Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Emerging evidence suggests that altered components and posttranslational modifications of proteins in the extracellular matrix (ECM) may both initiate and drive disease progression. The ECM is a complex grid consisting of multiple proteins, most of which play a vital role in containing the essential information needed for maintenance of a sophisticated structure anchoring the cells and sustaining normal function of tissues. Therefore, the matrix itself may be considered as a paracrine/endocrine entity, with more complex functions than previously appreciated. The aims of this review are to 1) explore key structural and functional components of the ECM as exemplified by monogenetic disorders leading to severe pathologies, 2) discuss selected pathological posttranslational modifications of ECM proteins resulting in altered functional (signaling) properties from the original structural proteins, and 3) discuss how these findings support the novel concept that an increasing number of components of the ECM harbor signaling functions that can modulate fibrotic liver disease. The ECM entails functions in addition to anchoring cells and modulating their migratory behavior. Key ECM components and their posttranslational modifications often harbor multiple domains with different signaling potential, in particular when modified during inflammation or wound healing. This signaling by the ECM should be considered a paracrine/endocrine function, as it affects cell phenotype, function, fate, and finally tissue homeostasis. These properties should be exploited to establish novel biochemical markers and antifibrotic treatment strategies for liver fibrosis as well as other fibrotic diseases.

Free full text 


Logo of ajpgiLink to Publisher's site
Am J Physiol Gastrointest Liver Physiol. 2015 May 15; 308(10): G807–G830.
Published online 2015 Mar 12. https://doi.org/10.1152/ajpgi.00447.2014
PMCID: PMC4437019
PMID: 25767261

Novel insights into the function and dynamics of extracellular matrix in liver fibrosis

Abstract

Emerging evidence suggests that altered components and posttranslational modifications of proteins in the extracellular matrix (ECM) may both initiate and drive disease progression. The ECM is a complex grid consisting of multiple proteins, most of which play a vital role in containing the essential information needed for maintenance of a sophisticated structure anchoring the cells and sustaining normal function of tissues. Therefore, the matrix itself may be considered as a paracrine/endocrine entity, with more complex functions than previously appreciated. The aims of this review are to 1) explore key structural and functional components of the ECM as exemplified by monogenetic disorders leading to severe pathologies, 2) discuss selected pathological posttranslational modifications of ECM proteins resulting in altered functional (signaling) properties from the original structural proteins, and 3) discuss how these findings support the novel concept that an increasing number of components of the ECM harbor signaling functions that can modulate fibrotic liver disease. The ECM entails functions in addition to anchoring cells and modulating their migratory behavior. Key ECM components and their posttranslational modifications often harbor multiple domains with different signaling potential, in particular when modified during inflammation or wound healing. This signaling by the ECM should be considered a paracrine/endocrine function, as it affects cell phenotype, function, fate, and finally tissue homeostasis. These properties should be exploited to establish novel biochemical markers and antifibrotic treatment strategies for liver fibrosis as well as other fibrotic diseases.

Keywords: collagen, cytokine, extracellular fibrogenesis, integrin, laminin, matrix metalloproteinase, posttranslational modification, proteoglycan, endocrine

45% of all deaths in the developed world are associated with chronic fibroproliferative diseases (256, 378). Thus there is an increasing need to address fibroproliferative diseases because of their strong impact on the quality of life and health costs consequent to pain and organ failure, with an increased need for organ transplants despite dwindling availability, often followed by death. Moreover, their severity and perceived irreversibility in view of a current paucity of treatment options, coupled with a high prevalence in most and an orphan status in some fibrotic diseases, have just begun to attract biotechnology and big pharmaceutical companies to the field.

The common denominator of fibroproliferative diseases is a dysregulated tissue remodeling leading to the excessive and abnormal accumulation of extracellular matrix (ECM) components, thereby generating an ECM with different structural and signaling properties in the affected tissues (285, 287, 289, 378380). Fibrosis can affect almost any organ or tissue and is therefore associated with a wide variety of diseases and injuries (287). Figure 1 illustrates the major fibroproliferative diseases with a significant impact on human health (20, 287, 323, 378, 379).

An external file that holds a picture, illustration, etc.
Object name is zh30101568860001.jpg

Examples of fibroproliferative diseases in different organs. NASH, nonalcoholic steatohepatitis; HCV, hepatitis C virus; HBV, hepatitis B virus; AMD, age-related macular degeneration; IPF, idiopathic pulmonary fibrosis; COPD, chronic obstructive pulmonary disease; ARDS, acute respiratory distress syndrome; FSGS, focal segmental glomerulosclerosis. [From Karsdal et al. (162).]

Fibrotic tissue was for a long period of time considered an inactive scaffold, precluding regenerative potential for the affected organ. However, this perception cannot be upheld because fibrosis is neither static nor irreversible but the result of a continuous remodeling process and thereby susceptible to intervention (176, 337, 378). Presently, there are no approved treatments that specifically target the mechanism underlying fibrosis, but, especially in the liver, reversibility of even advanced fibrosis has been demonstrated upon treatment of the major underlying cause. Examples are effective antiviral therapy for chronic hepatitis B (22, 208) or the eradication of chronic hepatitis C with interferon-α-based and interferon-free regimens (94, 263, 264). The major future challenge in hepatology will be to halt fibrogenesis and reverse advanced fibrosis without tissue homeostasis or interfering with normal wound healing. Consequently, our increased understanding of the ECM, its dynamics, and the potential of fibrotic microenvironments to reverse holds promise for the development of highly specific and side effect-free antifibrotic therapies.

Traditionally, growth factors, cytokines, hormones, and certain other small molecules have only been considered as relevant mediators of inter-, para-, and intracellular communication and signaling. However, the ECM fulfils direct and indirect paracrine or even endocrine roles. In addition to maintaining the structure of tissues, the ECM has properties that directly signal to cells. Even conceptual exclusively structural proteins such as fibrillar collagens or proteoglycans are emerging as specific signaling molecules that affect cell behavior and phenotype via cellular ECM receptors. In addition, the ECM can bind multiple otherwise soluble proteins, growth factors, cytokines, chemokines, or enzymes, restricting or regulating their access to cells, apart from specifically attracting and modulating the cells that produce these factors. Moreover, specific proteolysis can generate biologically active fragments from the ECM, whereas their parent molecules are inactive. The ECM thus can control cell phenotype by functioning as a precursor bank of potent signaling fragments in addition to the direct effect on cell phenotype through ECM-cell interactions mediated by receptors such as integrins and/or certain proteoglycans (137, 138, 276).

The aims of this review are to 1) explore key structural and functional components of the ECM, in part exemplified by monogenetic disorders leading to severe pathologies; 2) describe select posttranslational modifications (PTMs) of ECM proteins that result in altered functional (signaling) properties of the original ECM component; 3) discuss the novel concept that an increasing number of components of the ECM harbor cryptic signaling functions that may be viewed as endocrine functions; and 4) highlight how this knowledge can be exploited to modulate fibrotic disease.

Methods

The PubMed database was searched using the following keywords: fibrosis, collagen, cytokine, growth factor, laminin, liver, matrix metalloproteinase (MMP), proteoglycan, posttranslational modification, ECM, and neoepitope. Each author further selected key publications according to his/her specific expertise.

Clinical Significance of the ECM in Liver Fibrosis

The common denominator of most chronic liver diseases is altered remodeling of the ECM initiated by inflammation, with both quantitative and qualitative alterations of its composition, thereby gradually disrupting normal function and finally leading to increased morbidity and mortality attributable to organ failure. Figure 2 highlights the main cellular and structural components of and differences between healthy and fibrotic tissue. The ECM is now considered a biologically active system that tends to perpetuate inflammation and fibrosis, rather than a passive consequence of chronic liver injury. This, together with accumulating evidence that fibrosis and even cirrhosis are reversible conditions, has changed the field and paved the path for novel optimism and interest in antifibrotic therapies (287). The novel insights into the pathophysiology and function of the ECM in chronic fibroproliferative diseases in general, and in liver fibrosis in particular, are already beginning to be translated into the clinic (26, 101, 287, 334). Early diagnosis for high-risk populations combined with targeted, individualized, and more rational interventions, including specific antifibrotic drugs, to prevent progressive disease or induce its regression and thus the morbidity, mortality, and excessive cost associated with symptomatic treatment of advanced stage fibrosis is urgently needed. Thus a central question to be addressed in this review is whether and how far our fairly advanced knowledge on the pathophysiology and molecular biology of the ECM can be translated into clinical practice.

An external file that holds a picture, illustration, etc.
Object name is zh30101568860002.jpg

Illustration depicting the changes that occur in the extracellular matrix (ECM) remodeling during the development of fibrosis. The ECM may be divided into the loose basement membrane and the compact interstitial matrix. The basement membrane consists mainly of type IV collagen, laminins, entactin, and proteoglycans and functions to anchor cells and connect to the interstitial ECM. The interstitial matrix has a different composition and consists mainly of fibrillar collagens, numerous noncollagenous glycoproteins, proteoglycans, and elastin. In healthy tissue, ECM remodeling is tightly controlled to ensure homeostasis; the synthesis of ECM proteins by fibroblasts has a relatively slow metabolic turnover, and the proteolytic activity is limited. During fibrogenesis, however, this process is disturbed. As a result of chronic wound healing, immune cells infiltrate the interstitial matrix, which helps drive the profibrotic response. Fibroblasts and fibroblast precursors like hepatic stellate cells differentiate into myofibroblasts, which deposit excessive levels of interstitial collagens. Initially, they may release enhanced levels of ECM-degrading matrix metalloproteinases (MMPs), whereas, at later stages, most MMPs are downregulated. This in turn results in a change in the composition of the ECM, where high levels of ECM degradation products and increased collagen cross-linking can be observed.

Fibrotic liver diseases are usually silent with limited clinical symptoms until development of cirrhosis with portal hypertension and complications like ascites, bleeding varices, or hepatocellular carcinoma (HCC). Furthermore, the heterogeneity within different subtypes of liver diseases is a separate challenge. Fewer than 20% of chronic alcohol abusers develop fibrosis, and of these only few will progress to cirrhosis (293, 319). Similarly, ~50% of patients with hepatitis C virus infection will never develop significant fibrosis (169), and only 10–20% of patients with nonalcoholic fatty liver disease will develop nonalcoholic steatohepatitis (NASH); of these, 10–20% will progress toward cirrhosis within 5–15 yr (290). Although the individual predictors for fast fibrosis progression are diverse, including genetic predisposition, increased age at onset of disease, and especially (hepatic) comorbidities (second hits), there is still a lack of sensitive noninvasive biomarkers to predict the individual risk of fibrosis progression (288). However, the population at risk for chronic liver diseases is enormous and poses a global health challenge. Thus the prevalence of obesity alone is 10–30% in most Western and developing countries, as is the consumption of harmful quantities of alcohol (1, 61, 232, 393). On the other hand, with the advent of potent antivirals, the global epidemic of chronic viral hepatitis will likely be harnessed in the near future causally (180, 233). However, even in chronic viral hepatitis, antifibrotics will be useful in those patients with advanced disease in whom fast regression is desirable to further decrease the risk of hepatic decompensation and HCC.

Despite the emergence of more effective antifibrotic agents, there is a lack of agents that specifically derive from the ECM (287), the fibrotic structure itself. The dynamics of the ECM likely harbor important diagnostic information but, more importantly, also clues and pathway for a targeted intervention. Thus biomarkers of ECM remodeling are the most likely candidates to predict fibrosis progression or regression or to permit noninvasive monitoring of antifibrotic interventions (162). During tissue remodeling, proteases release small protein fragments into the circulation (52, 162, 259, 288) (Fig. 2) that may serve as biomarkers of the fibrotic process. Furthermore, the dynamic nature of ECM measures vs. current static measures obtained by imaging or histology should synergistically improve diagnostic and prognostic information on fibrotic liver diseases (52, 162, 259, 288).

Fibrosis and the ECM

The matrix composition itself is particularly important for the regulation of fibrosis. Because the ECM composition regulates the behavior and phenotype of cells housed in the ECM, any ECM remodeling will in turn influence adjacent cells and modify their behavior/phenotype. Consequently, abnormal ECM remodeling may be a prerequisite for fibrogenesis and/or fibrolysis.

Common to all fibrotic disorders is the characteristic alterations in ECM remodeling, which results in the excessive accumulation of ECM components in a given organ that can ultimately lead to organ malfunction (117, 151, 352). Fibrosis is also a dynamic condition with accelerated ECM turnover in which both tissue formation and tissue degradation are highly upregulated (15, 16, 297, 350), resulting in pathological collagen deposition and altered MMP expression profiles (353). Moreover, modifications to amino acids or proteolytic cleavage at specific locations by specific PTMs result in both immunologically and functionally different proteins (161).

Fibrogenesis and fibrolysis are driven by many cell types and molecular events. Fibrogenesis is intimately linked to wound healing, serving to prevent tissues from disassembly during inflammation, apoptosis, necrosis, and release of lytic enzymes. In the liver, the major downstream effectors of fibrosis are activated stellate cells and (portal) fibroblasts with a phenotype of wound myofibroblasts (175), which produce excessive amounts of ECM molecules, such as the prominent collagens, mainly the abundant fibrillar type I and III collagens (131), which are a hallmark of all fibrotic diseases increasing up to tenfold in advanced fibrosis of organs, such as lungs and liver (205, 370). Type IV collagen is also subject to extensive remodeling during fibrosis, and its quantity has been found to increase more than 10-fold during liver fibrogenesis (116, 289). Variants of type IV collagen together with isoforms of laminin, nidogen, and perlecan are the main components of all basement membranes and form a sheet-like scaffold at the basal site of epithelia and endothelia and around interstitial cells (135, 289) that maintain viability and differentiation of these cells (157, 231). Other functions of this network are the provision of interaction sites for other ECM components, inflammatory cells, chemokines, cytokines, and important functions in cellular signaling (394). The collagen formation observed during fibrosis is also accompanied by an increase in type IV collagen degradation and unfavorable remodeling of the basement membranes, resulting from an increased expression of proteases in the affected tissue (129, 194, 351). This favors myofibroblast activation and a net increase in the deposition of, for example, fibrillar collagens by myofibroblasts (379). This is in part accompanied by a deregulated MMP activity toward basement membranes and increased expression of tissue inhibitors of matrix metalloproteinases (TIMPs) by several cell types in the fibrotic liver, especially TIMP-1, which blocks local interstitial collagen degradation by, for example, MMP-1, -8, and -13 and further promotes myofibroblast activation directly (31, 129, 309). Importantly, collagen properties and quality are altered in fibrosis by protein modifications, including increased cross-linking, that result in enhanced tissue stiffness, which contributes to fibroblast activation and compromised hepatocyte function. Although there presently is no consensus on the most important proteins of the ECM that should be specifically addressed in fibrosis, type I and III collagens are the most abundant ones, followed by type IV, V, and VI collagens, nidogen, laminin, fibronectin, biglycan, mimican, versican, decorin, lumican, and elastin, to name a few (117119, 195, 287, 289), all clearly being altered in quantity and quality in fibrosis. Although myofibroblasts may be the central ECM-producing cells, macrophages are emerging as important upstream regulators of hepatic fibrogenesis and direct effectors of fibrosis resolution, controlled by specific subtypes, such as M1 and variant M2 macrophage subsets (20, 37, 84, 91, 223, 265, 266, 284, 332, 349, 379381). Specific functional macrophage subtypes that are highly dependent on the ECM and cellular environment and soluble mediators in the host tissue can also phagocytose cellular debris, which removes potential proinflammatory signals and can cause increased MMP expression and enhanced matrix degradation (260). In a simplified view, M1 macrophages are rather proinflammatory and antifibrotic, in contrast to (subtypes of) M2 macrophages, which can assume profibrotic properties.

The Importance of the ECM for Tissue Function: The ECM Is Controlling Cell Phenotype

The first evidence of a central functional role of the ECM in controlling cellular behavior was obtained when a malignant cellular genotype and phenotype could be repressed by a normal mouse embryonic ECM (81, 214). Generally the ECM is a 3-D scaffold that supports or encapsulates sessile or migrating cells and defines their microenvironment (11). It consists of a meshwork of proteins and nonprotein components (glycosaminoglycans) to which soluble factors, such as growth factors and cytokines, can bind, which regulates accessibility of common nutrients. The importance and the role of ECM in cell phenotype, tissue-specific differentiation is exemplified by the fact that cells grown as 2-D monolayers on top of either a plastic substrate or a glass coverslip, with or without ECM ligand, fail to assemble into the same tissue-like structures as those growing in or on top of the normal 3-D ECM. Cells grown on plastic or glass are also unable to express differentiated proteins upon stimulation (247) or to respond to growth factors or protease inhibitors in the same way as cells growing in a 3-D setting (163). These phenotypic disparities can be explained, at least in part, by ECM-derived signals that living tissues in 3-D emit and that are transmitted into the cell via ECM receptors such as integrins and via receptors for ECM-bound proteins such as growth factors. This temporospatial 3-D signaling is absent or altered in 2-D substrata. This has been convincingly shown for both epithelial and interstitial cells (122). The architecture of the interstitial matrix in vivo also differs substantially from that offered to or found typically in cells cultured on plastic (163). As an example, osteoblasts grown on plastic in 2-D do not rely on MMPs for survival, whereas osteoblasts embedded in an interstitial 3-D matrix containing type I collagen are critically dependent on MMP activation of latent transforming growth factor (TGF)-β for their survival (163). Moreover, the orientation and function of cells and collagen fibers are lost when cells are grown in 2-D compared with 3-D, which critically regulate cell and tissue behavior (238, 251, 252). Taken together, in vitro models need to replicate the naturally occurring 3-D environment, encompassing a sufficient physiological and pathophysiological variety of ECM components. Thus the effect of 2-D vs. 3-D ECM environments on central biological features of fibroblasts and myofibroblasts, e.g., contraction, migration, proliferation, ECM synthesis, and degradation, is remarkable, usually with a much higher fibrogenic activation under 2-D conditions (44, 71, 76, 149).

The Relation of Structural Proteins to Pathologies

Important information on the functional role of structural components of the ECM has been obtained from mutations in ECM genes that lead to pathologies. Table 1 contains a summary of key structural proteins and their known mutations leading to matrix and tissue failure. With the caveat that some of these components fulfil key functions only in development but may be dispensable later in life, these disease phenotypes provide pivotal information on ECM molecules important for tissue function and thus give insight into their function and dysfunction in the pathology of nongenomic disorders.

Table 1.

The relation of matrix components to connective tissue diseases

ProteinDiseaseAnimal ModelsReference
Type I collagenOI, Ehlers-Danlos syndrome type VIIOI model (303)(328, 368)
Type II collagenSeveral chondrodysplasias, osteoarthritisCIA model (39, 336)(4, 5, 95, 246)
Type III collagenEhlers-Danlos syndrome type IV, aortic aneurysmsKO/haploinsufficiency mice model (201, 312)(183, 338)
Type IV collagenKidney fibrosis, Alport syndromeCanine and murine models (166)(18, 331, 347, 348)
Type V collagenEhlers-Danlos syndrome type I and II(269, 369)
Type VI collagenBethlem myopathy, Ullrich congenital muscular dystrophyKO murine models (32)(187)
Type VII collagenEpidermolysis bullosa dystrophicaDEB mice model (128)(74)
Type IX collagenMEDKO mice model (6, 132)(70)
Type X collagenSMCD, Japanese-type SMDKO and transgenic mice models (145)(206, 373)
Type XV collagenCardiac and muscle phenotypesKO mice (87)(348)
Type XVII collagenGrowth retardationKO and nude mice (113, 235)(348)
Type XVIII collagenRenal filtration defects, Knobloch syndromeKO and loss-of function mice (103, 344)(302, 348)
ElastinLung, skin, and arterial defects, SVAS, WBS, CLVarious mice models (245)(147, 170, 213)
LamininAlport syndromeMurine and canine models (167)(18)
BiglycanCardiovascular disease, osteoporosisKO mice (392)(56, 127, 387)
Biglycan/DecorinOsteopenia, skin fragilityDouble KO mice (392)(392)
Biglycan/FibromodulinOsteoarthritisKO mice (357)(6a)
PerlecanMultiple developmental defects, myotonia, Schwartz-Jampel syndromeTransgenic and point mutation mice (324, 365)(348)
Nidogen 1 and 2Lung and kidney developmentSingle KO mice (224, 296)(347)
FibromodulinOsteoarthritisKO mice (326)(108)
Lumican/FibromodulinJoint laxity, impaired tendon integrityDouble KO mice (89)(150, 326)
LumicanReduced corneal transparency, skin fragilityKO mice, knockdown zebrafish (89, 390)(54)
DecorinIntestinal tumor, skin fragility, Ehlers-Danlos syndrome-likeKO mice (75)(28, 68, 75)
MimecanColorectal cancer early formation(364)
FibrillinMarfan syndromeMurine and bovine model (27, 199)(126)
COMPPSACH, MEDTransgenic mice (172)(41, 250) (146)
Matrillin-3MEDKO mice (346)(42)
FibronectinGlomerulopathy (proteinuria, microscopic hematuria, hypertension, renal failure)Knockdown mice (182)(295, 310)
Tenascin CCardiovascular diseases, liver fibrosisKO mice (186)(59, 60)

[Modified and extended from Karsdal et al. (164).] OI, osteogenesis imperfecta; CIA, collagen-induced arthritis; KO, knockout; DEB, dystrophic epidermolysis bullosa; MED, multiple epiphyseal dysplasia; SMCD, Schmid-type metaphyseal chondrodysplasia; SMD, spondometaphyseal dysplasia; SVAS, supravascular aortic stenosis; WBS, William-Beuren syndrome; CL, cutis laxa; COMP, cartilage oligomeric matrix protein; PSACH, pseudoachondroplasia.

