-
K.V., McCullough, A., Terrault, N., Clark, J.M., Tonascia, J., Brunt, E.M., Kleiner, macrophage paracrine lipid axis drives the lung toward fibrosis. Am. J. Respir. Cell
D.E., Doo, E., 2015. NASH Clinical Research Network. Farnesoid X nuclear receptor Mol. Biol. 53, 74–86. http://dx.doi.org/10.1165/rcmb.2014-0343OC.
ligand obeticholic acid for non-cirrhotic, non-alcoholic steatohepatitis (FLINT): a Rosenbloom, J., Macarak, E., Piera-Velazquez, S., Jimenez, S.A., 2017. Human fibrotic
multicentre, randomised, placebo-controlled trial. Lancet 385, 956–965. http://dx. diseases: current challenges in fibrosis research. Meth. Mol. Biol. 1627, 1–23. http://
doi.org/10.1016/S0140-6736(14)61933-4. Erratum in: Lancet 2015;385:946. Lancet dx.doi.org/10.1007/978-1-4939-7113-8_1.
2016;387:1618. Rotman, Y., Sanyal, A.J., 2017. Current and upcoming pharmacotherapy for non-alco-
Nielsen, M.J., Veidal, S.S., Karsdal, M.A., Ørsnes-Leeming, D.J., Vainer, B., Gardner, S.D., holic fatty liver disease. Gut 66, 180–190. http://dx.doi.org/10.1136/gutjnl-2016-
Hamatake, R., Goodman, Z.D., Schuppan, D., Patel, K., 2015. Plasma Pro-C3 (N- 312431.
terminal type III collagen propeptide) predicts fibrosis progression in patients with Rygiel, K.A., Robertson, H., Marshall, H.L., Pekalski, M., Zhao, L., Booth, T.A., Jones, D.E.,
chronic hepatitis C. Liver Int. 35, 429–437. http://dx.doi.org/10.1111/liv.12700. Burt, A.D., Kirby, J.A., 2008. Epithelial-mesenchymal transition contributes to portal
Ninichuk, V., Clauss, S., Kulkarni, O., Schmid, H., Segerer, S., Radomska, E., Eulberg, D., tract fibrogenesis during human chronic liver disease. Lab. Invest. 88, 112–123.
Buchner, K., Selve, N., Klussmann, S., Anders, H.J., 2008. Late onset of Ccl2 blockade http://dx.doi.org/10.1038/labinvest.3700704.
with the Spiegelmer mNOX-E36-3'PEG prevents glomerulosclerosis and improves Sánchez-Valle, V., Chávez-Tapia, N.C., Uribe, M., Méndez-Sánchez, N., 2012. Role of
glomerular filtration rate in db/db mice. Am. J. Pathol. 172, 628–637. http://dx.doi. oxidative stress and molecular changes in liver fibrosis: a review. Curr. Med. Chem.
org/10.2353/ajpath.2008.070601. 19, 4850–4860. http://dx.doi.org/10.2174/092986712803341520.
Nishioka, Y., Azuma, M., Kishi, M., Aono, Y., 2013. Targeting platelet-derived growth Schon, H.T., Bartneck, M., Borkham-Kamphorst, E., Nattermann, J., Lammers, T., Tacke,
factor as a therapeutic approach in pulmonary fibrosis. J. Med. Invest. 60, 175–183. F., Weiskirchen, R., 2016. Pharmacological intervention in hepatic stellate cell acti-
http://dx.doi.org/10.2152/jmi.60.175. vation and hepatic fibrosis. Front. Pharmacol. 7, 33. http://dx.doi.org/10.3389/
Ohkubo, H., Nakagawa, H., Niimi, A., 2018. Computer-based quantitative computed fphar.2016.00033.

13
R. Weiskirchen et al. Molecular Aspects of Medicine xxx (xxxx) xxx–xxx

Shull, M.M., Ormsby, I., Kier, A.B., Pawlowski, S., Diebold, R.J., Yin, M., Allen, R., fibrosis in mice. Kidney Int. 87, 948–962. http://dx.doi.org/10.1038/ki.2014.386.