Mutations within these structural proteins clearly suggest that these proteins have capacities that are important for the maintenance of a healthy ECM phenotype.

The ECM as Regulator of Cytokine and Growth Factor Activities

In addition to the direct effect of ECM structural molecules on cell phenotype and tissue function, the ECM also serves as a storage site for multiple otherwise soluble cytokines and growth factors (306). In particular, the small leucine-rich proteoglycans (SLRPs), which act as matricellular proteins, are active components of the ECM with a specific role in direct or indirect modulation of the cell-matrix crosstalk. These molecules are modulators of growth factor and cytokine functions, such as TGF-β1 (considered the most potent profibrotic cytokine), tumor necrosis factor (TNF)-α, Wnt-1 induced secreted protein 1, and bone morphogenic proteins (BMPs) (211), but they possess other signaling properties, both as whole proteins and as protein fragments. Specifically, decorin and biglycan are antifibrotic molecules, which by binding active TGF-β1 can interfere with its signaling and neutralize its activity (106). The biological activity of TGF-β1 is attenuated by binding to the core proteins of the SLRPs decorin and biglycan (280282). The SLRPs are also potent antiapoptotic molecules of tubular epithelial and endothelial cells, acting through binding to the insulin-like growth factor (IGF) type I receptor (211, 280). Notably, biglycan degradation was recently shown to be highly correlated to fibrosis in carbon tetrachloride (CCl4) and bile duct ligation models of liver fibrosis (106). The soluble factors are usually bound to specific ECM components via low-affinity noncovalent interactions, which create stable concentration gradients around the ECM-embedded cells that produce these cytokines and growth factors, guiding, for example, inflammatory cells but also (myo)fibroblasts toward the target site of inflammation and often excess ECM production. Moreover, ECM binding serves as acellular storage sites, from which these factors will be released during tissue injury, when inflammatory cells degrade the ECM, promoting tissue regeneration and remodeling. Many of the heparan sulfate (HS) (proteoglycan) binding growth factors promote angiogenesis [e.g., fibroblast growth factor (FGF)-1 and vascular endothelial growth factor (VEGF)] or epithelial cell proliferation and survival [e.g., epidermal growth factor (EGF), hepatocyte growth factor (HGF), and keratinocyte growth factor]; others that bind to the abundant collagens regulate fibroblast or immune cell activation [e.g., platelet-derived growth factor (PDGF)-B, oncostatin-M, interleukin-2, and HGF]. Most of these interactions have been studied in detail (289, 291, 315317), which is schematically drawn in Fig. 3 and listed in Table 2 and highlights that the ECM is a key specific storage facility for potent signaling molecules. Other important examples of molecules sequestered in the ECM are 1) the proinflammatory cytokine TNF-like weak inducer of apoptosis (TWEAK) that induces proliferation of hepatic progenitor cells (HPCs) directly through its receptor fibroblast growth factor-inducible 14, known to be overexpressed in chronic liver diseases (148, 243, 333). Lastly, the hedgehog and Wnt/β-catenin pathways together with Notch signaling have been documented to be important in HPC activation and differentiation (37, 154, 307, 320), as HPCs maintain viability via autocrine and/or paracrine Hedgehog signaling (307).

An external file that holds a picture, illustration, etc.
Object name is zh30101568860003.jpg

Binding of certain growth factors and cytokines to the extracellular matrix. The figure highlights prominent interactions. Shown is a selection of relevant ECM binding factors (also discussed in the text) and their association with either heparan sulfate or collagen and their target cells. The liberation of ECM-stored biologically active growth factors and cytokines can either trigger (inflammatory) cells to further degrade ECM or promote excess ECM deposition by (myo)fibroblasts (as is the case with decorin/biglycan-bound transforming growth factor, TGF-β1). EGF, epidermal growth factor; FGF, fibroblast growth factor; HGF, hepatocyte growth factor; IL-2, interleukin-2; KGF, keratinocyte growth factor; OsM, oncostatin-M; PDGF, platelet-derived growth factor; VEGF, vascular endothelial growth factor.

Table 2.

Key factors and their downstream effect on signaling molecules and cells

DownstreamEffectReference
Growth Factor SignalingPDGFsRas-MAPK, PI3K-Akt/PKB, PKCHSC activation(33, 90, 168, 242, 375)
TGFsJNK, NF-κB, CTGF, SmadHSC proliferation, fibrogenic/inhibitory(53, 191, 200, 287)
EGFsSTATs, EGFR, ERK1/2HSC proliferation, polypeptide mitogen(107, 159, 191, 270)
VEGFsMAPK, AktHSC activation, hepatic angiogenesis(67, 191, 397)
HGFsJAK/STAT, MAPK, c-MetHepatocyte mitogen, antifibrotic(105, 174, 382)
IGFsMAPK, PI3K, ERKAntifibrotic, HSC proliferation(49, 313, 325)
TNF-αNF-κB, JNK, ERKHepatocyte proliferation, inflammation, fibrogenic(262, 330)
ChemokinesCXCL1, CXCL9, CXCL10CXCR2, CXCR3, TGF-βInflammation, antifibrotic, chemoattractant(130, 191, 366, 395)
MCP1-3CCR2, p47HSC activation, NK cell activation, inflammation, chemoattractant(202, 301)
CCL5 (RANTES)CCR5, PI3K/AktNK cell activation, HSC migration and proliferation, inflammation, chemoattractant(173, 191, 202, 300)
Innate immune interactionsTLR2, TLR4TNF-α, NF-κB, JNK, BAMBIChemotactic/inflammatory(101, 191, 287)
CD40JNK, NF-κBSecretion of MCP-1 and IL-8, inflammation, chemoattractant(100, 101, 292)
IL-6JAK/STAT, ERK1/2, Akt, NF-κB, MAPKRegulation of inflammatory response, antifibrotic(229, 270)
IL-8 (CXCL8)CXCR1, CXCR2.Chemoattractant, inflammation(398)
IL-10JAK/STATAntifibrotic(22, 105, 396)
Adipokine pathwaysLeptinOB-R, JAK/STAT, TGF-βFibrogenic, HSC activation(191, 360)
AdiponectinAMPKAntifibrotic(46, 92, 158)
DevelopmentalHedgehogPtc, Smo, Gli familyFibrogenic, HSC activation and proliferation(255, 287, 389)
NotchNICD, RBP-JκFibrogenic, HPC differentiation, epithelial-mesenchymal transition(57, 217)
WntWnt/β-catenin, Wnt/Ca2+Fibrogenic, HSC activation(58, 197)

PDGF, platelet-derived growth factor; MAPK, mitogen-activated protein kinase; PI3K, phosphoinositide 3-kinase; PKB/C, protein kinase B/C; HSC, hepatic stellate cell; TGF, transforming growth factor; JNK, c-Jun N-terminal kinase; NF, nuclear factor; CTGF, connective tissue growth factor; EGF(R), epidermal growth factor (receptor); STAT, signal transducer and activator of transcription; EGFR, epidermal growth factor receptor; ERK, extracellular signal-regulated kinase; VEGF, vascular endothelial growth factor; HGF, hepatocyte growth factor; JAK, Janus kinase; IGF, insulin-like growth factor; TNF, tumor necrosis factor; CXCL/R, CXC chemokine ligand/receptor; MCP, monocyte chemoattractant protein; CCL/R, CC chemokine ligand/receptor; TLR, Toll-like receptor; BAMBI, bone morphogenic protein and activin membrane-bound inhibitor homolog; IL, interleukin; OB-R, OB/leptin receptor; AMPK, adenosine monophosphate-activated protein kinase; Ptc, Patched; Smo, Smoothened; NICD, Notch intracellular domain; RBP, recombinant signal-binding protein.

Tissue Turnover Generates Posttranslational Modifications with Signaling Function Affecting Cell Phenotype

To maintain healthy tissue, the ECM must regenerate itself by normal remodeling, in which aged or damaged proteins are broken down in a specific sequence of proteolytic events and replaced by new proteins. However, during pathological conditions, such as fibrosis and inflammation, the delicate repair-response balance is disturbed (142, 289). The constituents of the “aged” ECM are degraded, which in these disease states results in an array of protein fragments and altered proteins, some of which have documented signaling potential (216, 230, 268, 305, 358, 388), which will be discussed in depth on an individual protein level in the following sections. The original proteins of the ECM are replaced by different constituents, and, consequently, the composition and quality of the matrix are altered. This transformation from a healthy tissue to a pathological one may cause the matrix to become stiff, which has been shown to enhance tumor cell migration, myofibroblast activation, and collagen deposition (19, 21, 29, 66, 143, 171, 219, 253, 271), thereby linking the actual matrix quality to disease progression and changing cell phenotypes. Figure 4 illustrates the steps of abnormal ECM remodeling in fibrosis. Healthy ECM consists of a network of fibers organized in a highly ordered fashion, with binding of key growth factors and signaling molecules at specific interaction sites in the network. During high matrix turnover, the ECM is degraded, leaving fragments of the ECM in the matrix and releasing other fragments into the circulation. Multiple enzymes are released into the matrix by both resident and invading cells, modulating the ECM and generating an altered microenvironment. Consequently, the altered cell-ECM interactions result in altered cellular phenotypes. Different steps during ECM remodeling and fibrogenesis are likely characterized by unique patterns of protein formation, deposition, and degradation, which generates unique protein fingerprints, part of which should be released into the circulation to be exploited as stage-specific serum markers of liver fibrogenesis and/or fibrolysis (259, 288).

An external file that holds a picture, illustration, etc.
Object name is zh30101568860004.jpg

Schematic representation of the high rate of extracellular matrix remodeling in fibrosis. Healthy ECM consists of a network of fibers organized in a highly ordered fashion. During high matrix turnover, the ECM is degraded, leaving fragments of the ECM in the matrix and releasing other fragments into the circulation, while an accumulation of both new and already existing proteins occur, macroscopically described as fibrosis. ECM remodeling is a delicate equilibrium and a prerequisite for maintenance of a healthy tissue, in which senescent proteins are continuously degraded and replaced by new ones. This delicate balance is disturbed in fibrotic diseases, resulting in an increased ECM turnover (both formation and degradation). Thus a subset of pathological proteases is overexpressed in the affected tissue, resulting in the release of protease-specific fragments of signature proteins of the fibrotic ECM, referred to as protein fingerprints or neoepitopes. These fragments may be used as early diagnostic or prognostic serological markers of tissue degradation and in part formation. PTMs, posttranslational modifications. [From Karsdal et al. (164).]

One group of modifications affecting protein quality is PTMs. PTMs are non-DNA-encoded modifications and are a consequence of tissue physiology and pathophysiology (63, 64). Examples of physiological PTMs are isomerization of aspartate (seen in tissue aging), citrullination of arginine, nitrolysation of cysteine (occurring during inflammation), protease degradation at cleavage hotspots (observed in fibrosis and inflammation), glycosylation (in glycemia, type II diabetes), and the fibrosis-specific modifications made by polysialic acid that modulate the profibrotic ductular reaction in liver injury (62, 64, 339). Each of these modifications may change the function and signaling of the modified protein. Several lines of independent evidence suggest that PTMs to specific proteins contribute to abnormal cellular proliferation, adhesion, and morphology (185) and may cause many of the differences in fibrotic compared with normal tissue (36, 125, 185, 209, 279, 321). These specific PTMs made to specific proteins, in particular to structural proteins including collagens, have been shown to be an integrated part of disease progression, exemplified by citrullination of type II collagen in rheumatoid arthritis, cross-linking of collagens in fibrosis and acetylation, citrullination, isomerization, and phosphorylation of myelin basic protein in systemic lupus or multiple sclerosis (63, 164). Similar PTMs made to other collagens or key structural proteins could be important determinants of fibrosis progression, and evidence has recently been obtained for this role, as exemplified by modifications to type XVIII collagen, as discussed later.

There is a growing list of ECM molecules with documented effect on tissue function. This effect may be referred to as a newly discovered paracrine function of ECM proteins. A nonexhaustive list of ECM proteins (as this list continues to grow) and their exerted effects is shown in Table 3. These examples highlight those ECM proteins that serve as paracrine signaling molecules, often revealed during pathological processes that, in addition to cytokines, growth factors, and hormones, become essential players in tissue homeostasis, apart from their roles to anchor cells and transmit positional information and differentiation signals. Notably, some proteins do not change the cellular phenotypes in their native conformation, whereas, subsequent to a specific PTM, a highly potent and novel function of the same protein is revealed.

Table 3.

The matricellular effects of ECM components

Protein or PTMCellular PhenotypeResponsible ReceptorReference
Elastin-derived peptidesChemotaxis of monocytes, fibroblasts, endothelial cellsElastin-binding protein in complex with protective protein/cathepsin A and neuraminidase-1(83)
Proliferation of fibroblasts and smooth muscle cells
Protease release from fibroblasts and leukocytes
ThrombospondinInhibition of angiogenesisCD36 and CD47(7, 112, 363)
Type I collagenFibroblast migrationDDR2(277)
Acetylated Pro-Gly-Pro (acPGP), fragment of type I collagenNeutrophil chemotaxisCXCR1 and CXCR2(254, 367)
Arresten, canstatin and tumstatin, fragments of type IV collagenInhibition of angiogenesis, tumor growth, and endothelial cell proliferation and migrationVarious integrins(124, 222)
Induction of apoptosis
Endostatin, fragment of collagen type XVIIIInhibition of endothelial proliferation, angiogenesis, and tumor growthGlypicans, nucleolin(239)
Induction of endothelial cell apoptosis(78, 165, 305
RGD motif, present in collagens, laminin and fibronectinCell adhesion, angiogenesis, apoptosisVarious integrins(45, 278)
FibromodulinProliferation, migration, and chemotaxis of HSCsUnknown(221)
Laminin-332, elastase-generated fragment of γ2Neutrophil chemotaxisUnknown(227)
SIKVAV and ASKVKV (sequences in linker regions between coiled-coil and globular domains of laminin α1 and α5 chains)Neutrophil and macrophage chemotaxisUnknown receptors. SIKVAV interacts with integrins α1, α6, and β1 in salivary gland carcinoma cell line(3, 98)
LamininChemotactic migration of malignant cells toward laminin67LR (LamR)(79, 318)
LumicanRegulation of inflammation and innate immunityCD14, Fas ligand, CXCL1(50, 104, 355, 377)
Apoptosis induction
Fas
BiglycanRegulation of inflammation and innate immunity and effect on adhesion and migrationTLR2, TLR4, P2X4/P2X7, selectin L/CD44, C1q(12, 55, 121, 144, 178, 181, 218, 234, 281, 311, 341, 342)
Cytokine modulation (PDGF, TGF-β, TNF-α, WISP-1, BMP-4)
RhoA, Rac1
DecorinSignal transductionLRP-1, c-MET(40, 109)
Cytokine modulation (PDGF, TGF-β, TNF-α, VWF, WISP-1)TGF-β, C1q(77, 121, 181, 234, 341)
Regulation of inflammation and innate immunityIGF-IR(65, 311)
Antiapoptotic effectsEGF-R, VEGF-R2(282)
Antioncogenic effectsIGF-IR, integrin α2β1, RhoA, Rac1(139, 298)
Adhesion and migration effects(96, 342)
FibronectinA 40-kDa fragment prevents PDL cell spreading, thereby inducing anoikisVarious integrins(17, 72)
The 29- and 50-kDa amino terminal fragments mediate release of proteoglycan from articular cartilage by RGD-independent mechanisms
Fn fragments can induce fibroblast gene expression of MMPs or can act as proteinases themselves
Tenascin-CFragments are highly upregulated in arthritic cartilage, where they mediate cartilage degradation by the induction of aggrecanase activityαv-integrins and unknown(314)

[Extended and modified from Karsdal et al. (164).] ECM, extracellular matrix; PTM, posttranslational modification; DDR, discoidin domain receptor; WISP, Wnt1-inducible signaling pathway protein; BMP, bone morphogenic protein; VWF, Von Willebrand factor; LRP, lipoprotein receptor-related protein; PDL, periodontal ligament; MMP, matrix metalloproteinase. RGD, arginine-glycine-aspartic acid; Fn, fibroblast growth factor-inducible.

Posttranslationally Modified Proteins Affect Cell Phenotype: Examples of a Paracrine Function of the ECM in Relation to Fibrosis

Fragments of type XVIII collagen: endostatin.

In line with the PTMs of ECM molecules being able to control cell phenotype (388), a peptide derived from endostatin by MMP activity (239) was shown to ameliorate organ fibrosis. Here, peptide E4 (endostatin), derived from the noncollagenous COOH terminus of type XVIII collagen, which is present in the liver sinusoidal and basement membrane (286), prevented TGF-β1-induced dermal fibrosis and bleomycin-induced dermal and pulmonary fibrosis in mouse models and ex vivo in human skin. In addition, E4 significantly reduced existing fibrosis in these preclinical models. E4 amelioration of fibrosis was accompanied by reduced cell apoptosis and lower levels of lysyl oxidase-like 2 (LOXL2), a disease-related member of the LOX enzyme family that cross-links collagen providing resistance to proteolysis. Similar findings were observed in the lung bleomycin model where E4 inhibited the TGF-β and TNF-α pathways (358). Along the same line of thinking is restin, which is a close homolog to E4 derived from type XV collagen, another basement membrane collagen with predominant localization in the portal ECM of the liver (123). Restin has inhibitory effects on endothelial cell migration but not on their proliferation (268), whereas tumor suppression by type XV collagen is independent of the restin domain (225). Vastatin, the noncollagenous COOH-terminal fragment of type VIII collagen, inhibits endothelial cell proliferation and induces apoptosis in a bovine aortic endothelial cells (386). These protein fragments are all derived from the processing of collagen and are consequently PTMs of collagens involved in organ fibrosis. They add to the evidence that ECM molecules have signaling properties, which can be considered an endocrine function.

Fragments of type IV collagen, fibronectin, and plasminogen (arresten, canstatin, tumstatin, angostatin, and anastellin).

Angiogenesis generally precedes fibrosis, and consequently these early events are of pivotal importance for progression of fibrosis. Several molecules of the ECM modified by proteolysis affect angiogenesis by ECM-cell interactions through the fibrosis structure itself. The sprouting of new vessels from preexisting vasculature may enhance the deregulated and amplified ECM composition, support fibroblast proliferation, and constrain normal tissue repair. Numerous endogenous inhibitors of angiogenesis are derived from proteolysis of the ECM and vascular basement membrane. Notably, angiostatin, a fragment derived from blood coagulation factor plasminogen, is one of the most potent antagonists of angiogenesis and is shown to inhibit liver fibrosis in mice (356). In a similar fashion, anastellin, a peptide derived from the first type III repeat of fibronectin, is an example of a matrix-derived inhibitor of angiogenesis. It prevents angiogenesis and growth of human tumors when injected into mice (391). Anastellin binds to full-length fibronectin, augmenting formation of high polymerized fibronectin multimers termed superfibronectin (220). Additionally, it remodels already assembled fibronectin matrix, affecting apoptosis, cellular differentiation, and cell cycle progression, being thus critical for cell growth and survival. Other matrix-derived fragments are arresten, canstatin, and tumstatin, all derived from MMP cleavage of the NC1 domain of type IV collagen, α-1, -2, and -3 chain, respectively. These peptides are inhibitors of angiogenesis, tumor growth, and endothelial cell proliferation and migration (124, 222). Despite the fact that numerous matrix-derived fragments are angiogenesis inhibitors, they inhibit distinct aspects of angiogenesis. Endostatin and anastellin fragment of type XVIII collagen and fibronectin, respectively, both inhibit endothelial cell migration in response to VEGF, but only anastellin completely inhibited dermal endothelial proliferation (230).

LOX: cross-linking of the ECM.

Modifications such as cross-linking of ECM proteins by LOX family members have gained increasing attention in fibrosis because of their role in generating tissue stiffness, which promotes fibrogenesis (21). LOX is highly overexpressed in the local fibrotic microenvironment, in particular by myofibroblasts (374). Focus has been directed especially to LOXL2, which is the most expressed of the nine members of the LOX family (141), predominantly at advanced disease stages (69, 343). LOXL2 activity and expression correlate with the derangement of the ECM microenvironment of tissues that are associated with cancer and fibrosis (43). Apart from direct signaling roles in cancer proliferation and dedifferentiation, the increase in cross-links contributes to the stability of collagen accumulation in fibrosis (34, 226, 362). Although there is a high constitutive expression of LOX in most tissues, LOXL2 is specifically upregulated in fibrosis and cancer and is tightly linked to a worsening of tumor grade and fibrosis stage (21). The validity of targeting this enzyme and its cross-linking activity to inhibit fibrosis and cancer has been tested using a therapeutic monoclonal antibody (21), and a clinical study in patients with fibrotic NASH and primary sclerosing cholangitis is ongoing.

Transglutaminase-mediated cross-linking of the ECM.

In addition to LOX, there is a well-established role for transglutaminases (TGs) in the biochemical modification of ECM proteins (156). TGs constitute a family of at least eight related proteins with well-characterized transamidation activities forming largely irreversible Nε(γ-glutamyl) lysine isopeptide bonds between a glutamine residue on one and a lysine (histidine) residue on another protein (88, 204). TG2 is the most ubiquitously expressed enzyme of this family, also in liver, but is largely retained intracellularly in an inactive state. Its secretion is carefully regulated, in particular by myofibroblasts and endothelial cells, increasing dramatically following cellular damage, where it becomes activated by the high extracellular calcium concentrations (193, 308). Extracellular TG2-dependent matrix cross-linking has been closely linked with the pathogenesis of fibrotic disease of kidneys, liver, and lungs (152, 215, 244) and also with tumor progression, cardiovascular diseases, and intestinal inflammation, where it plays a central role in celiac disease (38, 80, 207). TG2 has been shown to cross-link certain ECM proteins, like type in the NH2-terminal portion of III procollagen, fibronectin, and elastin, conferring increased stability, rigidity, and resistance to degradation of the ECM, similar to the effects of LOXL2 (88, 354). However, a recent study comparing liver fibrosis progression and regression with the extent of collagen cross-linking in several mouse models demonstrated that TG2 does not significantly affect fibrosis in mice deleted of TG2 vs. wild-type controls (261). This may be different when TG2 activity is blocked in wild-type mice because the irreversible TG2 inhibitor NTU281 ameliorated the development of fibrosis and kidney failure in a model of diabetic nephropathy (133, 134). Given that this inhibitor acts extracellularly, its antifibrotic effects are likely attributable to inhibition of TG2 cross-linking activity. However, the extracellular functions of TG2 are more complex, including activation of TGF-β1 from its inactive precursor (134) and nonenzymatic interactions between ECM proteins such as fibronectin and cell surface growth factor receptors and integrins (80).

Type VI collagen as regulator of fibrogenesis.

Type VI collagen forms microfilaments that traverse interstitial connective tissues. These microfibrils are among the first ECM structures to be degraded upon tissue injury, resulting in larger fragments that, via engagement of integrins and perhaps NG2 proteoglycan, prevent apoptosis and induce fibrogenic activation of surrounding (myo)fibroblasts via activation of proproliferative mitogen-activated kinases and PDGF receptor-α (9, 273, 274). This mechanism likely serves to trigger an immediate wound-healing response, with type VI collagen as a sensor for ECM destruction. Modulation of type VI collagen degradation or blocking its receptors may prevent an overshooting wound-healing response, as occurs in fibrotic tissue remodeling, which is regularly coupled to deregulated ECM proteolysis, as illustrated in Fig. 5.

An external file that holds a picture, illustration, etc.
Object name is zh30101568860005.jpg

Type VI collagen as autoparacrine and paracrine activator of fibrogenesis. Type VI collagen microfibrils serve as sensor of early tissue injury and ECM destruction, usually initiated by release of ECM-degrading proteases like the MMPs by inflammatory cells. The generated proteolytic fragments activate type VI collagen receptors on (myo)fibroblasts that promote their fibrogenic activation, in part via integrins, focal adhesion kinase, and mitogen-activated kinase (Erk1 and Erk2) activation. There is also coactivation of the mitogenic PDGF-α receptor. The activated myofibroblasts then enhance their deposition of ECM components, including intact type VI collagen, whose ongoing degradation maintains a fibrogenic wound-healing response.

Perlecan, endorepellin.

The proteoglycan perlecan, a basement membrane HS proteoglycan, has been implicated in fibrosis (86) and has shown to have opposing terminal angiogenic activities. The NH2 terminus carrying three HS chains and the COOH terminus have proangiogenic and antiangiogenic properties, respectively (30). The COOH terminus is proteolytically processed by BMP-1 (110), yielding endorepellin, an 85-kDa perlecan domain V fragment (216). The COOH terminus V domain of perlecan is similar to another HS proteoglycan, agrin, which is the major proteoglycan of the glomerular basement membrane. The COOH-terminal agrin fragment is a known marker of neuronal muscular remodeling, but it has recently been shown to be a potential biomarker for renal function in renal transplant recipients (322). Like other mentioned angiogenesis inhibitors, such as arresten, canstatin, endostatin, and tumstatin, endorepellin mediates its antiangiogenesis functions through binding to α2β1-integrin, a key angiogenesis receptor (376). This interaction disintegrates the actin cytoskeleton and focal adhesions (237). The laminin-like globular domain (LG3) of endorepellin can be released by further proteolytic processing with BMP-1 (111). The LG3 has been associated with end-stage renal failure (241) and shown to be implicated in renal allograft rejection (240). Caspase-3 activation triggers cathepsin-L proteolytic processing of endorepellin to release the LG3 domain (48). LG3 induces α2β1-integrin and Src family kinase-dependent antiapoptotic pathways in fibroblasts (188). Thus the LG3 domain may affect, not only angiogenesis, but also collagen deposition and overall tissue stiffness, known hallmarks of fibrosis.

The ECM as Regulator of Cell Function: The ECM Interacts with Cells Through Integrins and DDRs and Induces Specific Signaling

Integrins and discoidin domain receptors (DDRs) are part of a network that enables the cell to sense and interact with the microenvironment and adapt to changes (102), through specific interactions and signaling, as outlined in Table 4. Integrins are involved in cell-cell interactions and the attachment of cells to the ECM. They are vital in development and tissue homeostasis, where they transmit “outside-in signals” and “inside-out signals,” which are important in the regulation of cellular proliferation, apoptosis, adhesion, migration, and growth (294). Their extracellular domain is responsible for sensing the microenvironment by ECM binding, which also modulates hepatocyte differentiation (283). Several proteins of the ECM, also the abundant collagens, signal through integrins and DDRs. In hepatic stellate cells (HSC), collagen type I signals through integrin-α1β1 (and to a minor degree through α2β1), whereas collagen type III engages mainly DDR1 (Table 4). The DDRs are receptor tyrosine kinases that specifically recognize collagens as their ligands (276). Similar to collagen-binding β1-integrins, the DDRs bind to specific secondary structural motifs within the collagen triple helix (383). DDRs are involved in tissue homeostasis and transduce signals regulating cell polarity, tissue morphogenesis, and cell differentiation (345). Here, DDR1 and DDR2, in particular, have demonstrated important functions in tissue homeostasis and cancers (243, 385), in which lack or inhibition of DDR1 attenuated fibrogenesis (35, 97, 120, 272, 276), whereas DDR2 deficiency promoted experimental liver fibrosis (101).

Table 4.

Selected key ECM proteins, their receptors, and downstream signaling

ProteinReceptorAssociated PathwaysEffectReference
Collagen type 1α1β1, α2β1, DDR1, DDR2ShcA, Nck2, Shp-2, STAT5, NF-κB, p37 MAPK, Src, Erk1/2, AP-1HSC activation, proliferation, and migration(73, 153, 275, 384)
Collagen type 3DDR1, DDR2ShcA, Nck2, Shp-2, STAT5, NF-κB, p37 MAPK, Src, Erk1/2, AP-1HSC activation, proliferation, and migration(73, 275, 384)
Collagen type 4α1β1, DDR1ShcA, Nck2, Shp-2, STAT5, NF-κB, p37 MAPKHSC activation, proliferation, and migration(73, 275, 384)
Collagen type 6NG2Paxillin, FAK, Erk2HSC and myofibroblast activation and proliferation(289)
Collagen type 18α1β1ILK, AktHepatocyte survival(85)
Fibronectinα5β1, αVβ1, αVβ3ERK, JNKHSC activation(262)
Fibronectin (Tenascin-C)αVβ6ERK, JNKCholangiocyte proliferation, local TGFβ1 activation(248, 258, 262)
Angiotensin IIAT1R, AT2RJAK2, Rho/RhoInflammation, HSC activation(22, 114, 177)
LumicanTLR4, β2 integrinsUnder investigationCollagen fibrillogenesis, requisite for hepatic fibrosis(184, 190)
Lamininα2β1, α6β1, α7β1, αVβ8FAK, ILKFibrogenic, HSC activation(136, 249)
FibromodulinUnder investigationUnder investigationHSC activation, collagen type 1 deposition, antifibrotic(136, 192, 221)
DecorinEGFRMetAntifibrotic by binding of TGF-β, collagen assembly(13, 14)
BiglycanTLR2, TLR4P38, ERK, NF-κBCollagen assembly, proinflammatory, antifibrotic by binding of TGF-β(106, 281)
SyndecansCD148, αVβ3PI3K, Src, MAPKHepatocyte proliferation, differentiation, and adhesion; HCV attachment, angiogenesis(10, 24, 25, 212, 371)
TWEAKFn14TRAF, JNK, NF-κBHSC proliferation, liver progenitor cell, proinflammatory(148, 333, 372)

AP, activator protein; FAK, focal adhesion kinase; ILK, integrin-linked kinase; AT1/2R, angiotensin type 1/2 receptor; HCV, hepatitis C virus; TWEAK, TNF-like weak inducer of apoptosis; TRAF, TNF receptor associated factor.

Liver Pathologies Associated with the ECM

Common to all underlying causes of fibrosis is the disruption of the normal ECM pattern attributable to activation of (myo)fibroblasts (activated HSCs) in the portal tracts initiated by tissue injury and inflammation. This enables proliferation and migration of these fibrogenic effector cells into the parenchyma along the sinusoids, which is the hallmark of incipient septum formation (289). The activated HSCs deposit excess ECM, serving to “close the wound” and to provide a scaffold for orderly liver regeneration in case of a short-term insult. However, with ongoing (inflammatory) insults, the excess ECM, which becomes dominated by fibrillar collagen (mainly type I and type III), generates scar tissue that merely maintains organ integrity but has lost the guiding function for orderly tissue regeneration (267). The deposition of ECM is dependent on etiology and consequently on activation of different subtypes of cells, including mainly myofibroblasts of variant activation state, suggesting that “fibrosis is not just fibrosis.” During fibrosis progression, the quantity and quality of the hepatic ECM changes with an up to 10-fold increase in collagenous and noncollagenous components followed by a shift in matrix composition from the low-density basement membrane-like matrix to an interstitial matrix containing mainly fibril-forming collagens (289). A specific characteristic of fibrosis development in the liver is the presence of specific growth points of the scar tissue, i.e., portal zone vs. central zone, and space of Disse, determining the development of portal, central, or pericellular fibrosis, with different clinical consequences as to development of liver failure, i.e., with relative preservation of liver function in portal fibrosis despite massive collagen accumulation (285). Thus the developmental pattern of fibrosis is dependent on the underlying etiology causing the fibrosis (51, 285).

The structural representation of fibrosis depends on the cell types and injury involved. Today, several fibroblast-like cell types have been identified in the liver, all of which also contribute to the development of fibrosis, including 1) septal myofibroblasts present in the inner part of fibrous septa, 2) activated HSCs located in capillarized sinusoids adjacent to expanded portal tracts, 3) interface myofibroblasts located at the edge of fibrous septa derived from activated HSCs (or portal fibroblasts) recruited at the site of injury where the ECM turnover (synthesis and degradation) and accompanying cell damage and inflammatory infiltration are highest (329), and 4) smooth muscle cells localized in the larger vessel walls (2, 99, 115, 198). Consequently, the activity of these different cell types results in distinct fibrosis patterns with formation of 1) portal-portal, 2) portal-central, or 3) central-central septa (Table 5).

Table 5.

The initial histological changes and developmental patterns of fibrosis in biliary fibrosis, viral hepatitis, and metabolic (nonalcoholic steatohepatitis) and alcoholic liver disease

EtiologyBiliary FibrosisViral HepatitisMetabolic/Alcoholic Liver Disease
Initial histological changesBile duct obstruction → fibroblast activation, enlargement of portal tracts with massive collagen deposition and mild inflammationPortal and parenchymal (hepatocellular) fibrogenesis → ECM deposition around portal vein and surrounding sinusoids, inflammatory cell infiltrate, bridging necrosisPerivenular and perihepatocellular fibrosis → HSC activation and collagen deposition in the space of Disse; advanced stages with extension and involvement of portal areas
Main fibrogenic cell typePortal myofibroblastsPortal myofibroblasts and HSCsHSCs and later portal myofibroblasts
Major ECM proteinsLaminin, fibronectin, tenascin, collagens (type I, III, IV, V, VI, XV, XIX), elastin, fibrillinLaminin, fibronectin, tenascin, collagens (type I, III, IV V, VI, XV, XIX), elastin, fibrillinCollagens (type I, III, IV, V, VI, XVIII), fibronectin, tenascin C, fibrillin
Fibrosis patternPortal-portal septumPortal-central septumChicken wire, later stages central-portal septum

Portal-portal septa.

In biliary fibrosis, portal fibroblasts residing next to the bile duct epithelium (or to biliary progenitors) are the major myofibroblast source. After injury, mainly to the bile ducts (but also severe damage to hepatocytes, which generates biliary progenitors), these fibroblasts undergo rapid activation, express prominent α-smooth muscle actin, and deposit a peribilary ECM (179, 285, 287, 288). The coproliferation of reactive bile ducts and periductular myofibroblasts results initially in periportal fibrosis and is followed by portal-portal septa formation surrounding the liver nodules, whereas the central vein and the sinusoids that connect from it to the portal tracts are anatomically preserved until later stages where HSCs and to a lesser extent fibrocytes contribute to disease progression.

Portal-central septa.

The initial histological changes in chronic viral hepatitis are characterized by inflammatory cell infiltration and matrix deposition around the portal tracts. The fibrotic pattern develops as portal-central septa because of portal-central bridging necrosis, suggesting that myofibroblasts as well as HSCs migrate from the portal tract and neighboring sinusoids into the developing septa (267, 285, 287, 288). In alcoholic and metabolic liver diseases, the fibrosis is characterized as the “chicken wire” pattern, in which the fibrillar matrix is deposited around groups of hepatocytes and sinusoids (51, 257).

Central-central septa.

These can be found secondary to venous outflow obstruction, such as in Budd Chiari syndrome. The central-central septa develop because of distal sinusoidal dilation, centrilobular fibrosis and necrosis, endothelialitis, and prominent platelet activation (196, 228).

How the organization of the structural and functional unit of the hepatic lobule is related to function and disease has been intensively discussed during the past decade. Several models describing the functional unit of the liver have been proposed. One model describes the hepatic acinus with the portal tract as its axis and its peripheral boundary circumscribed by an imaginary line connecting the neighboring terminal hepatic venules. The acinus is divided into three zones, each with different oxygen content and metabolic function (155). Another model describes the classic lobule as subdivided into several primary lobules, the so-called hepatic microvascular subunits. The subunits consist of a group of sinusoids supplied by a single venule and its associated termination of a branch of the hepatic arteriole from the adjacent portal space accompanied by hepatic parenchymal cells and associated cholangioles. Furthermore, a hepatocellular metabolic gradient has also been demonstrated in this proposed functional unit model (210).

Whether the distinct fibrosis patterns develop because of different cell phenotypes or specific mechanisms has to our knowledge not been elucidated. However, on the basis of the above mentioned models, one can assume that a combination of both phenotype and mechanisms, dependent on the local metabolic activity and oxygen content at the specific site of injury, is operative.

Primary cholangiopathies are of particular interest. Here changes are confined to the periportal area, with a mixed inflammatory cell infiltrate leading to periportal fibrosis. Interestingly, inflammation and fibrosis are not necessarily closely associated because the risk of biliary dysplasia and malignancy is not correlated with disease duration or severity (47, 327). Additionally, primary sclerosing cholangitis can lead to sinusoidal hypertension before development of varices and hemorrhages associated with cirrhosis (335).

ECM and HCC

Worldwide, HCC is the fifth most common cancer and the third most common cause of cancer-related death (304). In the US, the occurrence of HCC is increasing, mainly because of the rising prevalence of end-stage hepatitis C and NASH (160, 236). HCC is strongly correlated to fibrosis severity (203), with 90% of HCC cases arising in cirrhotic livers (299). Thus more than 80% of patients with hepatitis B and C that have HCC are cirrhotic (140). Similarly, HCC prevalence and progression is related to cirrhosis (93) in alcoholic steatohepatitis and NASH (8), with a yearly HCC incidence of 1.7% (93) and 2.6%, respectively (8). Mechanisms linking fibrosis and HCC are in need of further exploration, whereas there is increasing evidence that the fibrotic/cirrhotic and associated (immunosuppressive) inflammatory liver microenvironment appears to be a major player in HCC pathology, with several molecular pathways being shared between fibrogenesis and carcinogenesis (340). Thus, apart from direct carcinogens such as hepatitis B virus, microenvironmental stimuli include intrahepatic cell subpopulations, such as progenitor, immune and stellate cells, proliferative or differentiation-modulating cytokines, and growth factors like TGF-β1, EGF, IGF-1, and VEGF and altered ECM molecules and ECM receptors like the integrins. These interact and increase protease activities directed at the ECM or cell-associated structures, facilitating ECM remodeling, cancer cell proliferation, and migration. An example is MMP-2, which is able to degrade numerous components of the ECM that are closely correlated with tumor invasion and metastasis. Together with hypoxia-inducible factor 1α, which regulates proteolytic and angiogenic activities, MMP-2 is overexpressed and its activity enhanced in HCC compared with healthy tissue (359). Additionally, MMP-7 and MMP-26 levels are significantly higher in HCC tissue compared with adjacent healthy hepatic cells from patients with HCC (361) and strongly correlated to phosphorylated FGF receptor 2, supporting a direct link to angiogenesis (361).

Another class of enzymes that is linked to HCC development is LOXL2, which catalyzes collagen cross-linking, enhancing matrix stiffness and resistance to proteolysis, and promotes HCC metastasis (374). Induction of LOXL2 occurs via multiple regulators, including hypoxia, TGF-β1, and microRNAs (374). Matrix stiffness alone appears to determine fibrosis and HCC progression, as expression of procollagen, LOXL2, VEGF and the endothelial marker CD31 is highly correlated (82). Therefore, the ECM remains a yet insufficiently exploited target for HCC therapies.

Do We Need to Target the ECM?

We have highlighted the importance of the ECM in cell differentiation and function, fibrosis, and HCC development. The normal and especially fibrotic ECM affects cell phenotype through cell-ECM receptor interactions and liberated and stored growth factors/cytokines/chemokines and vice versa. Moreover, we have highlighted specific PTMs made to the ECM that themselves affect cell phenotype and fate differently than the parent molecules. Consequent to these observations, it should be possible to ameliorate progression of and reverse established fibrosis without crucially disrupting beneficial cell-ECM interactions by differentially targeting pathogenic ECM structures, PTMs, and protein fragments.

Clearly, interfering with the usual suspects TGF-β, PDGF, or VEGF may ameliorate fibrogenesis in certain settings, only in preclinical studies, to date. Direct targeting of the ECM and cells whose fibrogenic activation or fibrolytic activity depends on specific ECM structures that are prevalent in fibrosis would render a fibrosis-specific and potentially fibrolysis-inducing therapy. This appears feasible because the ECM is continuously remodeled, even in cirrhosis (289, 291, 301). Such therapy would ideally be coupled to elimination or suppression of the major cause of fibrosis, such as viral hepatitis B or C. Although this seems a long way to go, the advanced understanding of the pathophysiology of the ECM in fibrosis is already beginning to translate into the clinic, as exemplified in an ongoing clinical study with a LOXL2-blocking antibody.

Conclusions

There is a growing body of evidence that modifications made to the structural proteins of the matrix may be both a consequence of the disease as well as drivers of disease progression. Consequently, PTMs within specific ECM proteins (such as degradation products) may be more integrated in pathogenesis than previously thought. Evidence for the use of biomimetic peptides from the ECM, such as types IV and XVIIII collagen, as anti-ECM therapeutics, which block fibrogenesis and ECM remodeling, is emerging (189, 358, 388). These examples begin to suggest that matrix molecules themselves may be antifibrotic agents in addition to kinase inhibitors and receptor blockers. A further example on how PTMs regulate ECM function is cross-linking of collagens or elastin by LOXL2, which prevents degradation and increases stiffness of the tissue. These combined observations clearly suggest that the ECM is more than just a structural framework for tissues and may perform paracrine functions affecting cell phenotype and fate. It is of key interest to understand the role of each of the major ECM components and their PTMs as well as their signaling potential to grasp the full potential of the ECM and its importance in pathogenesis.

GRANTS

We acknowledge the funding from the Danish Ministry of Science, Technology and Innovation and the Danish Science Foundation (Den Danske Forskningsfond), the US National Institutes of Health, the German Research Foundation (DFG) and the European Research Foundation (ERC Advanced Grant to D. Schuppan).

DISCLOSURES

All authors except H. Krarup, A. Blanchard, and D. Schuppan are full-time employees of Nordic Bioscience. The authors declare that they have no other conflicts of interest.

AUTHOR CONTRIBUTIONS

Author contributions: M.A.K. and J.M.B.S. conception and design of research; M.A.K., F.G., M.J.N., C.L.B., A.B., D.J.L., and D.S. analyzed data; M.A.K., T.M.-J., F.G., J.H.K., M.J.N., N.-U.B.H., A.-C.B.-J., C.L.B., A.K., A.B., H.K., D.J.L., and D.S. interpreted results of experiments; M.A.K., T.M.-J., F.G., J.H.K., M.J.N., J.M.B.S., A.-C.B.-J., C.L.B., A.K., A.B., H.K., D.J.L., and D.S. prepared figures; M.A.K., T.M.-J., F.G., J.H.K., M.J.N., J.M.B.S., N.-U.B.H., A.-C.B.-J., C.L.B., A.K., A.B., H.K., D.J.L., and D.S. drafted manuscript; M.A.K., T.M.-J., F.G., J.H.K., M.J.N., J.M.B.S., N.-U.B.H., A.-C.B.-J., C.L.B., A.K., A.B., H.K., D.J.L., and D.S. edited and revised manuscript; M.A.K., T.M.-J., F.G., J.H.K., M.J.N., J.M.B.S., N.-U.B.H., A.-C.B.-J., C.L.B., A.K., A.B., H.K., D.J.L., and D.S. approved final version of manuscript; F.G. performed experiments.