Sidman, C., Proetzel, G., Calvin, D., et al., 1992. Targeted disruption of the mouse Weiskirchen, R., Meurer, S.K., 2013. BMP-7 counteracting TGF-β1 activities in organ fi-
transforming growth factor-β1 gene results in multifocal inflammatory disease. brosis. Front. Biosci. 18, 1407–1434. http://dx.doi.org/10.2741/4189.
Nature 359, 693–699. http://dx.doi.org/10.1038/359693a0. Weiskirchen, R., Tacke, F., 2014. Cellular and molecular functions of hepatic stellate cells
Strutz, F., Okada, H., Lo, C.W., Danoff, T., Carone, R.L., Tomaszewski, J.E., Neilson, E.G., in inflammatory responses and liver immunology. Hepatobiliary Surg. Nutr. 3,
1995. Identification and characterization of a fibroblast marker: FSP1. J. Cell Biol. 344–363. http://dx.doi.org/10.3978/j.issn.2304-3881.2014.11.03.
130, 393–405. Weiskirchen, R., Tacke, F., 2016. Liver fibrosis: from pathogenesis to novel therapies. Dig.
Su, H., Lei, C.T., Zhang, C., 2017. Interleukin-6 signaling pathway and its role in kidney Dis. 34, 410–422. http://dx.doi.org/10.1159/000444556.
disease: an update. Front. Immunol. 8, 405. http://dx.doi.org/10.3389/fimmu.2017. Weiskirchen, R., 2016. Hepatoprotective and anti-fibrotic agents: it's time to take the next
00405. step. Front. Pharmacol. 6, 303. http://dx.doi.org/10.3389/fphar.2015.00303.
Sun, M., Chen, M., Dawood, F., Zurawska, U., Li, J.Y., Parker, T., Kassiri, Z., Kirshenbaum, Weiskirchen, R., Tacke, F., 2017. Interleukin-33 in the pathogenesis of liver fibrosis:
L.A., Arnold, M., Khokha, R., Liu, P.P., 2007. Tumor necrosis factor-alpha mediates alarming ILC2 and hepatic stellate cells. Cell. Mol. Immunol. 14, 143–145. http://dx.
cardiac remodeling and ventricular dysfunction after pressure overload state. doi.org/10.1038/cmi.2016.62.
Circulation 115, 1398–1407. http://dx.doi.org/10.1161/CIRCULATIONAHA.106. Willis, B.C., Liebler, J.M., Luby-Phelps, K., Nicholson, A.G., Crandall, E.D., du Bois, R.M.,
643585. Borok, Z., 2005. Induction of epithelial-mesenchymal transition in alveolar epithelial
Sundberg, C., Ivarsson, M., Gerdin, B., Rubin, K., 1996. Pericytes as collagen-producing cells by transforming growth factor-beta1: potential role in idiopathic pulmonary
cells in excessive dermal scarring. Lab. Invest. 74, 452–466. fibrosis. Am. J. Pathol. 166, 1321–1332. http://dx.doi.org/10.1016/S0002-9440(10)
Svegliati-Baroni, G., Inagaki, Y., Rincon-Sanchez, A.R., Else, C., Saccomanno, S., 62351-6.
Benedetti, A., Ramirez, F., Rojkind, M., 2005. Early response of α2(I) collagen to Wu, J., Montaniel, K.R., Saleh, M.A., Xiao, L., Chen, W., Owens, G.K., Humphrey, J.D.,
acetaldehyde in human hepatic stellate cells is TGF-β independent. Hepatology 42, Majesky, M.W., Paik, D.T., Hatzopoulos, A.K., Madhur, M.S., Harrison, D.G., 2016.
343–352. http://dx.doi.org/10.1002/hep.20798. Origin of matrix-producing cells that contribute to aortic fibrosis in hypertension.