REFERENCES

1. Forskelligedødsårsager i forskellige aldre. REGIO - Danske Regioners Analysemagasin 4: 14–15, 2014. [Google Scholar]
2. Abdel-Aziz G, Rescan PY, Clement B, Lebeau G, Rissel M, Grimaud JA, Campion JP, Guillouzo A. Cellular sources of matrix proteins in experimentally induced cholestatic rat liver. J Pathol 164: 167–174, 1991. [Abstract] [Google Scholar]
3. Adair-Kirk TL, Atkinson JJ, Broekelmann TJ, Doi M, Tryggvason K, Miner JH, Mecham RP, Senior RM. A site on laminin alpha 5, AQARSAASKVKVSMKF, induces inflammatory cell production of matrix metalloproteinase-9 and chemotaxis. J Immunol 171: 398–406, 2013. [Abstract] [Google Scholar]
4. Ahmad NN, Donald-McGinn DM, Zackai EH, Knowlton RG, LaRossa D, DiMascio J, Prockop DJ. A second mutation in the type II procollagen gene (COL2AI) causing stickler syndrome (arthro-ophthalmopathy) is also a premature termination codon. Am J Hum Genet 52: 39–45, 1993. [Europe PMC free article] [Abstract] [Google Scholar]
5. Ala-Kokko L, Baldwin CT, Moskowitz RW, Prockop DJ. Single base mutation in the type II procollagen gene (COL2A1) as a cause of primary osteoarthritis associated with a mild chondrodysplasia. Proc Natl Acad Sci USA 87: 6565–6568, 1990. [Europe PMC free article] [Abstract] [Google Scholar]
6. Allen KD, Griffin TM, Rodriguiz RM, Wetsel WC, Kraus VB, Huebner JL, Boyd LM, Setton LA. Decreased physical function and increased pain sensitivity in mice deficient for type IX collagen. Arthritis Rheum 60: 2684–2693, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
6a. Ameye L, Aria D, Jepsen K, Oldberg A, Xu T, Young MF. Abnormal collagen fibrils in tendons of biglycan/fibromodulin-deficient mice lead to gait impairment, ectopic ossification, and osteoarthritis. FASEB J 16: 673–680, 2002. [Abstract] [Google Scholar]
7. Asch AS, Barnwell J, Silverstein RL, Nachman RL. Isolation of the thrombospondin membrane receptor. J Clin Invest 79: 1054–1061, 1987. [Europe PMC free article] [Abstract] [Google Scholar]
8. Ascha MS, Hanouneh IA, Lopez R, Tamimi TA, Feldstein AF, Zein NN. The incidence and risk factors of hepatocellular carcinoma in patients with nonalcoholic steatohepatitis. Hepatology 51: 1972–1978, 2010. [Abstract] [Google Scholar]
9. Atkinson JC, Ruhl M, Becker J, Ackermann R, Schuppan D. Collagen VI regulates normal and transformed mesenchymal cell proliferation in vitro. Exp Cell Res 228: 283–291, 1996. [Abstract] [Google Scholar]
10. Auguste P, Vincent F, Babbiani G, Desmouliere A. The fibroblast and myofibroblast in inflammatory angiogenesis. In: Angiogenesis in Inflammation: Mechanisms and Clinical Correlates, edited by MP Seed and DA Walsh. New York, NY: Springer, 2008. p. 59–82. [Google Scholar]
11. Aumailley M, Gayraud B. Structure and biological activity of the extracellular matrix. J Mol Med 76: 253–265, 1998. [Abstract] [Google Scholar]
12. Babelova A, Moreth K, Tsalastra-Greul W, Zeng-Brouwers J, Eickelberg O, Young MF, Bruckner P, Pfeilschifter J, Schaefer RM, Gröne HJ, Schaefer L. Biglycan, a danger signal that activates the NLRP3 inflammasome via toll-like and P2X receptors. J Biol Chem 284: 24035–24048, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
13. Baghy K, Dezso K, Laszlo V, Fullar A, Peterfia B, Paku S, Nagy P, Schaff Z, Iozzo RV, Kovalszky I. Ablation of the decorin gene enhances experimental hepatic fibrosis and impairs hepatic healing in mice. Lab Invest 91: 439–451, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
14. Baghy K, Horvath Z, Regos E, Kiss K, Schaff Z, Iozzo RV, Kovalszky I. Decorin interferes with platelet-derived growth factor receptor signaling in experimental hepatocarcinogenesis. FEBS J 280: 2150–2164, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
15. Barascuk N, Vassiliadis E, Larsen L, Wang J, Zheng Q, Xing R, Cao Y, Crespo C, Lapret I, Sabatini M, Villeneuve N, Vilaine JP, Rasmussen LM, Register TC, Karsdal MA. Development and validation of an enzyme-linked immunosorbent assay for the quantification of a specific MMP-9 mediated degradation fragment of type III collagen–A novel biomarker of atherosclerotic plaque remodeling. Clin Biochem 44: 900–906, 2011. [Abstract] [Google Scholar]
16. Barascuk N, Veidal SS, Larsen L, Larsen DV, Larsen MR, Wang J, Zheng Q, Xing R, Cao Y, Rasmussen LM, Karsdal MA. A novel assay for extracellular matrix remodeling associated with liver fibrosis: An enzyme-linked immunosorbent assay (ELISA) for a MMP-9 proteolytically revealed neo-epitope of type III collagen. Clin Biochem 43: 899–904, 2010. [Abstract] [Google Scholar]
17. Barilla ML, Carsons SE. Fibronectin fragments and their role in inflammatory arthritis. Semin Arthritis Rheum 29: 252–265, 2000. [Abstract] [Google Scholar]
18. Barker DF, Hostikka SL, Zhou J, Chow LT, Oliphant AR, Gerken SC, Gregory MC, Skolnick MH, Atkin CL, Tryggvason K. Identification of mutations in the COL4A5 collagen gene in Alport syndrome. Science 248: 1224–1227, 1990. [Abstract] [Google Scholar]
19. Barker HE, Erler JT. The potential for LOXL2 as a target for future cancer treatment. Future Oncol 7: 707–710, 2011. [Abstract] [Google Scholar]
20. Barron L, Wynn TA. Macrophage activation governs schistosomiasis-induced inflammation and fibrosis. Eur J Immunol 41: 2509–2514, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
21. Barry-Hamilton V, Spangler R, Marshall D, McCauley S, Rodriguez HM, Oyasu M, Mikels A, Vaysberg M, Ghermazien H, Wai C, Garcia CA, Velayo AC, Jorgensen B, Biermann D, Tsai D, Green J, Zaffryar-Eilot S, Holzer A, Ogg S, Thai D, Neufeld G, Van Vlasselaer P, Smith V. Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a pathologic microenvironment. Nat Med 16: 1009–1017, 2010. [Abstract] [Google Scholar]
22. Bataller R, Brenner DA. Liver fibrosis. J Clin Invest 115: 209–218, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
24. Beauvais DM, Ell BJ, McWhorter AR, Rapraeger AC. Syndecan-1 regulates alphavbeta3 and alphavbeta5 integrin activation during angiogenesis and is blocked by synstatin, a novel peptide inhibitor. J Exp Med 206: 691–705, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
25. Beauvais DM, Rapraeger AC. Syndecan-1 couples the insulin-like growth factor-1 receptor to inside-out integrin activation. J Cell Sci 123: 3796–3807, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
26. Berenguer M, Schuppan D. Progression of liver fibrosis in post-transplant hepatitis C: mechanisms, assessment and treatment. J Hepatol 58: 1028–1041, 2013. [Abstract] [Google Scholar]
27. Besser TE, Potter KA, Bryan GM, Knowlen GG. An animal model of the Marfan syndrome. Am J Med Genet 37: 159–165, 1990. [Abstract] [Google Scholar]
28. Bi X, Tong C, Dockendorff A, Bancroft L, Gallagher L, Guzman G, Iozzo RV, Augenlicht LH, Yang W. Genetic deficiency of decorin causes intestinal tumor formation through disruption of intestinal cell maturation. Carcinogenesis 29: 1435–1440, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
29. Bignon M, Pichol-Thievend C, Hardouin J, Malbouyres M, Brechot N, Nasciutti L, Barret A, Teillon J, Guillon E, Etienne E, Caron M, Joubert-Caron R, Monnot C, Ruggiero F, Muller L, Germain S. Lysyl oxidase-like protein-2 regulates sprouting angiogenesis and type IV collagen assembly in the endothelial basement membrane. Blood 118: 3979–3989, 2011. [Abstract] [Google Scholar]
30. Bix G, Iozzo RV. Novel interactions of perlecan: unraveling perlecan's role in angiogenesis. Microsc Res Tech 71: 339–348, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
31. Boeker KH, Haberkorn CI, Michels D, Flemming P, Manns MP, Lichtinghagen R. Diagnostic potential of circulating TIMP-1 and MMP-2 as markers of liver fibrosis in patients with chronic hepatitis C. Clin Chim Acta 316: 71–81, 2002. [Abstract] [Google Scholar]
32. Bonaldo P, Braghetta P, Zanetti M, Piccolo S, Volpin D, Bressan GM. Collagen VI deficiency induces early onset myopathy in the mouse: an animal model for Bethlem myopathy. Hum Mol Genet 7: 2135–2140, 1998. [Abstract] [Google Scholar]
33. Borkham-Kamphorst E, Kovalenko E, van Roeyen CR, Gassler N, Bomble M, Ostendorf T, Floege J, Gressner AM, Weiskirchen R. Platelet-derived growth factor isoform expression in carbon tetrachloride-induced chronic liver injury. Lab Invest 88: 1090–1100, 2008. [Abstract] [Google Scholar]
34. Borkham-Kamphorst E, Stoll D, Gressner AM, Weiskirchen R. Antisense strategy against PDGF B-chain proves effective in preventing experimental liver fibrogenesis. Biochem Biophys Res Commun 321: 413–423, 2004. [Abstract] [Google Scholar]
35. Borza CM, Pozzi A. Discoidin domain receptors in disease. Matrix Biol 34: 185–192, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
36. Bosques CJ, Raguram S, Sasisekharan R. The sweet side of biomarker discovery. Nat Biotechnol 24: 1100–1101, 2006. [Abstract] [Google Scholar]
37. Boulter L, Govaere O, Bird TG, Radulescu S, Ramachandran P, Pellicoro A, Ridgway RA, Seo SS, Spee B, Van Rooijen N, Sansom OJ, Iredale JP, Lowell S, Roskams T, Forbes SJ. Macrophage-derived Wnt opposes Notch signaling to specify hepatic progenitor cell fate in chronic liver disease. Nat Med 18: 572–579, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
38. Bowness JM, Venditti M, Tarr AH, Taylor JR. Increase in epsilon(gamma-glutamyl)lysine crosslinks in atherosclerotic aortas. Atherosclerosis 111: 247–253, 1994. [Abstract] [Google Scholar]
39. Brand DD, Latham KA, Rosloniec EF. Collagen-induced arthritis. Nat Protoc 2: 1269–1275, 2007. [Abstract] [Google Scholar]
40. Brandan E, Retamal C, Cabello-Verrugio C, Marzolo MP. The low density lipoprotein receptor-related protein functions as an endocytic receptor for decorin. J Biol Chem 281: 31562–31571, 2006. [Abstract] [Google Scholar]
41. Briggs MD, Rasmussen IM, Weber JL, Yuen J, Reinker K, Garber AP, Rimoin DL, Cohn DH. Genetic linkage of mild pseudoachondroplasia (PSACH) to markers in the pericentromeric region of chromosome 19. Genomics 18: 656–660, 1993. [Abstract] [Google Scholar]
42. Briggs MD, Wright MJ, Mortier GR. Multiple Epiphyseal Dysplasia. New Delhi, India: Dominant, 1993. [Google Scholar]
43. Brinckmann J, Neess CM, Gaber Y, Sobhi H, Notbohm H, Hunzelmann N, Fietzek PP, Müller PK, Risteli J, Gebker R, Scharffetter-Kochanek K. Different pattern of collagen cross-links in two sclerotic skin diseases: lipodermatosclerosis and circumscribed scleroderma. J Invest Dermatol 117: 269–273, 2001. [Abstract] [Google Scholar]
44. Brown RA. In the beginning there were soft collagen-cell gels: towards better 3D connective tissue models? Exp Cell Res 319: 2460–2469, 2013. [Abstract] [Google Scholar]
45. Buckley CD, Pilling D, Henriquez NV, Parsonage G, Threlfall K, Scheel-Toellner D, Simmons DL, Akbar AN, Lord JM, Salmon M. RGD peptides induce apoptosis by direct caspase-3 activation. Nature 397: 534–539, 1999. [Abstract] [Google Scholar]
46. Buechler C, Wanninger J, Neumeier M. Adiponectin, a key adipokine in obesity related liver diseases. World J Gastroenterol 17: 2801–2811, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
47. Burak K, Angulo P, Pasha TM, Egan K, Petz J, Lindor KD. Incidence and risk factors for cholangiocarcinoma in primary sclerosing cholangitis. Am J Gastroenterol 99: 523–526, 2004. [Abstract] [Google Scholar]
48. Cailhier JF, Sirois I, Laplante P, Lepage S, Raymond MA, Brassard N, Prat A, Iozzo RV, Pshezhetsky AV, Hébert MJ. Caspase-3 activation triggers extracellular cathepsin L release and endorepellin proteolysis. J Biol Chem 283: 27220–27229, 2008. [Abstract] [Google Scholar]
49. Canturk NZ, Canturk Z, Ozden M, Dalcik H, Yardimoglu M, Tulubas F. Protective effect of IGF-1 on experimental liver cirrhosis-induced common bile duct ligation. Hepatogastroenterology 50: 2061–2066, 2003. [Abstract] [Google Scholar]
50. Carlson EC, Lin M, Liu CY, Kao WW, Perez VL, Pearlman E. Keratocan and lumican regulate neutrophil infiltration and corneal clarity in lipopolysaccharide-induced keratitis by direct interaction with CXCL1. J Biol Chem 282: 35502–35509, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
51. Cassiman D, Roskams T. Beauty is in the eye of the beholder: emerging concepts and pitfalls in hepatic stellate cell research. J Hepatol 37: 527–535, 2002. [Abstract] [Google Scholar]
52. Castera L, Bedossa P. How to assess liver fibrosis in chronic hepatitis C: serum markers or transient elastography vs. liver biopsy? Liver Int 31, Suppl 1: 13–17, 2011. [Abstract] [Google Scholar]
53. Castilla A, Prieto J, Fausto N. Transforming growth factors beta 1 and alpha in chronic liver disease. Effects of interferon alpha therapy. N Engl J Med 324: 933–940, 1991. [Abstract] [Google Scholar]
54. Chakravarti S. Functions of lumican and fibromodulin: lessons from knockout mice. Glycoconj J 19: 287–293, 2002. [Abstract] [Google Scholar]
55. Chen XD, Fisher LW, Robey PG, Young MF. The small leucine-rich proteoglycan biglycan modulates BMP-4-induced osteoblast differentiation. FASEB J 18: 948–958, 2004. [Abstract] [Google Scholar]
56. Chen XD, Shi S, Xu T, Robey PG, Young MF. Age-related osteoporosis in biglycan-deficient mice is related to defects in bone marrow stromal cells. J Bone Miner Res 17: 331–340, 2002. [Abstract] [Google Scholar]
57. Chen Y, Zheng S, Qi D, Zheng S, Guo J, Zhang S, Weng Z. Inhibition of Notch signaling by a gamma-secretase inhibitor attenuates hepatic fibrosis in rats. PLoS One 7: e46512, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
58. Cheng JH, She H, Han YP, Wang J, Xiong S, Asahina K, Tsukamoto H. Wnt antagonism inhibits hepatic stellate cell activation and liver fibrosis. Am J Physiol Gastrointest Liver Physiol 294: G39–G49, 2008. [Abstract] [Google Scholar]
59. Chiquet-Ehrismann R, Tucker RP. Connective tissues: signaling by tenascins. Int J Biochem Cell Biol 36: 1085–1089, 2004. [Abstract] [Google Scholar]
60. Chiquet-Ehrismann R, Tucker RP. Tenascins and the importance of adhesion modulation. Cold Spring Harb Perspect Biol 3: a004960, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
61. Christensen AI, Davidsen M, Ekholm O, Pedersen PV, Juel K. Danskernes Sundhed—Den Nationale Sundhedsprofil 2013. Copenhagen, Denmark: Sundhedsstyrelsen, 2014. [Google Scholar]
62. Cloos PA, Christgau S. Characterization of aged osteocalcin fragments derived from bone resorption. Clin Lab 50: 585–598, 2014. [Abstract] [Google Scholar]
63. Cloos PA, Christgau S. Post-translational modifications of proteins: implications for aging, antigen recognition, and autoimmunity. Biogerontology 5: 139–158, 2004. [Abstract] [Google Scholar]
64. Cloos PA, Jensen AL. Age-related de-phosphorylation of proteins in dentin: a biological tool for assessment of protein age. Biogerontology 1: 341–356, 2000. [Abstract] [Google Scholar]
65. Comalada M, Cardo M, Xaus J, Valledor AF, Lloberas J, Ventura F, Celada A. Decorin reverses the repressive effect of autocrine-produced TGF-beta on mouse macrophage activation. J Immunol 170: 4450–4456, 2003. [Abstract] [Google Scholar]
66. Condeelis J, Pollard JW. Macrophages: obligate partners for tumor cell migration, invasion, and metastasis. Cell 124: 263–266, 2006. [Abstract] [Google Scholar]
67. Cong M, Iwaisako K, Jiang C, Kisseleva T. Cell signals influencing hepatic fibrosis. Int J Hepatol 2012: 158547, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
68. Corsi A, Xu T, Chen XD, Boyde A, Liang J, Mankani M, Sommer B, Iozzo RV, Eichstetter I, Robey PG, Bianco P, Young MF. Phenotypic effects of biglycan deficiency are linked to collagen fibril abnormalities, are synergized by decorin deficiency, and mimic Ehlers-Danlos-like changes in bone and other connective tissues. J Bone Miner Res 17: 1180–1189, 2002. [Abstract] [Google Scholar]
69. Counts DF, Evans JN, Dipetrillo TA, Sterling KM Jr, Kelley J. Collagen lysyl oxidase activity in the lung increases during bleomycin-induced lung fibrosis. J Pharmacol Exp Ther 219: 675–678, 1981. [Abstract] [Google Scholar]
70. Czarny-Ratajczak M, Lohiniva J, Rogala P, Kozlowski K, Perala M, Carter L, Spector TD, Kolodziej L, Seppänen U, Glazar R, Królewski J, Latos-Bielenska A, Ala-Kokko L. A mutation in COL9A1 causes multiple epiphyseal dysplasia: further evidence for locus heterogeneity. Am J Hum Genet 69: 969–980, 2001. [Europe PMC free article] [Abstract] [Google Scholar]
71. da Rocha-Azevedo B, Grinnell F. Fibroblast morphogenesis on 3D collagen matrices: the balance between cell clustering and cell migration. Exp Cell Res 319: 2440–2446, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
72. Dai R, Iwama A, Wang S, Kapila YL. Disease-associated fibronectin matrix fragments trigger anoikis of human primary ligament cells: p53 and c-myc are suppressed. Apoptosis 10: 503–512, 2005. [Abstract] [Google Scholar]
73. Dancygier H, Schirmacher P. Fibrogenic reaction. In: Clinical Hepatology: Principles and Practice of Hepatobiliary Diseases, edited by H Dancygier., editor. New York, NY: Springer, 2010, p. 251–268. [Google Scholar]
74. Dang N, Murrell DF. Mutation analysis and characterization of COL7A1 mutations in dystrophic epidermolysis bullosa. Exp Dermatol 17: 553–568, 2008. [Abstract] [Google Scholar]
75. Danielson KG, Baribault H, Holmes DF, Graham H, Kadler KE, Iozzo RV. Targeted disruption of decorin leads to abnormal collagen fibril morphology and skin fragility. J Cell Biol 136: 729–743, 1997. [Europe PMC free article] [Abstract] [Google Scholar]
76. De Minicis S, Seki E, Uchinami H, Kluwe J, Zhang Y, Brenner DA, Schwabe RF. Gene expression profiles during hepatic stellate cell activation in culture and in vivo. Gastroenterology 132: 1937–1946, 2007. [Abstract] [Google Scholar]
77. Desnoyers L, Arnott D, Pennica D. WISP-1 binds to decorin and biglycan. J Biol Chem 276: 47599–47607, 2001. [Abstract] [Google Scholar]
78. Dhanabal M, Ramchandran R, Waterman MJ, Lu H, Knebelmann B, Segal M, Sukhatme VP. Endostatin induces endothelial cell apoptosis. J Biol Chem 274: 11721–11726, 1999. [Abstract] [Google Scholar]
79. Di Giovanni C, Grottesi A, Lavecchia A. Conformational switch of a flexible loop in human laminin receptor determines laminin-1 interaction. Eur Biophys J 41: 353–358, 2012. [Abstract] [Google Scholar]
80. Dieterich W, Ehnis T, Bauer M, Donner P, Volta U, Riecken EO, Schuppan D. Identification of tissue transglutaminase as the autoantigen of celiac disease. Nat Med 3: 797–801, 1997. [Abstract] [Google Scholar]
81. Dolberg DS, Bissell MJ. Inability of Rous sarcoma virus to cause sarcomas in the avian embryo. Nature 309: 552–556, 1984. [Abstract] [Google Scholar]
82. Dong Y, Xie X, Wang Z, Hu C, Zheng Q, Wang Y, Chen R, Xue T, Chen J, Gao D, Wu W, Ren Z, Cui J. Increasing matrix stiffness upregulates vascular endothelial growth factor expression in hepatocellular carcinoma cells mediated by integrin beta1. Biochem Biophys Res Commun 444: 427–432, 2014. [Abstract] [Google Scholar]
83. Duca L, Floquet N, Alix AJ, Haye B, Debelle L. Elastin as a matrikine. Crit Rev Oncol Hematol 49: 235–244, 2004. [Abstract] [Google Scholar]
84. Duffield JS, Lupher M, Thannickal VJ, Wynn TA. Host responses in tissue repair and fibrosis. Annu Rev Pathol 8: 241–276, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