Swirski, F.K., Nahrendorf, M., 2013. Leukocyte behavior in atherosclerosis, myocardial Hypertension 67, 461–468. http://dx.doi.org/10.1161/HYPERTENSIONAHA.115.
infarction, and heart failure. Science 339, 161–166. http://dx.doi.org/10.1126/ 06123.
science.1230719. Wu, L., Luo, Z., Zheng, J., Yao, P., Yuan, Z., Lv, X., Zhao, J., Wang, M., 2018 Feb 7. IL-33
Szabo, G., Momen-Heravi, F., 2017. Extracellular vesicles in liver disease and potential as can promote the process of pulmonary fibrosis by inducing the imbalance between
biomarkers and therapeutic targets. Nat. Rev. Gastroenterol. Hepatol. 14, 455–466. MMP-9 and TIMP-1. Inflammation. http://dx.doi.org/10.1007/s10753-018-0742-6.
http://dx.doi.org/10.1038/nrgastro.2017.71. Wynn, T.A., 2004. Fibrotic disease and the TH1/TH2 paradigm. Nat. Rev. Immunol. 4,
Tacke, F., Trautwein, C., 2015. Mechanisms of liver fibrosis resolution. J. Hepatol. 63, 583–594. http://dx.doi.org/10.1038/nri1412.
1038–1039. http://dx.doi.org/10.1016/j.jhep.2015.03.039. Xu, J., Gonzalez, E.T., Iyer, S.S., Mac, V., Mora, A.L., Sutliff, R.L., Reed, A., Brigham, K.L.,
Tacke, F., 2017. Targeting hepatic macrophages to treat liver diseases. J. Hepatol. 66, Kelly, P., Rojas, M., 2009. Use of senescence-accelerated mouse model in bleomycin-
1300–1312. http://dx.doi.org/10.1016/j.jhep.2017.02.026. induced lung injury suggests that bone marrow-derived cells can alter the outcome of
Tacke, F., 2018. Cenicriviroc for the treatment of non-alcoholic steatohepatitis and liver lung injury in aged mice. J Gerontol A Biol Sci Med Sci. 64, 731–739. http://dx.doi.
fibrosis. Expet Opin. Invest. Drugs 27, 301–311. http://dx.doi.org/10.1080/ org/10.1093/gerona/glp040.
13543784.2018.1442436. Yang, A.H., Chen, J.Y., Lin, J.K., 2003. Myofibroblastic conversion of mesothelial cells.
Tan, Y., Ge, G., Pan, T., Wen, D., Gan, J., 2014. A pilot study of serum microRNAs panel as Kidney Int. 63, 1530–1539. http://dx.doi.org/10.1046/j.1523-1755.2003.00861.x.
potential biomarkers for diagnosis of nonalcoholic fatty liver disease. PLoS One 9 (8). Yang, J., Savvatis, K., Kang, J.S., Fan, P., Zhong, H., Schwartz, K., Barry, V., Mikels-
http://dx.doi.org/10.1371/journal.pone.0105192. e105192. Vigdal, A., Karpinski, S., Kornyeyev, D., Adamkewicz, J., Feng, X., Zhou, Q., Shang,
Tang, L., Tanaka, Y., Marumo, F., Sato, C., 1994. Phenotypic change in portal fibroblasts C., Kumar, P., Phan, D., Kasner, M., López, B., Diez, J., Wright, K.C., Kovacs, R.L.,
in biliary fibrosis. Liver 14, 76–82. http://dx.doi.org/10.1111/j.1600-0676.1994. Chen, P.S., Quertermous, T., Smith, V., Yao, L., Tschöpe, C., Chang, C.P., 2016.
tb00051.x. Targeting LOXL2 for cardiac interstitial fibrosis and heart failure treatment. Nat.
Tecklenborg, J., Clayton, D., Siebert, S., Coley, S.M., 2018. The role of the immune system Commun. 7, 13710. http://dx.doi.org/10.1038/ncomms13710.
in kidney disease. Clin. Exp. Immunol. http://dx.doi.org/10.1111/cei.13119. Feb 17. Yazdani, S., Bansal, R., Prakash, J., 2017. Drug targeting to myofibroblasts: implications
Thanigaimani, S., Lau, D.H., Agbaedeng, T., Elliott, A.D., Mahajan, R., Sanders, P., 2017. for fibrosis and cancer. Adv. Drug Deliv. Rev. 121, 101–116. http://dx.doi.org/10.
Molecular mechanisms of atrial fibrosis: implications for the clinic. Expert Rev. 1016/j.addr.2017.07.010.