85. Duncan MB, Yang C, Tanjore H, Boyle PM, Keskin D, Sugimoto H, Zeisberg M, Olsen BR, Kalluri R. Type XVIII collagen is essential for survival during acute liver injury in mice. Dis Model Mech 6: 942–951, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
86. Ebefors K, Granqvist A, Ingelsten M, Molne J, Haraldsson B, Nystrom J. Role of glomerular proteoglycans in IgA nephropathy. PLoS One 6: e18575, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
87. Eklund L, Piuhola J, Komulainen J, Sormunen R, Ongvarrasopone C, Fassler R, Muona A, Ilves M, Ruskoaho H, Takala TE, Pihlajaniemi T. Lack of type XV collagen causes a skeletal myopathy and cardiovascular defects in mice. Proc Natl Acad Sci USA 98: 1194–1199, 2001. [Europe PMC free article] [Abstract] [Google Scholar]
88. Elli L, Bergamini CM, Bardella MT, Schuppan D. Transglutaminases in inflammation and fibrosis of the gastrointestinal tract and the liver. Dig Liver Dis 41: 541–550, 2009. [Abstract] [Google Scholar]
89. Ezura Y, Chakravarti S, Oldberg A, Chervoneva I, Birk DE. Differential expression of lumican and fibromodulin regulate collagen fibrillogenesis in developing mouse tendons. J Cell Biol 151: 779–788, 2000. [Europe PMC free article] [Abstract] [Google Scholar]
90. Failli P, Ruocco C, De FR, Caligiuri A, Gentilini A, Giotti A, Gentilini P, Pinzani M. The mitogenic effect of platelet-derived growth factor in human hepatic stellate cells requires calcium influx. Am J Physiol 269: C1133–C1139, 1995. [Abstract] [Google Scholar]
91. Fallowfield DJ. Macrophage-derived VEGF and angiogenesis within the hepatic scar—new pathways unmasked in the resolution of fibrosis. Hepatology. In press. [Abstract] [Google Scholar]
92. Fang F, Liu L, Yang Y, Tamaki Z, Wei J, Marangoni RG, Bhattacharyya S, Summer RS, Ye B, Varga J. The adipokine adiponectin has potent anti-fibrotic effects mediated via adenosine monophosphate-activated protein kinase: novel target for fibrosis therapy. Arthritis Res Ther 14: R229, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
93. Fattovich G, Stroffolini T, Zagni I, Donato F. Hepatocellular carcinoma in cirrhosis: incidence and risk factors. Gastroenterology 127: S35–S50, 2004. [Abstract] [Google Scholar]
94. Feeney ER, Chung RT. Antiviral treatment of hepatitis C. BMJ 348: g3308, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
95. Fernandes RJ, Wilkin DJ, Weis MA, Wilcox WR, Cohn DH, Rimoin DL, Eyre DR. Incorporation of structurally defective type II collagen into cartilage matrix in kniest chondrodysplasia. Arch Biochem Biophys 355: 282–290, 1998. [Abstract] [Google Scholar]
96. Fiedler LR, Schonherr E, Waddington R, Niland S, Seidler DG, Aeschlimann D, Eble JA. Decorin regulates endothelial cell motility on collagen I through activation of insulin-like growth factor I receptor and modulation of alpha2beta1 integrin activity. J Biol Chem 283: 17406–17415, 2008. [Abstract] [Google Scholar]
97. Franco C, Hou G, Ahmad PJ, Fu EY, Koh L, Vogel WF, Bendeck MP. Discoidin domain receptor 1 (DDR1) deletion decreases atherosclerosis by accelerating matrix accumulation and reducing inflammation in low-density lipoprotein receptor-deficient mice. Circ Res 102: 1202–1211, 2008. [Abstract] [Google Scholar]
98. Freitas VM, Vilas-Boas VF, Pimenta DC, Loureiro V, Juliano MA, Carvalho MR, Pinheiro JJ, Camargo AC, Moriscot AS, Hoffman MP, Jaeger RG. SIKVAV, a laminin alpha1-derived peptide, interacts with integrins and increases protease activity of a human salivary gland adenoid cystic carcinoma cell line through the ERK 1/2 signaling pathway. Am J Pathol 171: 124–138, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
99. Friedman SL. Seminars in medicine of the Beth Israel Hospital, Boston. The cellular basis of hepatic fibrosis Mechanisms and treatment strategies. N Engl J Med 328: 1828–1835, 1993. [Abstract] [Google Scholar]
100. Friedman SL. Hepatic stellate cells: protean, multifunctional, and enigmatic cells of the liver. Physiol Rev 88: 125–172, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
101. Friedman SL, Sheppard D, Duffield JS, Violette S. Therapy for fibrotic diseases: nearing the starting line. Sci Transl Med 5: 167sr1, 2013. [Abstract] [Google Scholar]
102. Fu HL, Valiathan RR, Arkwright R, Sohail A, Mihai C, Kumarasiri M, Mahasenan KV, Mobashery S, Huang P, Agarwal G, Fridman R. Discoidin domain receptors: unique receptor tyrosine kinases in collagen-mediated signaling. J Biol Chem 288: 7430–7437, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
103. Fukai N, Eklund L, Marneros AG, Oh SP, Keene DR, Tamarkin L, Niemelä M, Ilves M, Li E, Pihlajaniemi T, Olsen BR. Lack of collagen XVIII/endostatin results in eye abnormalities. EMBO J 21: 1535–1544, 2002. [Europe PMC free article] [Abstract] [Google Scholar]
104. Funderburgh JL, Mitschler RR, Funderburgh ML, Roth MR, Chapes SK, Conrad GW. Macrophage receptors for lumican. A corneal keratan sulfate proteoglycan. Invest Ophthalmol Vis Sci 38: 1159–1167, 1997. [Abstract] [Google Scholar]
105. Gao B. Cytokines, STATs and liver disease. Cell Mol Immunol 2: 92–100, 2005. [Abstract] [Google Scholar]
106. Genovese F, Barascuk N, Larsen L, Larsen MR, Nawrocki A, Li Y, Zheng Q, Wang J, Veidal SS, Leeming DJ, Karsdal MA. Biglycan fragmentation in pathologies associated with extracellular matrix remodeling by matrix metalloproteinases. Fibrogenesis Tissue Repair 6: 9, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
107. Giambartolomei S, Covone F, Levrero M, Balsano C. Sustained activation of the Raf/MEK/Erk pathway in response to EGF in stable cell lines expressing the Hepatitis C Virus (HCV) core protein. Oncogene 20: 2606–2610, 2001. [Abstract] [Google Scholar]
108. Gill MR, Oldberg A, Reinholt FP. Fibromodulin-null murine knee joints display increased incidences of osteoarthritis and alterations in tissue biochemistry. Osteoarthritis Cartilage 10: 751–757, 2002. [Abstract] [Google Scholar]
109. Goldoni S, Humphries A, Nystrom A, Sattar S, Owens RT, McQuillan DJ, Ireton K, Iozzo RV. Decorin is a novel antagonistic ligand of the Met receptor. J Cell Biol 185: 743–754, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
110. Gonzalez D, Ghiringhelli G, Mautalen C. Acute antiosteoclastic effect of salmon calcitonin in osteoporotic women. Calcif Tissue Int 38: 71–75, 1986. [Abstract] [Google Scholar]
111. Gonzalez EM, Reed CC, Bix G, Fu J, Zhang Y, Gopalakrishnan B, Greenspan DS, Iozzo RV. BMP-1/Tolloid-like metalloproteases process endorepellin, the angiostatic C-terminal fragment of perlecan. J Biol Chem 280: 7080–7087, 2005. [Abstract] [Google Scholar]
112. Good DJ, Polverini PJ, Rastinejad F, Le Beau MM, Lemons RS, Frazier WA, Bouck NP. A tumor suppressor-dependent inhibitor of angiogenesis is immunologically and functionally indistinguishable from a fragment of thrombospondin. Proc Natl Acad Sci USA 87: 6624–6628, 1990. [Europe PMC free article] [Abstract] [Google Scholar]
113. Gostynski A, Llames S, Garcia M, Escamez MJ, Martinez-Santamaria L, Nijenhuis M, Meana A, Pas HH, Larcher F, Pasmooij AM, Jonkman MF, Del Rio M. Long-term survival of type XVII collagen revertant cells in an animal model of revertant cell therapy. J Invest Dermatol 134: 571–574, 2014. [Abstract] [Google Scholar]
114. Granzow M, Schierwagen R, Klein S, Kowallick B, Huss S, Linhart M, Mazar IG, Görtzen J, Vogt A, Schildberg FA, Gonzalez-Carmona MA, Wojtalla A, Krämer B, Nattermann J, Siegmund SV, Werner N, Fürst DO, Laleman W, Knolle P, Shah VH, Sauerbruch T, Trebicka J. Angiotensin-II type 1 receptor-mediated Janus kinase 2 activation induces liver fibrosis. Hepatology 60: 334–348, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
115. Gressner AM. Perisinusoidal lipocytes and fibrogenesis. Gut 35: 1331–1333, 1994. [Europe PMC free article] [Abstract] [Google Scholar]
116. Gressner AM, Weiskirchen R. Modern pathogenetic concepts of liver fibrosis suggest stellate cells and TGF-beta as major players and therapeutic targets. J Cell Mol Med 10: 76–99, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
117. Gressner AM, Yagmur E, Lahme B, Gressner O, Stanzel S. Connective tissue growth factor in serum as a new candidate test for assessment of hepatic fibrosis. Clin Chem 52: 1815–1817, 2006. [Abstract] [Google Scholar]
118. Gressner OA, Weiskirchen R, Gressner AM. Biomarkers of hepatic fibrosis, fibrogenesis and genetic pre-disposition pending between fiction and reality. J Cell Mol Med 11: 1031–1051, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
119. Gressner OA, Weiskirchen R, Gressner AM. Biomarkers of liver fibrosis: clinical translation of molecular pathogenesis or based on liver-dependent malfunction tests. Clin Chim Acta 381: 107–113, 2007. [Abstract] [Google Scholar]
120. Guerrot D, Kerroch M, Placier S, Vandermeersch S, Trivin C, Mael-Ainin M, Chatziantoniou C, Dussaule JC. Discoidin domain receptor 1 is a major mediator of inflammation and fibrosis in obstructive nephropathy. Am J Pathol 179: 83–91, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
121. Guidetti GF, Bartolini B, Bernardi B, Tira ME, Berndt MC, Balduini C, Torti M. Binding of von Willebrand factor to the small proteoglycan decorin. FEBS Lett 574: 95–100, 2004. [Abstract] [Google Scholar]
122. Haass NK, Smalley KS, Herlyn M. The role of altered cell-cell communication in melanoma progression. J Mol Histol 35: 309–318, 2004. [Abstract] [Google Scholar]
123. Hagg PM, Hagg PO, Peltonen S, Autio-Harmainen H, Pihlajaniemi T. Location of type XV collagen in human tissues and its accumulation in the interstitial matrix of the fibrotic kidney. Am J Pathol 150: 2075–2086, 1997. [Europe PMC free article] [Abstract] [Google Scholar]
124. Hamano Y, Zeisberg M, Sugimoto H, Lively JC, Maeshima Y, Yang C, Hynes RO, Werb Z, Sudhakar A, Kalluri R. Physiological levels of tumstatin, a fragment of collagen IV alpha3 chain, are generated by MMP-9 proteolysis and suppress angiogenesis via alphaV beta3 integrin. Cancer Cell 3: 589–601, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
125. Hanash SM, Pitteri SJ, Faca VM. Mining the plasma proteome for cancer biomarkers. Nature 452: 571–579, 2008. [Abstract] [Google Scholar]
126. Hayward C, Brock DJ. Fibrillin-1 mutations in Marfan syndrome and other type-1 fibrillinopathies. Hum Mutat 10: 415–423, 1997. [Abstract] [Google Scholar]
127. Heegaard AM, Corsi A, Danielsen CC, Nielsen KL, Jorgensen HL, Riminucci M, Young MF, Bianco P. Biglycan deficiency causes spontaneous aortic dissection and rupture in mice. Circulation 115: 2731–2738, 2007. [Abstract] [Google Scholar]
128. Heinonen S, Mannikko M, Klement JF, Whitaker-Menezes D, Murphy GF, Uitto J. Targeted inactivation of the type VII collagen gene (Col7a1) in mice results in severe blistering phenotype: a model for recessive dystrophic epidermolysis bullosa. J Cell Sci 112: 3641–3648, 1999. [Abstract] [Google Scholar]
129. Hemmann S, Graf J, Roderfeld M, Roeb E. Expression of MMPs and TIMPs in liver fibrosis—a systematic review with special emphasis on anti-fibrotic strategies. J Hepatol 46: 955–975, 2007. [Abstract] [Google Scholar]
130. Hintermann E, Bayer M, Pfeilschifter JM, Luster AD, Christen U. CXCL10 promotes liver fibrosis by prevention of NK cell mediated hepatic stellate cell inactivation. J Autoimmun 35: 424–435, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
131. Hinz B, Phan SH, Thannickal VJ, Galli A, Bochaton-Piallat ML, Gabbiani G. The myofibroblast: one function, multiple origins. Am J Pathol 170: 1807–1816, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
132. Hu K, Xu L, Cao L, Flahiff CM, Brussiau J, Ho K, Setton LA, Youn I, Guilak F, Olsen BR, Li Y. Pathogenesis of osteoarthritis-like changes in the joints of mice deficient in type IX collagen. Arthritis Rheum 54: 2891–2900, 2006. [Abstract] [Google Scholar]
133. Huang L, Haylor JL, Fisher M, Hau Z, El Nahas AM, Griffin M, Johnson TS. Do changes in transglutaminase activity alter latent transforming growth factor beta activation in experimental diabetic nephropathy? Nephrol Dial Transplant 25: 3897–3910, 2010. [Abstract] [Google Scholar]
134. Huang L, Haylor JL, Hau Z, Jones RA, Vickers ME, Wagner B, Griffin M, Saint RE, Coutts IG, El Nahas AM, Johnson TS. Transglutaminase inhibition ameliorates experimental diabetic nephropathy. Kidney Int 76: 383–394, 2009. [Abstract] [Google Scholar]
135. Hudson BG, Tryggvason K, Sundaramoorthy M, Neilson EG. Alport's syndrome, Goodpasture's syndrome, and type IV collagen. N Engl J Med 348: 2543–2556, 2003. [Abstract] [Google Scholar]
136. Hui AY, Friedman SL. Extracellular matrix. In: Signaling Pathways in Liver Diseases, edited by Dufour JF, Clavien PA, Trautwien C, Graf R. New York, NY: Springer, 2005. [Google Scholar]
137. Hynes RO. Integrins: bidirectional, allosteric signaling machines. Cell 110: 673–687, 2002. [Abstract] [Google Scholar]
138. Hynes RO. The extracellular matrix: not just pretty fibrils. Science 326: 1216–1219, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
139. Iacob D, Cai J, Tsonis M, Babwah A, Chakraborty C, Bhattacharjee RN, Lala PK. Decorin-mediated inhibition of proliferation and migration of the human trophoblast via different tyrosine kinase receptors. Endocrinology 149: 6187–6197, 2008. [Abstract] [Google Scholar]
140. Idilman R, De MN, Colantoni A, Van Thiel DH. Pathogenesis of hepatitis B and C-induced hepatocellular carcinoma. J Viral Hepat 5: 285–299, 1998. [Abstract] [Google Scholar]
141. Iismaa SE, Mearns BM, Lorand L, Graham RM. Transglutaminases and disease: lessons from genetically engineered mouse models and inherited disorders. Physiol Rev 89: 991–1023, 2009. [Abstract] [Google Scholar]
142. Ingber DE. Can cancer be reversed by engineering the tumor microenvironment? Semin Cancer Biol 18: 356–364, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
143. Ingman WV, Wyckoff J, Gouon-Evans V, Condeelis J, Pollard JW. Macrophages promote collagen fibrillogenesis around terminal end buds of the developing mammary gland. Dev Dyn 235: 3222–3229, 2006. [Abstract] [Google Scholar]
144. Inkson CA, Ono M, Bi Y, Kuznetsov SA, Fisher LW, Young MF. The potential functional interaction of biglycan and WISP-1 in controlling differentiation and proliferation of osteogenic cells. Cells Tissues Organs 189: 153–157, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
145. Jacenko O, Roberts DW, Campbell MR, McManus PM, Gress CJ, Tao Z. Linking hematopoiesis to endochondral skeletogenesis through analysis of mice transgenic for collagen X. Am J Pathol 160: 2019–2034, 2002. [Europe PMC free article] [Abstract] [Google Scholar]
146. Jackson GC, Mittaz-Crettol L, Taylor JA, Mortier GR, Spranger J, Zabel B, Le Merrer M, Cormier-Daire V, Hall CM, Offiah A, Wright MJ, Savarirayan R, Nishimura G, Ramsden SC, Elles R, Bonafe L, Superti-Furga A, Unger S, Zankl A, Briggs MD. Pseudoachondroplasia and multiple epiphyseal dysplasia: a 7-year comprehensive analysis of the known disease genes identify novel and recurrent mutations and provides an accurate assessment of their relative contribution. Hum Mutat 33: 144–157, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
147. Jakob A, Unger S, Arnold R, Grohmann J, Kraus C, Schlensak C, Stiller B. A family with a new elastin gene mutation: broad clinical spectrum, including sudden cardiac death. Cardiol Young 21: 62–65, 2011. [Abstract] [Google Scholar]
148. Jakubowski A, Ambrose C, Parr M, Lincecum JM, Wang MZ, Zheng TS, Browning B, Michaelson JS, Baetscher M, Wang B, Bissell DM, Burkly LC. TWEAK induces liver progenitor cell proliferation. J Clin Invest 115: 2330–2340, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
149. Janmey PA, Wells RG, Assoian RK, McCulloch CA. From tissue mechanics to transcription factors. Differentiation 86: 112–120, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
150. Jepsen KJ, Wu F, Peragallo JH, Paul J, Roberts L, Ezura Y, Oldberg A, Birk DE, Chakravarti S. A syndrome of joint laxity and impaired tendon integrity in lumican- and fibromodulin-deficient mice. J Biol Chem 277: 35532–35540, 2002. [Abstract] [Google Scholar]
151. Jiao J, Friedman SL, Aloman C. Hepatic fibrosis. Curr Opin Gastroenterol 25: 223–229, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
152. Johnson TS, El-Koraie AF, Skill NJ, Baddour NM, El Nahas AM, Njloma M, Adam AG, Griffin M. Tissue transglutaminase and the progression of human renal scarring. J Am Soc Nephrol 14: 2052–2062, 2003. [Abstract] [Google Scholar]
153. Jokinen J, Dadu E, Nykvist P, Kapyla J, White DJ, Ivaska J, Vehviläinen P, Reunanen H, Larjava H, Häkkinen L, Heino J. Integrin-mediated cell adhesion to type I collagen fibrils. J Biol Chem 279: 31956–31963, 2004. [Abstract] [Google Scholar]
154. Jung Y, McCall SJ, Li YX, Diehl AM. Bile ductules and stromal cells express hedgehog ligands and/or hedgehog target genes in primary biliary cirrhosis. Hepatology 45: 1091–1096, 2007. [Abstract] [Google Scholar]
155. Jungermann K, Kietzmann T. Zonation of parenchymal and nonparenchymal metabolism in liver. Annu Rev Nutr 16: 179–203, 1996. [Abstract] [Google Scholar]
156. Kadler KE, Hill A, Canty-Laird EG. Collagen fibrillogenesis: fibronectin, integrins, and minor collagens as organizers and nucleators. Curr Opin Cell Biol 20: 495–501, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
157. Kallis YN, Robson AJ, Fallowfield JA, Thomas HC, Alison MR, Wright NA, Goldin RD, Iredale JP, Forbes SJ. Remodelling of extracellular matrix is a requirement for the hepatic progenitor cell response. Gut 60: 525–533, 2011. [Abstract] [Google Scholar]
158. Kamada Y, Tamura S, Kiso S, Matsumoto H, Saji Y, Yoshida Y, Fukui K, Maeda N, Nishizawa H, Nagaretani H, Okamoto Y, Kihara S, Miyagawa J, Shinomura Y, Funahashi T, Matsuzawa Y. Enhanced carbon tetrachloride-induced liver fibrosis in mice lacking adiponectin. Gastroenterology 125: 1796–1807, 2003. [Abstract] [Google Scholar]
159. Kannangai R, Sahin F, Torbenson MS. EGFR is phosphorylated at Ty845 in hepatocellular carcinoma. Mod Pathol 19: 1456–1461, 2006. [Abstract] [Google Scholar]
160. Kanwal F, Hoang T, Kramer JR, Asch SM, Goetz MB, Zeringue A, Richardson P, El-Serag HB. Increasing prevalence of HCC and cirrhosis in patients with chronic hepatitis C virus infection. Gastroenterology 140: 1182–1188, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
161. Karsdal MA, Henriksen K, Leeming DJ, Woodworth T, Vassiliadis E, Bay-Jensen AC. Novel combinations of Post-Translational Modification (PTM) neo-epitopes provide tissue-specific biochemical markers—are they the cause or the consequence of the disease? Clin Biochem 43: 793–804, 2010. [Abstract] [Google Scholar]
162. Karsdal MA, Krarup H, Sand JM, Christensen PB, Gerstoft J, Leeming DJ, Weis N, Schaffalitzky de Muckadell OB, Krag A. Review article: the efficacy of biomarkers in chronic fibroproliferative diseases—early diagnosis and prognosis, with liver fibrosis as an exemplar. Aliment Pharmacol Ther 40: 233–249, 2014. [Abstract] [Google Scholar]
163. Karsdal MA, Larsen L, Engsig MT, Lou H, Ferreras M, Lochter A, Delaissé JM, Foged NT. Matrix metalloproteinase-dependent activation of latent transforming growth factor-beta controls the conversion of osteoblasts into osteocytes by blocking osteoblast apoptosis. J Biol Chem 277: 44061–44067, 2002. [Abstract] [Google Scholar]