Cardiovasc Ther. 15, 247–256. http://dx.doi.org/10.1080/14779072.2017.1299005. Ying, Q., Wu, G., 2017. Molecular mechanisms involved in podocyte EMT and con-
Tomankova, T., Kriegova, E., Liu, M., 2015. Chemokine receptors and their therapeutic comitant diabetic kidney diseases: an update. Ren. Fail. 39, 474–483. http://dx.doi.
opportunities in diseased lung: far beyond leukocyte trafficking. Am. J. Physiol. Lung org/10.1080/0886022X.2017.1313164.
Cell Mol. Physiol. 308, L603–L618. http://dx.doi.org/10.1152/ajplung.00203.2014. Zaffanello, M., Tardivo, S., Cataldi, L., Fanos, V., Biban, P., Malerba, G., 2011. Genetic
Troeger, J.S., Mederacke, I., Gwak, G.Y., Dapito, D.H., Mu, X., Hsu, C.C., Pradere, J.P., susceptibility to renal scar formation after urinary tract infection: a systematic review
Friedman, R.A., Schwabe, R.F., 2012. Deactivation of hepatic stellate cells during and meta-analysis of candidate gene polymorphisms. Pediatr. Nephrol. 26,
liver fibrosis resolution in mice. Gastroenterology 143, 1073–1083. http://dx.doi. 1017–1029. http://dx.doi.org/10.1007/s00467-010-1695-7.
org/10.1053/j.gastro.2012.06.036. e22. Zeisberg, M., Yang, C., Martino, M., Duncan, M.B., Rieder, F., Tanjore, H., Kalluri, R.,
Tsuchida, T., Friedman, S.L., 2017. Mechanisms of hepatic stellate cell activation. Nat. 2007. Fibroblasts derive from hepatocytes in liver fibrosis via epithelial to me-
Rev. Gastroenterol. Hepatol. 14, 397–411. http://dx.doi.org/10.1038/nrgastro. senchymal transition. J. Biol. Chem. 282, 23337–23347. http://dx.doi.org/10.1074/
2017.38. jbc.M700194200.
Tuchweber, B., Desmouliere, A., Bochaton-Piallat, M.L., Rubbia-Brandt, L., Gabbiani, G., Zhang, W., Zhou, X., Zhang, H., Yao, Q., Liu, Y., Dong, Z., 2016. Extracellular vesicles in
1996. Proliferation and phenotypic modulation of portal fibroblasts in the early diagnosis and therapy of kidney diseases. Am. J. Physiol. Ren. Physiol. 311,
stages of cholestatic fibrosis in the rat. Lab. Invest. 74, 265–278. F844–F851. http://dx.doi.org/10.1152/ajprenal.00429.2016.
Vanhaverbeke, M., Gal, D., Holvoet, P., 2017. Functional role of cardiovascular exosomes Zhu, J., Carver, W., 2012. Effects of interleukin-33 on cardiac fibroblast gene expression
in myocardial injury and atherosclerosis. Adv. Exp. Med. Biol. 998, 45–58. http://dx. and activity. Cytokine 58, 368–379. http://dx.doi.org/10.1016/j.cyto.2012.02.008.
doi.org/10.1007/978-981-10-4397-0_3. Zimmer, V., Lammert, F., 2011. Genetics in liver disease: new concepts. Curr. Opin.
Vilar-Gomez, E., Chalasani, N., 2018. Non-invasive assessment of non-alcoholic fatty liver Gastroenterol. 27, 231–239. http://dx.doi.org/10.1097/MOG.0b013e3283444862.
disease: clinical prediction rules and blood-based biomarkers. J. Hepatol. 68, Zlotnik, A., Yoshie, O., 2000. Chemokines: a new classification system and their role in
305–315. http://dx.doi.org/10.1016/j.jhep.2017.11.013. immunity. Immunity 12, 121–127. http://dx.doi.org/10.1016/S1074-7613(00)
Wang, H., Chen, X., Su, Y., Paueksakon, P., Hu, W., Zhang, M.Z., Harris, R.C., Blackwell, 80165-X.
T.S., Zent, R., Pozzi, A., 2015. p47(phox) contributes to albuminuria and kidney

14

You might also like