164. Karsdal MA, Nielsen MJ, Sand JM, Henriksen K, Genovese F, Bay-Jensen AC, Smith V, Adamkewicz JI, Christiansen C, Leeming DJ. Extracellular matrix remodeling: the common denominator in connective tissue diseases. Possibilities for evaluation and current understanding of the matrix as more than a passive architecture, but a key player in tissue failure. Assay Drug Dev Technol 11: 70–92, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
165. Karumanchi SA, Jha V, Ramchandran R, Karihaloo A, Tsiokas L, Chan B, Dhanabal M, Hanai JI, Venkataraman G, Shriver Z, Keiser N, Kalluri R, Zeng H, Mukhopadhyay D, Chen RL, Lander AD, Hagihara K, Yamaguchi Y, Sasisekharan R, Cantley L, Sukhatme VP. Cell surface glypicans are low-affinity endostatin receptors. Mol Cell 7: 811–822, 2001. [Abstract] [Google Scholar]
166. Kashtan CE. Animal models of Alport syndrome. Nephrol Dial Transplant 17: 1359–1362, 2002. [Abstract] [Google Scholar]
167. Kashtan CE, Kim Y, Lees GE, Thorner PS, Virtanen I, Miner JH. Abnormal glomerular basement membrane laminins in murine, canine, and human Alport syndrome: aberrant laminin alpha2 deposition is species independent. J Am Soc Nephrol 12: 252–260, 2001. [Abstract] [Google Scholar]
168. Kelly JD, Haldeman BA, Grant FJ, Murray MJ, Seifert RA, Bowen-Pope DF, Cooper JA, Kazlauskas A. Platelet-derived growth factor (PDGF) stimulates PDGF receptor subunit dimerization and intersubunit trans-phosphorylation. J Biol Chem 266: 8987–8992, 1991. [Abstract] [Google Scholar]
169. Kielland KB, Delaveris GJ, Rogde S, Eide TJ, Amundsen EJ, Dalgard O. Liver fibrosis progression at autopsy in injecting drug users infected by hepatitis C: a longitudinal long-term cohort study. J Hepatol 60: 260–266, 2014. [Abstract] [Google Scholar]
170. Kielty CM. Elastic fibres in health and disease. Expert Rev Mol Med 8: 1–23, 2006. [Abstract] [Google Scholar]
171. Kim YM, Kim EC, Kim Y. The human lysyl oxidase-like 2 protein functions as an amine oxidase toward collagen and elastin. Mol Biol Rep 38: 145–149, 2011. [Abstract] [Google Scholar]
172. Kipnes JR, Xu L, Han F, Rallapalli R, Jimenez S, Hall DJ, Tuan RS, Li Y. Molecular cloning and expression patterns of mouse cartilage oligomeric matrix protein gene. Osteoarthritis Cartilage 8: 236–239, 2000. [Abstract] [Google Scholar]
173. Kirovski G, Gabele E, Dorn C, Moleda L, Niessen C, Weiss TS, Wobser H, Schacherer D, Buechler C, Wasmuth HE, Hellerbrand C. Hepatic steatosis causes induction of the chemokine RANTES in the absence of significant hepatic inflammation. Int J Clin Exp Pathol 3: 675–680, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
174. Kisseleva T, Bhattacharya S, Braunstein J, Schindler CW. Signaling through the JAK/STAT pathway, recent advances and future challenges. Gene 285: 1–24, 2002. [Abstract] [Google Scholar]
175. Kisseleva T, Brenner DA. Fibrogenesis of parenchymal organs. Proc Am Thorac Soc 5: 338–342, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
176. Kisseleva T, Cong M, Paik Y, Scholten D, Jiang C, Benner C, Iwaisako K, Moore-Morris T, Scott B, Tsukamoto H, Evans SM, Dillmann W, Glass CK, Brenner DA. Myofibroblasts revert to an inactive phenotype during regression of liver fibrosis. Proc Natl Acad Sci USA 109: 9448–9453, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
177. Kitamura K, Tada S, Nakamoto N, Toda K, Horikawa H, Kurita S, Tsunematsu S, Kumagai N, Ishii H, Saito H, Hibi T. Rho/Rho kinase is a key enzyme system involved in the angiotensin II signaling pathway of liver fibrosis and steatosis. J Gastroenterol Hepatol 22: 2022–2033, 2007. [Abstract] [Google Scholar]
178. Kitaya K, Yasuo T. Dermatan sulfate proteoglycan biglycan as a potential selectin L/CD44 ligand involved in selective recruitment of peripheral blood CD16(-) natural killer cells into human endometrium. J Leukoc Biol 85: 391–400, 2009. [Abstract] [Google Scholar]
179. Knittel T, Kobold D, Saile B, Grundmann A, Neubauer K, Piscaglia F, Ramadori G. Rat liver myofibroblasts and hepatic stellate cells: different cell populations of the fibroblast lineage with fibrogenic potential. Gastroenterology 117: 1205–1221, 1999. [Abstract] [Google Scholar]
180. Kohli A, Shaffer A, Sherman A, Kottilil S. Treatment of hepatitis C: a systematic review. JAMA 312: 631–640, 2014. [Abstract] [Google Scholar]
181. Kolb M, Margetts PJ, Sime PJ, Gauldie J. Proteoglycans decorin and biglycan differentially modulate TGF-β-mediated fibrotic responses in the lung. Am J Physiol Lung Cell Mol Physiol 280: L1327–L1334, 2001. [Abstract] [Google Scholar]
182. Konstandin MH, Volkers M, Collins B, Quijada P, Quintana M, De La Torre A, Ormachea L, Din S, Gude N, Toko H, Sussman MA. Fibronectin contributes to pathological cardiac hypertrophy but not physiological growth. Basic Res Cardiol 108: 375, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
183. Kontusaari S, Tromp G, Kuivaniemi H, Romanic AM, Prockop DJ. A mutation in the gene for type III procollagen (COL3A1) in a family with aortic aneurysms. J Clin Invest 86: 1465–1473, 1990. [Europe PMC free article] [Abstract] [Google Scholar]
184. Krishnan A, Li X, Kao WY, Viker K, Butters K, Masuoka H, Knudsen B, Gores G, Charlton M. Lumican, an extracellular matrix proteoglycan, is a novel requisite for hepatic fibrosis. Lab Invest 92: 1712–1725, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
185. Krueger KE, Srivastava S. Posttranslational protein modifications: current implications for cancer detection, prevention, and therapeutics. Mol Cell Proteomics 5: 1799–1810, 2006. [Abstract] [Google Scholar]
186. Kuriyama N, Duarte S, Hamada T, Busuttil RW, Coito AJ. Tenascin-C: a novel mediator of hepatic ischemia and reperfusion injury. Hepatology 54: 2125–2136, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
187. Lampe AK, Bushby KM. Collagen VI related muscle disorders. J Med Genet 42: 673–685, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
188. Laplante P, Raymond MA, Labelle A, Abe J, Iozzo RV, Hebert MJ. Perlecan proteolysis induces an alpha2beta1 integrin- and Src family kinase-dependent anti-apoptotic pathway in fibroblasts in the absence of focal adhesion kinase activation. J Biol Chem 281: 30383–30392, 2006. [Abstract] [Google Scholar]
189. Lee E, Lee SJ, Koskimaki JE, Han Z, Pandey NB, Popel AS. Inhibition of breast cancer growth and metastasis by a biomimetic peptide. Sci Rep 4: 7139, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
190. Lee S, Bowrin K, Hamad AR, Chakravarti S. Extracellular matrix lumican deposited on the surface of neutrophils promotes migration by binding to beta2 integrin. J Biol Chem 284: 23662–23669, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
191. Lee UE, Friedman SL. Mechanisms of hepatic fibrogenesis. Best Pract Res Clin Gastroenterol 25: 195–206, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
192. Lee YH, Schiemann WP. Fibromodulin suppresses nuclear factor-kappaB activity by inducing the delayed degradation of IKBA via a JNK-dependent pathway coupled to fibroblast apoptosis. J Biol Chem 286: 6414–6422, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
193. Lee ZW, Kwon SM, Kim SW, Yi SJ, Kim YM, Ha KS. Activation of in situ tissue transglutaminase by intracellular reactive oxygen species. Biochem Biophys Res Commun 305: 633–640, 2003. [Abstract] [Google Scholar]
194. Leeming DJ, Byrjalsen I, Jimenez W, Christiansen C, Karsdal MA. Protein fingerprinting of the extracellular matrix remodelling in a rat model of liver fibrosis—a serological evaluation. Liver Int 33: 439–447, 2013. [Abstract] [Google Scholar]
195. Leeming DJ, Karsdal MA, Byrjalsen I, Bendtsen F, Trebicka J, Nielsen MJ, Christiansen C, Møller S, Krag A. Novel serological neo-epitope markers of extracellular matrix proteins for the detection of portal hypertension. Aliment Pharmacol Ther 38: 1086–1096, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
196. Lefkowitch JH, Mendez L. Morphologic features of hepatic injury in cardiac disease and shock. J Hepatol 2: 313–327, 1986. [Abstract] [Google Scholar]
197. Li JT, Liao ZX, Ping J, Xu D, Wang H. Molecular mechanism of hepatic stellate cell activation and antifibrotic therapeutic strategies. J Gastroenterol 43: 419–428, 2008. [Abstract] [Google Scholar]
198. Libbrecht L, Cassiman D, Desmet V, Roskams T. The correlation between portal myofibroblasts and development of intrahepatic bile ducts and arterial branches in human liver. Liver 22: 252–258, 2002. [Abstract] [Google Scholar]
199. Lima BL, Santos EJ, Fernandes GR, Merkel C, Mello MR, Gomes JP, Soukoyan M, Kerkis A, Massironi SM, Visintin JA, Pereira LV. A new mouse model for marfan syndrome presents phenotypic variability associated with the genetic background and overall levels of Fbn1 expression. PLoS One 5: e14136, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
200. Liu X, Hu H, Yin JQ. Therapeutic strategies against TGF-beta signaling pathway in hepatic fibrosis. Liver Int 26: 8–22, 2006. [Abstract] [Google Scholar]
201. Liu X, Wu H, Byrne M, Krane S, Jaenisch R. Type III collagen is crucial for collagen I fibrillogenesis and for normal cardiovascular development. Proc Natl Acad Sci USA 94: 1852–1856, 1997. [Europe PMC free article] [Abstract] [Google Scholar]
202. Loetscher P, Seitz M, Clark-Lewis I, Baggiolini M, Moser B. Activation of NK cells by CC chemokines. Chemotaxis, Ca2+ mobilization, and enzyme release. J Immunol 156: 322–327, 1996. [Abstract] [Google Scholar]
203. Lok AS, Seeff LB, Morgan TR, di Bisceglie AM, Sterling RK, Curto TM, Curto TM, Everson GT, Lindsay KL, Lee WM, Bonkovsky HL, Dienstag JL, Ghany MG, Morishima C, Goodman ZD.; HALT-C Trial Group. Incidence of hepatocellular carcinoma and associated risk factors in hepatitis C-related advanced liver disease. Gastroenterology 136: 138–148, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
204. Lorand L, Graham RM. Transglutaminases: crosslinking enzymes with pleiotropic functions. Nat Rev Mol Cell Biol 4: 140–156, 2003. [Abstract] [Google Scholar]
205. Lucey EC, Goldstein RH, Stone PJ, Snider GL. Remodeling of alveolar walls after elastase treatment of hamsters. Results of elastin and collagen mRNA in situ hybridization. Am J Respir Crit Care Med 158: 555–564, 1998. [Abstract] [Google Scholar]
206. Makitie O, Susic M, Cole WG. Early-onset metaphyseal chondrodysplasia type Schmid associated with a COL10A1 frame-shift mutation and impaired trimerization of wild-type alpha1(X) protein chains. J Orthop Res 28: 1497–1501, 2010. [Abstract] [Google Scholar]
207. Mangala LS, Mehta K. Tissue transglutaminase (TG2) in cancer biology. Prog Exp Tumor Res 38: 125–138, 2005. [Abstract] [Google Scholar]
208. Marcellin P, Gane E, Buti M, Afdhal N, Sievert W, Jacobson IM, Washington MK, Germanidis G, Flaherty JF, Schall RA, Bornstein JD, Kitrinos KM, Subramanian GM, McHutchison JG, Heathcote EJ. Regression of cirrhosis during treatment with tenofovir disoproxil fumarate for chronic hepatitis B: a 5-year open-label follow-up study. Lancet 381: 468–475, 2013. [Abstract] [Google Scholar]
209. Marx J. Cancer research. Inflammation and cancer: the link grows stronger. Science 306: 966–968, 2004. [Abstract] [Google Scholar]
210. McCuskey R. Anatomy of the liver. In: Zakim and Boyer's Hepatology: A Textbook of Liver Disease, ed. 6, edited by Boyer TD, Manns MP, Sanyal AJ. Philadelphia, PA: Elsevier Saunders; 2012, p. 3–19. [Google Scholar]
211. Merline R, Schaefer RM, Schaefer L. The matricellular functions of small leucine-rich proteoglycans (SLRPs). J Cell Commun Signal 3: 323–335, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
212. Metwaly HA, Al-Gayyar MM, Eletreby S, Ebrahim MA, El-Shishtawy MM. Relevance of serum levels of interleukin-6 and syndecan-1 in patients with hepatocellular carcinoma. Sci Pharm 80: 179–188, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
213. Milewicz DM, Urban Z, Boyd C. Genetic disorders of the elastic fiber system. Matrix Biol 19: 471–480, 2000. [Abstract] [Google Scholar]
214. Mintz B, Illmensee K. Normal genetically mosaic mice produced from malignant teratocarcinoma cells. Proc Natl Acad Sci USA 2: 3585–3589, 1975. [Europe PMC free article] [Abstract] [Google Scholar]
215. Mirza A, Liu SL, Frizell E, Zhu J, Maddukuri S, Martinez J, Davies P, Schwarting R, Norton P, Zern MA. A role for tissue transglutaminase in hepatic injury and fibrogenesis, and its regulation by NF-κB. Am J Physiol 272: G281–G288, 1997. [Abstract] [Google Scholar]
216. Mongiat M, Sweeney SM, San Antonio JD, Fu J, Iozzo RV. Endorepellin, a novel inhibitor of angiogenesis derived from the C terminus of perlecan. J Biol Chem 278: 4238–4249, 2003. [Abstract] [Google Scholar]
217. Morell CM, Strazzabosco M. Notch signaling and new therapeutic options in liver disease. J Hepatol 60: 885–890, 2014. [Abstract] [Google Scholar]
218. Moreno M, Munoz R, Aroca F, Labarca M, Brandan E, Larrain J. Biglycan is a new extracellular component of the Chordin-BMP4 signaling pathway. EMBO J 24: 1397–1405, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
219. Moreno-Bueno G, Salvador F, Martin A, Floristan A, Cuevas EP, Santos V, Csiszar K, Dubus P, Haigh JJ, Bigas A, Portillo F, Cano A. Lysyl oxidase-like 2 (LOXL2), a new regulator of cell polarity required for metastatic dissemination of basal-like breast carcinomas. EMBO Mol Med 3: 528–544, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
220. Morla A, Zhang Z, Ruoslahti E. Superfibronectin is a functionally distinct form of fibronectin. Nature 367: 193–196, 1994. [Abstract] [Google Scholar]
221. Mormone E, Lu Y, Ge X, Fiel MI, Nieto N. Fibromodulin, an oxidative stress-sensitive proteoglycan, regulates the fibrogenic response to liver injury in mice. Gastroenterology 142: 612–621, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
222. Mundel TM, Kalluri R. Type IV collagen-derived angiogenesis inhibitors. Microvasc Res 74: 85–89, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
223. Murray PJ, Wynn TA. Protective and pathogenic functions of macrophage subsets. Nat Rev Immunol 11: 723–737, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
224. Murshed M, Smyth N, Miosge N, Karolat J, Krieg T, Paulsson M, Nischt R. The absence of nidogen 1 does not affect murine basement membrane formation. Mol Cell Biol 20: 7007–7012, 2000. [Europe PMC free article] [Abstract] [Google Scholar]
225. Mutolo MJ, Morris KJ, Leir SH, Caffrey TC, Lewandowska MA, Hollingsworth MA, Harris A. Tumor suppression by collagen XV is independent of the restin domain. Matrix Biol 31: 285–289, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
226. Mutsaers SE, Marshall RP, Goldsack NR, Laurent GJ, McAnulty RJ. Effect of endothelin receptor antagonists (BQ-485, Ro 47-0203) on collagen deposition during the development of bleomycin-induced pulmonary fibrosis in rats. Pulm Pharmacol Ther 11: 221–225, 1998. [Abstract] [Google Scholar]
227. Mydel P, Shipley JM, Adair-Kirk TL, Kelley DG, Broekelmann TJ, Mecham RP, Senior RM. Neutrophil elastase cleaves laminin-332 (laminin-5) generating peptides that are chemotactic for neutrophils. J Biol Chem 283: 9513–9522, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
228. Myers RP, Cerini R, Sayegh R, Moreau R, Degott C, Lebrec D, Lee SS. Cardiac hepatopathy: clinical, hemodynamic, and histologic characteristics and correlations. Hepatology 37: 393–400, 2003. [Abstract] [Google Scholar]
229. Nasir GA, Mohsin S, Khan M, Shams S, Ali G, Khan SN, Riazuddin S. Mesenchymal stem cells and Interleukin-6 attenuate liver fibrosis in mice. J Transl Med 11: 78, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
230. Neskey DM, Ambesi A, Pumiglia KM, Keown-Longo PJ. Endostatin and anastellin inhibit distinct aspects of the angiogenic process. J Exp Clin Cancer Res 27: 61, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
231. Neubauer K, Kruger M, Quondamatteo F, Knittel T, Saile B, Ramadori G. Transforming growth factor-beta1 stimulates the synthesis of basement membrane proteins laminin, collagen type IV and entactin in rat liver sinusoidal endothelial cells. J Hepatol 31: 692–702, 1999. [Abstract] [Google Scholar]
232. Nguyen DM, El-Serag HB. The epidemiology of obesity. Gastroenterol Clin North Am 39: 1–7, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
233. Niederau C. Chronic hepatitis B in 2014: Great therapeutic progress, large diagnostic deficit. World J Gastroenterol 20: 11595–11617, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
234. Nili N, Cheema AN, Giordano FJ, Barolet AW, Babaei S, Hickey R, Eskandarian MR, Smeets M, Butany J, Pasterkamp G, Strauss BH. Decorin inhibition of PDGF-stimulated vascular smooth muscle cell function: potential mechanism for inhibition of intimal hyperplasia after balloon angioplasty. Am J Pathol 163: 869–878, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
235. Nishie W, Sawamura D, Goto M, Ito K, Shibaki A, McMillan JR, Sakai K, Nakamura H, Olasz E, Yancey KB, Akiyama M, Shimizu H. Humanization of autoantigen. Nat Med 13: 378–383, 2007. [Abstract] [Google Scholar]
236. Nordenstedt H, White DL, El-Serag HB. The changing pattern of epidemiology in hepatocellular carcinoma. Dig Liver Dis 42, Suppl 3: S206–S214, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
237. Nystrom A, Shaik ZP, Gullberg D, Krieg T, Eckes B, Zent R, Pozzi A, Iozzo RV. Role of tyrosine phosphatase SHP-1 in the mechanism of endorepellin angiostatic activity. Blood 114: 4897–4906, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
238. O'Brien LE, Jou TS, Pollack AL, Zhang Q, Hansen SH, Yurchenco P, Mostov KE. Rac1 orientates epithelial apical polarity through effects on basolateral laminin assembly. Nat Cell Biol 3: 831–838, 2001. [Abstract] [Google Scholar]
239. O'Reilly MS, Boehm T, Shing Y, Fukai N, Vasios G, Lane WS, Flynn E, Birkhead JR, Olsen BR, Folkman J. Endostatin: an endogenous inhibitor of angiogenesis and tumor growth. Cell 88: 277–285, 1997. [Abstract] [Google Scholar]
240. O'Riordan E, Orlova TN, Podust VN, Chander PN, Yanagi S, Nakazato M, Hu R, Butt K, Delaney V, Goligorsky MS. Characterization of urinary peptide biomarkers of acute rejection in renal allografts. Am J Transplant 7: 930–940, 2007. [Abstract] [Google Scholar]
241. Oda O, Shinzato T, Ohbayashi K, Takai I, Kunimatsu M, Maeda K, Yamanaka N. Purification and characterization of perlecan fragment in urine of end-stage renal failure patients. Clin Chim Acta 255: 119–132, 1996. [Abstract] [Google Scholar]
242. Ogawa S, Ochi T, Shimada H, Inagaki K, Fujita I, Nii A, Moffat MA, Katragadda M, Violand BN, Arch RH, Masferrer JL. Anti-PDGF-B monoclonal antibody reduces liver fibrosis development. Hepatol Res 40: 1128–1141, 2010. [Abstract] [Google Scholar]
243. Olaso E, Arteta B, Benedicto A, Crende O, Friedman SL. Loss of discoidin domain receptor 2 promotes hepatic fibrosis after chronic carbon tetrachloride through altered paracrine interactions between hepatic stellate cells and liver-associated macrophages. Am J Pathol 179: 2894–2904, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
244. Olsen KC, Sapinoro RE, Kottmann RM, Kulkarni AA, Iismaa SE, Johnson GV, Thatcher TH, Phipps RP, Sime PJ. Transglutaminase 2 and its role in pulmonary fibrosis. Am J Respir Crit Care Med 184: 699–707, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
245. Osborne LR. Animal models of Williams syndrome. Am J Med Genet C Semin Med Genet 154C: 209–219, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
246. Palotie A, Vaisanen P, Ott J, Ryhanen L, Elima K, Vikkula M, Cheah K, Vuorio E, Peltonen L. Predisposition to familial osteoarthrosis linked to type II collagen gene. Lancet 1: 924–927, 1989. [Abstract] [Google Scholar]
247. Paszek MJ, Zahir N, Johnson KR, Lakins JN, Rozenberg GI, Gefen A, Reinhart-King CA, Margulies SS, Dembo M, Boettiger D, Hammer DA, Weaver VM. Tensional homeostasis and the malignant phenotype. Cancer Cell 8: 241–254, 2005. [Abstract] [Google Scholar]
248. Patsenker E, Popov Y, Stickel F, Jonczyk A, Goodman SL, Schuppan D. Inhibition of integrin alphavbeta6 on cholangiocytes blocks transforming growth factor-beta activation and retards biliary fibrosis progression. Gastroenterology 135: 660–670, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
249. Patsenker E, Stickel F. Role of integrins in fibrosing liver diseases. Am J Physiol Gastrointest Liver Physiol 301: G425–G434, 2011. [Abstract] [Google Scholar]
250. Paulsson M, Heinegard D. Purification and structural characterization of a cartilage matrix protein. Biochem J 197: 367–375, 1981. [Europe PMC free article] [Abstract] [Google Scholar]
251. Pedersen JA, Lichter S, Swartz MA. Cells in 3D matrices under interstitial flow: effects of extracellular matrix alignment on cell shear stress and drag forces. J Biomech 43: 900–905, 2010. [Abstract] [Google Scholar]
252. Pedersen JA, Swartz MA. Mechanobiology in the third dimension. Ann Biomed Eng 33: 1469–1490, 2005. [Abstract] [Google Scholar]
253. Peng L, Ran YL, Hu H, Yu L, Liu Q, Zhou Z, Sun YM, Sun LC, Pan J, Sun LX, Zhao P, Yang ZH. Secreted LOXL2 is a novel therapeutic target that promotes gastric cancer metastasis via the Src/FAK pathway. Carcinogenesis 30: 1660–1669, 2009. [Abstract] [Google Scholar]
254. Pfister RR, Haddox JL, Sommers CI, Lam KW. Identification and synthesis of chemotactic tripeptides from alkali-degraded whole cornea. A study of N-acetyl-proline-glycine-proline and N-methyl-proline-glycine-proline. Invest Ophthalmol Vis Sci 36: 1306–1316, 1995. [Abstract] [Google Scholar]
255. Philips GM, Chan IS, Swiderska M, Schroder VT, Guy C, Karaca GF, Moylan C, Venkatraman T, Feuerlein S, Syn WK, Jung Y, Witek RP, Choi S, Michelotti GA, Rangwala F, Merkle E, Lascola C, Diehl AM. Hedgehog signaling antagonist promotes regression of both liver fibrosis and hepatocellular carcinoma in a murine model of primary liver cancer. PLoS One 6: e23943, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
256. Pinzani M. Welcome to fibrogenesis & tissue repair. Fibrogenesis Tissue Repair 1: 1, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
257. Pinzani M, Rombouts K. Liver fibrosis: from the bench to clinical targets. Dig Liver Dis 36: 231–242, 2004. [Abstract] [Google Scholar]
258. Popov Y, Patsenker E, Stickel F, Zaks J, Bhaskar KR, Niedobitek G, Kolb A, Friess H, Schuppan D. Integrin alphavbeta6 is a marker of the progression of biliary and portal liver fibrosis and a novel target for antifibrotic therapies. J Hepatol 48: 453–464, 2008. [Abstract] [Google Scholar]
259. Popov Y, Schuppan D. Targeting liver fibrosis: strategies for development and validation of antifibrotic therapies. Hepatology 50: 1294–1306, 2009. [Abstract] [Google Scholar]
260. Popov Y, Sverdlov DY, Bhaskar KR, Sharma AK, Millonig G, Patsenker E, Krahenbuhl S, Krahenbuhl L, Schuppan D. Macrophage-mediated phagocytosis of apoptotic cholangiocytes contributes to reversal of experimental biliary fibrosis. Am J Physiol Gastrointest Liver Physiol 298: G323–G334, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
261. Popov Y, Sverdlov DY, Sharma AK, Bhaskar KR, Li S, Freitag TL, Lee J, Dieterich W, Melino G, Schuppan D. Tissue transglutaminase does not affect fibrotic matrix stability or regression of liver fibrosis in mice. Gastroenterology 140: 1642–1652, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
262. Poulos JE, Weber JD, Bellezzo JM, di Bisceglie AM, Britton RS, Bacon BR, Baldassare JJ. Fibronectin and cytokines increase JNK, ERK, AP-1 activity, and transin gene expression in rat hepatic stellate cells. Am J Physiol 273: G804–G811, 1997. [Abstract] [Google Scholar]
263. Poynard T, Marcellin P, Lee SS, Niederau C, Minuk GS, Ideo G, Bain V, Heathcote J, Zeuzem S, Trepo C, Albrecht J. Randomised trial of interferon alpha2b plus ribavirin for 48 wk or for 24 wk versus interferon alpha2b plus placebo for 48 wk for treatment of chronic infection with hepatitis C virus. International Hepatitis Interventional Therapy Group (IHIT). Lancet 352: 1426–1432, 1998. [Abstract] [Google Scholar]
264. Poynard T, Ratziu V, McHutchison J, Manns M, Goodman Z, Zeuzem S, Younossi Z, Albrecht J. Effect of treatment with peginterferon or interferon alpha-2b and ribavirin on steatosis in patients infected with hepatitis C. Hepatology 38: 75–85, 2003. [Abstract] [Google Scholar]
265. Ramachandran P, Iredale JP. Reversibility of liver fibrosis. Ann Hepatol 8: 283–291, 2009. [Abstract] [Google Scholar]
266. Ramachandran P, Iredale JP. Macrophages: central regulators of hepatic fibrogenesis and fibrosis resolution. J Hepatol 56: 1417–1419, 2012. [Abstract] [Google Scholar]
267. Ramadori G, Saile B. Portal tract fibrogenesis in the liver. Lab Invest 84: 153–159, 2004. [Abstract] [Google Scholar]
268. Ramchandran R, Dhanabal M, Volk R, Waterman MJ, Segal M, Lu H, Knebelmann B, Sukhatme VP. Antiangiogenic activity of restin, NC10 domain of human collagen XV: comparison to endostatin. Biochem Biophys Res Commun 255: 735–739, 1999. [Abstract] [Google Scholar]
269. Richards AJ, Martin S, Nicholls AC, Harrison JB, Pope FM, Burrows NP. A single base mutation in COL5A2 causes Ehlers-Danlos syndrome type II. J Med Genet 35: 846–848, 1998. [Europe PMC free article] [Abstract] [Google Scholar]
270. Riehle KJ, Campbell JS, McMahan RS, Johnson MM, Beyer RP, Bammler TK, Fausto N. Regulation of liver regeneration and hepatocarcinogenesis by suppressor of cytokine signaling 3. J Exp Med 205: 91–103, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
271. Rodriguez HM, Vaysberg M, Mikels A, McCauley S, Velayo AC, Garcia C, Smith V. Modulation of lysyl oxidase-like 2 enzymatic activity by an allosteric antibody inhibitor. J Biol Chem 285: 20964–20974, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
272. Rubel D, Frese J, Martin M, Leibnitz A, Girgert R, Miosge N, Eckes B, Müller GA, Gross O. Collagen receptors integrin alpha2beta1 and discoidin domain receptor 1 regulate maturation of the glomerular basement membrane and loss of integrin alpha2beta1 delays kidney fibrosis in COL4A3 knockout mice. Matrix Biol 34: 13–21, 2014. [Abstract] [Google Scholar]
273. Ruhl M, Johannsen M, Atkinson J, Manski D, Sahin E, Somasundaram R, Riecken EO, Schuppan D. Soluble collagen VI induces tyrosine phosphorylation of paxillin and focal adhesion kinase and activates the MAP kinase Erk2 in fibroblasts. Exp Cell Res 250: 548–557, 1999. [Abstract] [Google Scholar]
274. Ruhl M, Sahin E, Johannsen M, Somasundaram R, Manski D, Riecken EO, Schuppan D. Soluble collagen VI drives serum-starved fibroblasts through S phase and prevents apoptosis via down-regulation of Bax. J Biol Chem 274: 34361–34368, 1999. [Abstract] [Google Scholar]
275. Ruiz PA, Jarai G. Collagen I induces discoidin domain receptor (DDR) 1 expression through DDR2 and a JAK2-ERK1/2-mediated mechanism in primary human lung fibroblasts. J Biol Chem 286: 12912–12923, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
276. Ruiz PA, Jarai G. Discoidin domain receptors regulate the migration of primary human lung fibroblasts through collagen matrices. Fibrogenesis Tissue Repair 5: 3, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
278. Ruoslahti E. RGD and other recognition sequences for integrins. Annu Rev Cell Dev Biol 12: 697–715, 1996. [Abstract] [Google Scholar]
279. Sawyers CL. The cancer biomarker problem. Nature 452: 548–552, 2008. [Abstract] [Google Scholar]
280. Schaefer L. Small leucine-rich proteoglycans in kidney disease. J Am Soc Nephrol: 1200–1207, 2011. [Abstract] [Google Scholar]
281. Schaefer L, Babelova A, Kiss E, Hausser HJ, Baliova M, Krzyzankova M, Marsche G, Young MF, Mihalik D, Götte M, Malle E, Schaefer RM, Gröne HJ. The matrix component biglycan is proinflammatory and signals through Toll-like receptors 4 and 2 in macrophages. J Clin Invest 115: 2223–2233, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
282. Schaefer L, Tsalastra W, Babelova A, Baliova M, Minnerup J, Sorokin L, Gröne HJ, Reinhardt DP, Pfeilschifter J, Iozzo RV, Schaefer RM. Decorin-mediated regulation of fibrillin-1 in the kidney involves the insulin-like growth factor-I receptor and Mammalian target of rapamycin. Am J Pathol 170: 301–315, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
283. Schrader J, Gordon-Walker TT, Aucott RL, Van DM, Quaas A, Walsh S, Benten D, Forbes SJ, Wells RG, Iredale JP. Matrix stiffness modulates proliferation, chemotherapeutic response, and dormancy in hepatocellular carcinoma cells. Hepatology 53: 1192–1205, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
284. Schupp JC, Binder H, Jager B, Cillis G, Zissel G, Muller-Quernheim J, Prasse A. Macrophage activation in acute exacerbation of idiopathic pulmonary fibrosis. PLoS One 10: e0116775, 2015. [Europe PMC free article] [Abstract] [Google Scholar]
285. Schuppan D, Afdhal NH. Liver cirrhosis. Lancet 371: 838–851, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
286. Schuppan D, Cramer T, Bauer M, Strefeld T, Hahn EG, Herbst H. Hepatocytes as a source of collagen type XVIII endostatin. Lancet 352: 879–880, 1998. [Abstract] [Google Scholar]
287. Schuppan D, Kim YO. Evolving therapies for liver fibrosis. J Clin Invest 123: 1887–1901, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
288. Schuppan D, Pinzani M. Anti-fibrotic therapy: lost in translation? J Hepatol 56, Suppl 1: S66–S74, 2012. [Abstract] [Google Scholar]
289. Schuppan D, Ruehl M, Somasundaram R, Hahn EG. Matrix as a modulator of hepatic fibrogenesis. Semin Liver Dis 21: 351–372, 2001. [Abstract] [Google Scholar]
290. Schuppan D, Schattenberg JM. Non-alcoholic steatohepatitis: pathogenesis and novel therapeutic approaches. J Gastroenterol Hepatol 28, Suppl 1: 68–76, 2013. [Abstract] [Google Scholar]
291. Schuppan D, Schmid M, Somasundaram R, Ackermann R, Ruehl M, Nakamura T, Riecken EO. Collagens in the liver extracellular matrix bind hepatocyte growth factor. Gastroenterology 114: 139–152, 1998. [Abstract] [Google Scholar]
292. Schwabe RF, Schnabl B, Kweon YO, Brenner DA. CD40 activates NF-kappa B and c-Jun N-terminal kinase and enhances chemokine secretion on activated human hepatic stellate cells. J Immunol 166: 6812–6819, 2001. [Abstract] [Google Scholar]
293. Schwartz JM, Reinus JF. Prevalence and natural history of alcoholic liver disease. Clin Liver Dis 16: 659–666, 2012. [Abstract] [Google Scholar]
294. Schwartz MA, Schaller MD, Ginsberg MH. Integrins: emerging paradigms of signal transduction. Annu Rev Cell Dev Biol 11: 549–599, 1995. [Abstract] [Google Scholar]
295. Schwarzbauer JE, DeSimone DW. Fibronectins, their fibrillogenesis, and in vivo functions. Cold Spring Harb Perspect Biol 3: a005041, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
296. Schymeinsky J, Nedbal S, Miosge N, Poschl E, Rao C, Beier DR, Skarnes WC, Timpl R, Bader BL. Gene structure and functional analysis of the mouse nidogen-2 gene: nidogen-2 is not essential for basement membrane formation in mice. Mol Cell Biol 22: 6820–6830, 2002. [Europe PMC free article] [Abstract] [Google Scholar]
297. Segovia-Silvestre T, Reichenbach V, Fernandez-Varo G, Vassiliadis E, Barascuk N, Morales-Ruiz M, Karsdal MA, Jiménez W. Circulating CO3-610, a degradation product of collagen III, closely reflects liver collagen and portal pressure in rats with fibrosis. Fibrogenesis Tissue Repair 4: 19, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
298. Seidler DG, Goldoni S, Agnew C, Cardi C, Thakur ML, Owens RT, McQuillan DJ, Iozzo RV. Decorin protein core inhibits in vivo cancer growth and metabolism by hindering epidermal growth factor receptor function and triggering apoptosis via caspase-3 activation. J Biol Chem 281: 26408–26418, 2006. [Abstract] [Google Scholar]
299. Seitz HK, Stickel F. Risk factors and mechanisms of hepatocarcinogenesis with special emphasis on alcohol and oxidative stress. Biol Chem 387: 349–360, 2006. [Abstract] [Google Scholar]
300. Seki E, De Minicis S, Gwak GY, Kluwe J, Inokuchi S, Bursill CA, Llovet JM, Brenner DA, Schwabe RF. CCR1 and CCR5 promote hepatic fibrosis in mice. J Clin Invest 119: 1858–1870, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
301. Seki E, De Minicis S, Inokuchi S, Taura K, Miyai K, Van Rooijen N, Schwabe RF, Brenner DA. CCR2 promotes hepatic fibrosis in mice. Hepatology 50: 185–197, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
302. Seppinen L, Pihlajaniemi T. The multiple functions of collagen XVIII in development and disease. Matrix Biol 30: 83–92, 2011. [Abstract] [Google Scholar]
303. Shapiro JR, Mcbride DJ Jr, Fedarko NS. OIM and related animal models of osteogenesis imperfecta. Connect Tissue Res 31: 265–268, 1995. [Abstract] [Google Scholar]
304. Shariff MI, Cox IJ, Gomaa AI, Khan SA, Gedroyc W, Taylor-Robinson SD. Hepatocellular carcinoma: current trends in worldwide epidemiology, risk factors, diagnosis and therapeutics. Expert Rev Gastroenterol Hepatol 3: 353–367, 2009. [Abstract] [Google Scholar]
305. Shi H, Huang Y, Zhou H, Song X, Yuan S, Fu Y, Luo Y. Nucleolin is a receptor that mediates antiangiogenic and antitumor activity of endostatin. Blood 110: 2899–2906, 2007. [Abstract] [Google Scholar]
306. Shute J. Glycosaminoglycan and chemokine/growth factor interactions. Hand Exp Pharmacol 2012: 307–324, 2012. [Abstract] [Google Scholar]
307. Sicklick JK, Li YX, Melhem A, Schmelzer E, Zdanowicz M, Huang J, Caballero M, Fair JH, Ludlow JW, McClelland RE, Reid LM, Diehl AM. Hedgehog signaling maintains resident hepatic progenitors throughout life. Am J Physiol Gastrointest Liver Physiol 290: G859–G870, 2006. [Abstract] [Google Scholar]
308. Siegel M, Strnad P, Watts RE, Choi K, Jabri B, Omary MB, Khosla C. Extracellular transglutaminase 2 is catalytically inactive, but is transiently activated upon tissue injury. PLoS One 3: e1861, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
309. Singh KP, Gerard HC, Hudson AP, Boros DL. Dynamics of collagen, MMP and TIMP gene expression during the granulomatous, fibrotic process induced by Schistosoma mansoni eggs. Ann Trop Med Parasitol 98: 581–593, 2004. [Abstract] [Google Scholar]
310. Singh P, Schwarzbauer JE. Fibronectin and stem cell differentiation - lessons from chondrogenesis. J Cell Sci 125: 3703–3712, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
311. Sjoberg AP, Manderson GA, Morgelin M, Day AJ, Heinegard D, Blom AM. Short leucine-rich glycoproteins of the extracellular matrix display diverse patterns of complement interaction and activation. Mol Immunol 46: 830–839, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
312. Smith LB, Hadoke PW, Dyer E, Denvir MA, Brownstein D, Miller E, Nelson N, Wells S, Cheeseman M, Greenfield A. Haploinsufficiency of the murine Col3a1 locus causes aortic dissection: a novel model of the vascular type of Ehlers-Danlos syndrome. Cardiovasc Res 90: 182–190, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
313. Sobrevals L, Rodriguez C, Romero-Trevejo JL, Gondi G, Monreal I, Paneda A, Juanarena N, Arcelus S, Razquin N, Guembe L, González-Aseguinolaza G, Prieto J, Fortes P. Insulin-like growth factor I gene transfer to cirrhotic liver induces fibrolysis and reduces fibrogenesis leading to cirrhosis reversion in rats. Hepatology 51: 912–921, 2010. [Abstract] [Google Scholar]
314. Sofat N, Robertson SD, Hermansson M, Jones J, Mitchell P, Wait R. Tenascin-C fragments are endogenous inducers of cartilage matrix degradation. Rheumatol Int 32: 2809–2817, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
315. Somasundaram R, Ruehl M, Schaefer B, Schmid M, Ackermann R, Riecken EO, Zeitz M, Schuppan D. Interstitial collagens I, III, and VI sequester and modulate the multifunctional cytokine oncostatin M. J Biol Chem 277: 3242–3246, 2002. [Abstract] [Google Scholar]
316. Somasundaram R, Ruehl M, Tiling N, Ackermann R, Schmid M, Riecken EO, Schuppan D. Collagens serve as an extracellular store of bioactive interleukin 2. J Biol Chem 275: 38170–38175, 2000. [Abstract] [Google Scholar]
317. Somasundaram R, Schuppan D. Type I, II, III, IV, V, and VI collagens serve as extracellular ligands for the isoforms of platelet-derived growth factor (AA, BB, and AB). J Biol Chem 271: 26884–26891, 1996. [Abstract] [Google Scholar]
318. Song T, Choi CH, Cho YJ, Sung CO, Song SY, Kim TJ, Bae DS, Lee JW, Kim BG. Expression of 67-kDa laminin receptor was associated with tumor progression and poor prognosis in epithelial ovarian cancer. Gynecol Oncol 125: 427–432, 2012. [Abstract] [Google Scholar]
319. Sorensen TI, Orholm M, Bentsen KD, Hoybye G, Eghoje K, Christoffersen P. Prospective evaluation of alcohol abuse and alcoholic liver injury in men as predictors of development of cirrhosis. Lancet 2: 241–244, 1984. [Abstract] [Google Scholar]
320. Spee B, Carpino G, Schotanus BA, Katoonizadeh A, Vander BS, Gaudio E, Roskams T. Characterisation of the liver progenitor cell niche in liver diseases: potential involvement of Wnt and Notch signalling. Gut 59: 247–257, 2010. [Abstract] [Google Scholar]
321. Spickett CM, Pitt AR, Morrice N, Kolch W. Proteomic analysis of phosphorylation, oxidation and nitrosylation in signal transduction. Biochim Biophys Acta 1764: 1823–1841, 2006. [Abstract] [Google Scholar]
322. Steubl D, Hettwer S, Vrijbloed W, Dahinden P, Wolf P, Luppa P, Wagner CA, Renders L, Heemann U, Roos M. C-terminal agrin fragment—a new fast biomarker for kidney function in renal transplant recipients. Am J Nephrol 38: 501–508, 2013. [Abstract] [Google Scholar]
323. Stone JH, Zen Y, Deshpande V. IgG4-related disease. N Engl J Med 366: 539–551, 2012. [Abstract] [Google Scholar]
324. Stum M, Girard E, Bangratz M, Bernard V, Herbin M, Vignaud A, Ferry A, Davoine CS, Echaniz-Laguna A, René F, Marcel C, Molgó J, Fontaine B, Krejci E, Nicole S. Evidence of a dosage effect and a physiological endplate acetylcholinesterase deficiency in the first mouse models mimicking Schwartz-Jampel syndrome neuromyotonia. Hum Mol Genet 17: 3166–3179, 2008. [Abstract] [Google Scholar]
325. Svegliati-Baroni G, Ridolfi F, Di Sario A, Casini A, Marucci L, Gaggiotti G, Orlandoni P, Macarri G, Perego L, Benedetti A, Folli F. Insulin and insulin-like growth factor-1 stimulate proliferation and type I collagen accumulation by human hepatic stellate cells: differential effects on signal transduction pathways. Hepatology 29: 1743–1751, 1999. [Abstract] [Google Scholar]
326. Svensson L, Aszodi A, Reinholt FP, Fassler R, Heinegard D, Oldberg A. Fibromodulin-null mice have abnormal collagen fibrils, tissue organization, and altered lumican deposition in tendon. J Biol Chem 274: 9636–9647, 1999. [Abstract] [Google Scholar]
327. Syal G, Fausther M, Dranoff JA. Advances in cholangiocyte immunobiology. Am J Physiol Gastrointest Liver Physiol 303: G1077–G1086, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
328. Sykes B, Ogilvie D, Wordsworth P, Wallis G, Mathew C, Beighton P, Nicholls A, Pope FM, Thompson E, Tsipouras P, Schwartz R, Jensson O, Arnason A, Børresen AL, Heiberg A, Frey D, Steinmann B. Consistent linkage of dominantly inherited osteogenesis imperfecta to the type I collagen loci: COL1A1 and COL1A2. Am J Hum Genet 46: 293–307, 1990. [Europe PMC free article] [Abstract] [Google Scholar]
329. Tao LH, Enzan H, Hayashi Y, Miyazaki E, Saibara T, Hiroi M, Toi M, Kuroda N, Naruse K, Jin YL, Guo LM. Appearance of denuded hepatic stellate cells and their subsequent myofibroblast-like transformation during the early stage of biliary fibrosis in the rat. Med Electron Microsc 33: 217–230, 2000. [Abstract] [Google Scholar]
330. Tarrats N, Moles A, Morales A, Garcia-Ruiz C, Fernandez-Checa JC, Mari M. Critical role of tumor necrosis factor receptor 1, but not 2, in hepatic stellate cell proliferation, extracellular matrix remodeling, and liver fibrogenesis. Hepatology 54: 319–327, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
331. Taylor SH, Al-Youha S, Van AT, Lu Y, Wong J, McGrouther DA, Kadler KE. Tendon is covered by a basement membrane epithelium that is required for cell retention and the prevention of adhesion formation. PLoS One 6: e16337, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
332. Thomas JA, Pope C, Wojtacha D, Robson AJ, Gordon-Walker TT, Hartland S, Ramachandran P, Van Deemter M, Hume DA, Iredale JP, Forbes SJ. Macrophage therapy for murine liver fibrosis recruits host effector cells improving fibrosis, regeneration, and function. Hepatology 53: 2003–2015, 2011. [Abstract] [Google Scholar]
333. Tirnitz-Parker JE, Viebahn CS, Jakubowski A, Klopcic BR, Olynyk JK, Yeoh GC, Knight B. Tumor necrosis factor-like weak inducer of apoptosis is a mitogen for liver progenitor cells. Hepatology 52: 291–302, 2010. [Abstract] [Google Scholar]
334. Torok N, Dranoff JA, Schuppan D, Friedman SL. Strategies and endpoints of antifibrotic drug trials: Summary and recommendations from the AASLD Emerging Trends Conference, Chicago, June 2014. Hepatology. In press. [Europe PMC free article] [Abstract] [Google Scholar]
335. Treeprasertsuk S, Kowdley KV, Luketic VA, Harrison ME, McCashland T, Befeler AS, Harnois D, Jorgensen R, Petz J, Keach J, Schmoll J, Hoskin T, Thapa P, Enders F, Lindor KD. The predictors of the presence of varices in patients with primary sclerosing cholangitis. Hepatology 51: 1302–1310, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
336. Trentham DE, Townes AS, Kang AH. Autoimmunity to type II collagen an experimental model of arthritis. J Exp Med 146: 857–868, 1977. [Europe PMC free article] [Abstract] [Google Scholar]
337. Troeger JS, Mederacke I, Gwak GY, Dapito DH, Mu X, Hsu CC, Pradere JP, Friedman RA, Schwabe RF. Deactivation of hepatic stellate cells during liver fibrosis resolution in mice. Gastroenterology 143: 1073–1083, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
338. Tromp G, Kuivaniemi H, Stolle C, Pope FM, Prockop DJ. Single base mutation in the type III procollagen gene that converts the codon for glycine 883 to aspartate in a mild variant of Ehlers-Danlos syndrome IV. J Biol Chem 264: 19313–19317, 1989. [Abstract] [Google Scholar]
339. Tsuchiya A, Lu WY, Weinhold B, Boulter L, Stutchfield BM, Williams MJ, Guest RV, Minnis-Lyons SE, MacKinnon AC, Schwarzer D, Ichida T, Nomoto M, Aoyagi Y, Gerardy-Schahn R, Forbes SJ. Polysialic acid/neural cell adhesion molecule modulates the formation of ductular reactions in liver injury. Hepatology 60: 1727–1740, 2014. [Abstract] [Google Scholar]
340. Tu T, Budzinska MA, Maczurek AE, Cheng R, Di BA, Warner FJ, McCaughan GW, McLennan SV, Shackel NA. Novel aspects of the liver microenvironment in hepatocellular carcinoma pathogenesis and development. Int J Mol Sci 15: 9422–9458, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
341. Tufvesson E, Westergren-Thorsson G. Tumour necrosis factor-alpha interacts with biglycan and decorin. FEBS Lett 530: 124–128, 2002. [Abstract] [Google Scholar]
342. Tufvesson E, Westergren-Thorsson G. Biglycan and decorin induce morphological and cytoskeletal changes involving signalling by the small GTPases RhoA and Rac1 resulting in lung fibroblast migration. J Cell Sci 116: 4857–4864, 2003. [Abstract] [Google Scholar]
343. Tzortzaki EG, Tischfield JA, Sahota A, Siafakas NM, Gordon MK, Gerecke DR. Expression of FACIT collagens XII and XIV during bleomycin-induced pulmonary fibrosis in mice. Anat Rec A Discov Mol Cell Evol Biol 275: 1073–1080, 2003. [Abstract] [Google Scholar]
344. Utriainen A, Sormunen R, Kettunen M, Carvalhaes LS, Sajanti E, Eklund L, Kauppinen R, Kitten GT, Pihlajaniemi T. Structurally altered basement membranes and hydrocephalus in a type XVIII collagen deficient mouse line. Hum Mol Genet 13: 2089–2099, 2004. [Abstract] [Google Scholar]
345. Valiathan RR, Marco M, Leitinger B, Kleer CG, Fridman R. Discoidin domain receptor tyrosine kinases: new players in cancer progression. Cancer Metastasis Rev 31: 295–321, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
346. van der Weyden L, Wei L, Luo J, Yang X, Birk DE, Adams DJ, Bradley A, Chen Q. Functional knockout of the matrilin-3 gene causes premature chondrocyte maturation to hypertrophy and increases bone mineral density and osteoarthritis. Am J Pathol 169: 515–527, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
347. Van Agtmael T, Bailey MA, Schlotzer-Schrehardt U, Craigie E, Jackson IJ, Brownstein DG, Megson IL, Mullins JJ. Col4a1 mutation in mice causes defects in vascular function and low blood pressure associated with reduced red blood cell volume. Hum Mol Genet 19: 1119–1128, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
348. Van Agtmael T, Bruckner-Tuderman L. Basement membranes and human disease. Cell Tissue Res 339: 167–188, 2010. [Abstract] [Google Scholar]
349. Vannella KM, Barron L, Borthwick LA, Kindrachuk KN, Narasimhan PB, Hart KM, Thompson RW, White S, Cheever AW, Ramalingam TR, Wynn TA. Incomplete deletion of IL-4Ralpha by LysM(Cre) reveals distinct subsets of M2 macrophages controlling inflammation and fibrosis in chronic schistosomiasis. PLoS Pathog 10: e1004372, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
350. Vassiliadis E, Larsen DV, Clausen RE, Veidal SS, Barascuk N, Larsen L, Simonsen H, Silvestre TS, Hansen C, Overgaard T, Leeming DJ, Karsdal MA. Measurement of CO3-610, a potential liver biomarker derived from matrix metalloproteinase-9 degradation of collagen type III, in a rat model of reversible carbon-tetrachloride-induced fibrosis. Biomark Insights 6: 49–58, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
351. Veidal SS, Karsdal MA, Nawrocki A, Larsen MR, Dai Y, Zheng Q, Hägglund P, Vainer B, Skjøt-Arkil H, Leeming DJ. Assessment of proteolytic degradation of the basement membrane: A fragment of type IV collagen as a biochemical marker for liver fibrosis. Fibrogenesis Tissue Repair 4: 22, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
352. Veidal SS, Larsen DV, Chen X, Sun S, Zheng Q, Bay-Jensen AC, Leeming DJ, Nawrocki A, Larsen MR, Schett G, Karsdal MA. MMP mediated type V collagen degradation (C5M) is elevated in ankylosing spondylitis. Clin Biochem 45: 541–546, 2012. [Abstract] [Google Scholar]
353. Veidal SS, Nielsen MJ, Leeming DJ, Karsdal MA. Phosphodiesterase inhibition mediates matrix metalloproteinase activity and the level of collagen degradation fragments in a liver fibrosis ex vivo rat model. BMC Res Notes 5: 686, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
354. Verderio EA, Johnson T, Griffin M. Tissue transglutaminase in normal and abnormal wound healing: review article. Amino Acids 26: 387–404, 2004. [Abstract] [Google Scholar]
355. Vij N, Roberts L, Joyce S, Chakravarti S. Lumican suppresses cell proliferation and aids Fas-Fas ligand mediated apoptosis: implications in the cornea. Exp Eye Res 78: 957–971, 2004. [Abstract] [Google Scholar]
356. Vogten JM, Drixler TA, te Velde EA, Schipper ME, van Vroonhoven TJ, Voest EE, Borel Rinkes IH. Angiostatin inhibits experimental liver fibrosis in mice. Int J Colorectal Dis 19: 387–394, 2004. [Abstract] [Google Scholar]
357. Wadhwa S, Embree M, Ameye L, Young MF. Mice deficient in biglycan and fibromodulin as a model for temporomandibular joint osteoarthritis. Cells Tissues Organs 181: 136–143, 2005. [Abstract] [Google Scholar]
358. Wan YY, Tian GY, Guo HS, Kang YM, Yao ZH, Li XL, Lin DJ. Endostatin, an angiogenesis inhibitor, ameliorates bleomycin-induced pulmonary fibrosis in rats. Respir Res 14: 56, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
359. Wang B, Ding YM, Fan P, Wang B, Xu JH, Wang WX. Expression and significance of MMP2 and HIF-1alpha in hepatocellular carcinoma. Oncol Lett 8: 539–546, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
360. Wang J, Leclercq I, Brymora JM, Xu N, Ramezani-Moghadam M, London RM, Brigstock D, George J. Kupffer cells mediate leptin-induced liver fibrosis. Gastroenterology 137: 713–723, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
361. Wang J, Su H, Han X, Xu K. Inhibition of fibroblast growth factor receptor signaling impairs metastasis of hepatocellular carcinoma. Tumour Biol 35: 11005–11011, 2014. [Abstract] [Google Scholar]
362. Wang Q, Wang Y, Hyde DM, Gotwals PJ, Koteliansky VE, Ryan ST, Giri SN. Reduction of bleomycin induced lung fibrosis by transforming growth factor beta soluble receptor in hamsters. Thorax 54: 805–812, 1999. [Europe PMC free article] [Abstract] [Google Scholar]
363. Wang XQ, Frazier WA. The thrombospondin receptor CD47 (IAP) modulates and associates with alpha2 beta1 integrin in vascular smooth muscle cells. Mol Biol Cell 9: 865–874, 1998. [Europe PMC free article] [Abstract] [Google Scholar]
364. Wang Y, Ma Y, Lu B, Xu E, Huang Q, Lai M. Differential expression of mimecan and thioredoxin domain-containing protein 5 in colorectal adenoma and cancer: a proteomic study. Exp Biol Med (Maywood) 232: 1152–1159, 2007. [Abstract] [Google Scholar]
365. Wansink DG, Wieringa B. Transgenic mouse models for myotonic dystrophy type 1 (DM1). Cytogenet Genome Res 100: 230–242, 2003. [Abstract] [Google Scholar]
366. Wasmuth HE, Lammert F, Zaldivar MM, Weiskirchen R, Hellerbrand C, Scholten D, Berres ML, Zimmermann H, Streetz KL, Tacke F, Hillebrandt S, Schmitz P, Keppeler H, Berg T, Dahl E, Gassler N, Friedman SL, Trautwein C. Antifibrotic effects of CXCL9 and its receptor CXCR3 in livers of mice and humans. Gastroenterology 137: 309–319, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
367. Weathington NM, van Houwelingen AH, Noerager BD, Jackson PL, Kraneveld AD, Galin FS, Folkerts G, Nijkamp FP, Blalock JE. A novel peptide CXCR ligand derived from extracellular matrix degradation during airway inflammation. Nat Med 12: 317–323, 2006. [Abstract] [Google Scholar]
368. Weil D, D'Alessio M, Ramirez F, Eyre DR. Structural and functional characterization of a splicing mutation in the pro-alpha 2(I) collagen gene of an Ehlers-Danlos type VII patient. J Biol Chem 265: 16007–16011, 1990. [Abstract] [Google Scholar]
369. Wenstrup RJ, Florer JB, Willing MC, Giunta C, Steinmann B, Young F, Susic M, Cole WG. COL5A1 haploinsufficiency is a common molecular mechanism underlying the classical form of EDS. Am J Hum Genet 66: 1766–1776, 2000. [Europe PMC free article] [Abstract] [Google Scholar]
370. Westergren-Thorsson G, Hernnas J, Sarnstrand B, Oldberg A, Heinegard D, Malmstrom A. Altered expression of small proteoglycans, collagen, and transforming growth factor-beta 1 in developing bleomycin-induced pulmonary fibrosis in rats. J Clin Invest 92: 632–637, 1993. [Europe PMC free article] [Abstract] [Google Scholar]
371. Whiteford JR, Xian X, Chaussade C, Vanhaesebroeck B, Nourshargh S, Couchman JR. Syndecan-2 is a novel ligand for the protein tyrosine phosphatase receptor CD148. Mol Biol Cell 22: 3609–3624, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
372. Wilhelm A, Munir M, Humphreys E, Adams D, Burkly L, Afford S, Weston C. The TWEAK and Fn14 pathway as potential mediator of liver fibrosis. Gut 63, Suppl 1: A16–A17, 2014. [Google Scholar]
373. Woelfle JV, Brenner RE, Zabel B, Reichel H, Nelitz M. Schmid-type metaphyseal chondrodysplasia as the result of a collagen type X defect due to a novel COL10A1 nonsense mutation: A case report of a novel COL10A1 mutation. J Orthop Sci 16: 245–249, 2011. [Abstract] [Google Scholar]
374. Wong CC, Tse AP, Huang YP, Zhu YT, Chiu DK, Lai RK, Au SL, Kai AK, Lee JM, Wei LL, Tsang FH, Lo RC, Shi J, Zheng YP, Wong CM, Ng IO. Lysyl oxidase-like 2 is critical to tumor microenvironment and metastatic niche formation in hepatocellular carcinoma. Hepatology 60: 1645–1658, 2014. [Abstract] [Google Scholar]
375. Wong L, Yamasaki G, Johnson RJ, Friedman SL. Induction of beta-platelet-derived growth factor receptor in rat hepatic lipocytes during cellular activation in vivo and in culture. J Clin Invest 94: 1563–1569, 1994. [Europe PMC free article] [Abstract] [Google Scholar]
376. Woodall BP, Nystrom A, Iozzo RA, Eble JA, Niland S, Krieg T, Eckes B, Pozzi A, Iozzo RV. Integrin alpha2beta1 is the required receptor for endorepellin angiostatic activity. J Biol Chem 283: 2335–2343, 2008. [Abstract] [Google Scholar]
377. Wu F, Vij N, Roberts L, Lopez-Briones S, Joyce S, Chakravarti S. A novel role of the lumican core protein in bacterial lipopolysaccharide-induced innate immune response. J Biol Chem 282: 26409–26417, 2007. [Abstract] [Google Scholar]
378. Wynn TA. Common and unique mechanisms regulate fibrosis in various fibroproliferative diseases. J Clin Invest 117: 524–529, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
379. Wynn TA. Cellular and molecular mechanisms of fibrosis. J Pathol 214: 199–210, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
380. Wynn TA, Barron L. Macrophages: master regulators of inflammation and fibrosis. Semin Liver Dis 30: 245–257, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
381. Wynn TA, Chawla A, Pollard JW. Macrophage biology in development, homeostasis and disease. Nature 496: 445–455, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
382. Xia JL, Dai C, Michalopoulos GK, Liu Y. Hepatocyte growth factor attenuates liver fibrosis induced by bile duct ligation. Am J Pathol 168: 1500–1512, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
383. Xu H, Bihan D, Chang F, Huang PH, Farndale RW, Leitinger B. Discoidin domain receptors promote alpha1beta1- and alpha2beta1-integrin mediated cell adhesion to collagen by enhancing integrin activation. PLoS One 7: e52209, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
384. Xu H, Raynal N, Stathopoulos S, Myllyharju J, Farndale RW, Leitinger B. Collagen binding specificity of the discoidin domain receptors: binding sites on collagens II and III and molecular determinants for collagen IV recognition by DDR1. Matrix Biol 30: 16–26, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
385. Xu J, Lu W, Zhang S, Zhu C, Ren T, Zhu T, Zhao H, Liu Y, Su J. Overexpression of DDR2 contributes to cell invasion and migration in head and neck squamous cell carcinoma. Cancer Biol Ther 15: 612–622, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
386. Xu R, Yao ZY, Xin L, Zhang Q, Li TP, Gan RB. NC1 domain of human type VIII collagen (alpha 1) inhibits bovine aortic endothelial cell proliferation and causes cell apoptosis. Biochem Biophys Res Commun 289: 264–268, 2001. [Abstract] [Google Scholar]
387. Xu T, Bianco P, Fisher LW, Longenecker G, Smith E, Goldstein S, Bonadio J, Boskey A, Heegaard AM, Sommer B, Satomura K, Dominguez P, Zhao C, Kulkarni AB, Robey PG, Young MF. Targeted disruption of the biglycan gene leads to an osteoporosis-like phenotype in mice. Nat Genet 20: 78–82, 1998. [Abstract] [Google Scholar]
388. Yamaguchi Y, Takihara T, Chambers RA, Veraldi KL, Larregina AT, Feghali-Bostwick CA. A peptide derived from endostatin ameliorates organ fibrosis. Sci Transl Med 4: 136ra71, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
389. Yang JJ, Tao H, Li J. Hedgehog signaling pathway as key player in liver fibrosis: new insights and perspectives. Expert Opin Ther Targets 18: 1011–1021, 2014. [Abstract] [Google Scholar]
390. Yeh LK, Liu CY, Kao WW, Huang CJ, Hu FR, Chien CL, Wang IJ. Knockdown of zebrafish lumican gene (zlum) causes scleral thinning and increased size of scleral coats. J Biol Chem 285: 28141–28155, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
391. Yi M, Ruoslahti E. A fibronectin fragment inhibits tumor growth, angiogenesis, and metastasis. Proc Natl Acad Sci USA 98: 620–624, 2001. [Europe PMC free article] [Abstract] [Google Scholar]
392. Young MF, Bi Y, Ameye L, Chen XD. Biglycan knockout mice: new models for musculoskeletal diseases. Glycoconj J 19: 257–262, 2002. [Abstract] [Google Scholar]
393. Younossi ZM, Stepanova M, Afendy M, Fang Y, Younossi Y, Mir H, Srishord M. Changes in the prevalence of the most common causes of chronic liver diseases in the United States from 1988 to 2008. Clin Gastroenterol Hepatol 9: 524–530, 2011. [Abstract] [Google Scholar]
394. Yurchenco PD. Basement membranes: cell scaffoldings and signaling platforms. Cold Spring Harb Perspect Biol 3: a004911, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
395. Zeremski M, Dimova R, Astemborski J, Thomas DL, Talal AH. CXCL9 and CXCL10 chemokines as predictors of liver fibrosis in a cohort of primarily African-American injection drug users with chronic hepatitis C. J Infect Dis 204: 832–836, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
396. Zhang LJ, Zheng WD, Chen YX, Huang YH, Chen ZX, Zhang SJ, Shi MN, Wang XZ. Antifibrotic effects of interleukin-10 on experimental hepatic fibrosis. Hepatogastroenterology 54: 2092–2098, 2007. [Abstract] [Google Scholar]
397. Zhao Y, Wang Y, Wang Q, Liu Z, Liu Q, Deng X. Hepatic stellate cells produce vascular endothelial growth factor via phospho-p44/42 mitogen-activated protein kinase/cyclooxygenase-2 pathway. Mol Cell Biochem 359: 217–223, 2012. [Abstract] [Google Scholar]
398. Zimmermann HW, Seidler S, Gassler N, Nattermann J, Luedde T, Trautwein C, Tacke F. Interleukin-8 is activated in patients with chronic liver diseases and associated with hepatic macrophage accumulation in human liver fibrosis. PLoS One 6: e21381, 2011. [Europe PMC free article] [Abstract] [Google Scholar]

Articles from American Journal of Physiology - Gastrointestinal and Liver Physiology are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/3808302
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/3808302

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1152/ajpgi.00447.2014

Supporting
Mentioning
Contrasting
2
168
0

Article citations


Go to all (131) article citations

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.