• Non ci sono risultati.

Extracellular MatrixAlex Y. Hui, Scott L. Friedman

N/A
N/A
Protected

Academic year: 2022

Condividi "Extracellular MatrixAlex Y. Hui, Scott L. Friedman"

Copied!
12
0
0

Testo completo

(1)

Extracellular Matrix

Alex Y. Hui, Scott L . Friedman

6

6.1

Introduction

The hepatic extracellular matrix (ECM) is a complex network of macromolecules that not only provides cells with an extracellular scaffold but also plays an important role in the regulation of cellular activities (Fig. 6.1) [50]. In normal liver, the ECM comprises less than 3% of the relative area on a tissue section and approximately 0.5% of the wet weight [13, 21]. In addition to Glisson’s capsule, ECM is found main- ly in the portal tracts and the central veins. Small amounts of ECM, the perisinusoidal matrix, are also found in the space of Disse. The sinusoids are lined by fenestrated endothelial cells, which lack an underlying basement membrane (BM). This facili- tates the flow of plasma between sinusoidal lumen and the hepatocytes, and vice versa. The strategic position of the perisinusoidal matrix at the inter- face between blood and the epithelial components of the liver explains why quantitative or qualitative change of ECM may significantly influence hepatic function [3].

Generation of ECM, or fibrogenesis, occurs in re- sponse to different chronic injuries to the liver. This is a wound-healing response, which is reversible by ECM degradation upon elimination of the primary insult. Transition of a normal to a fibrotic liver in- volves both quantitative and compositional changes in ECM [16]. Intense research over the past 15 years has established that hepatic stellate cells (HSC) play a central role in the process (Fig. 6.2; see Part 1, Chap. 3). Other cell types including hepatocytes and hepatic sinusoidal cells have at most a modest contribution to the production of ECM (see Part 1, Chaps. 1 and 5).

6.2

Components of the ECM in Liver

The components of ECM in liver include collagens, non-collagenous glycoproteins, glycosaminogly- cans, proteoglycans, matrix-bound growth factors (Fig. 6.3) and matricellular proteins. In normal liv- er, the dense, interstitial ECM is largely confined to the capsule, around large vessels and in the portal triad. The perisinusoidal matrix, on the other hand, is composed of both an interstitial and a BM-like low-density ECM.

6.2.1

The Collagen Scaffold

More than 20 genetically distinct collagens have been

identified and are grouped into two main molecular

classes: the relatively homogeneous group of fibril-

forming collagens (collagens I, II, III, V and XI) and

the rather heterogeneous group of non-fibrillar col-

lagens. The fibril-forming collagens, which consist

of a triple helix of approximately 300 nm in length

and 1.5 nm in diameter, self-assemble into fibrils

in the extracellular space through the cleavage of

terminal procollagen peptides by C-propeptidase

and N-propeptidase. Types I, III and V are the main

components in the dense interstitial ECM in portal

tract and central vein wall of normal liver. Among

the non-fibrillar collagens, types IV, VI, VIII, XIV,

XIX, XV and XVIII are found in the liver with dif-

ferent locations and functions. Both fibril-forming

and non-fibrillar collagens are found in the perisi-

nusoidal matrix. These include fibrillar types I, III

and V, microfibrillar collagen VI, BM collagens IV

and XVIII, and FACIT (fibril-associated collagens

with interrupted triple helices) collagen [9, 50].

(2)

6.2.2

Proteoglycans

Proteoglycans belong to a distinct subset of non-collagenous glycoproteins that contain gly- cosaminoglycan (GAG) side chains. They interact with other ECM molecules via specific GAG-bind- ing domains in these molecules. By virtue of such properties, they regulate matrix architecture and spatial arrangement of structural polymers. They bind cytokines and growth factors and thus control their availability and biological activities [9].

Proteoglycans identified in liver so far include aggrecan, fibromodulin, decorin, biglycan, perle- can, betaglycan, glypicans and syndecan-1, -2, -3 and -4. Aggrecan belongs to the family of proteogly- cans characterized by an N-terminal globular do- main that interacts with hyaluronan, and a C-termi-

nal selectin domain. It is found in the interstitium with structural function.

Fibromodulin, decorin and biglycan are char- acterized by a protein core composed of leucine- rich repeats. These provide a horseshoe-like struc- ture, which favors protein–protein interaction. In fact, these small proteoglycans bind transforming growth factor- β1 (TGF-β1), a potent fibrogenic cy- tokine to HSC. In normal liver tissue, biglycan and decorin are detected in the space of Disse, while in liver of patients with chronic hepatitis, they are also found in fibrotic areas [26].

Betaglycan, syndecans and glypicans are mem- brane-anchored heparin sulfate proteoglycans. Be- taglycan is the type III TGF- β receptor, while syn- decans may function as co-receptors of cytokines [15, 49]. They are transmembrane proteins with an amino-terminal extracellular domain, a single transmembrane domain and a short cytoplasmic

Fig. 6.1. Molecules of the hepatic extracellular matrix. The he- patic ECM consists of collagens, non-collagenous glycoproteins, elastin, glycosaminoglycans and proteoglycans. ECM-bound molecules include fibrin, plasmin, urokinase plasminogen acti- vator (upa), plasminogen activator inhibitor (PAI)-1, tissue trans- glutaminase, lysyl oxidase, growth factors/cytokines, metallo- proteinases (MMPs), and tissue inhibitors of metalloproteinase (TIMP)-3. In addition, transmembrane proteoglycans, which may serve as cell surface receptors, are cleaved by proteases, becom- ing ECM-bound, including CTGF connective tissue growth fac-

tor, a/bFGF acidic/basic fibroblast growth factor, GM-CSF granu- locyte macrophage colony-stimulating factor, IFN interferon, IL interleukin, HGF hepatocyte growth factor, KGF keratinocyte growth factor, OSM oncostatin M, PDGF platelet-derived growth factor, SPARC secreted protein acidic and rich in cysteine (syn- onymous with osteonectin or BM-40), TGF transforming growth factor, TSP thrombospondin, VEGF vascular endothelial growth factor, BM basement membrane, S sulphate. (Reproduced from reference 50, with permission from the author and publisher)

(3)

tail. Glypicans are integral membrane proteogly- cans that are anchored to the membrane via glycosyl phosphatidylinositol. Overexpression of glypican-3 is seen in hepatocellular carcinoma, which lends evidence to the suggestion that glypican-3 is mainly involved in growth-regulatory functions [63].

6.2.3 Laminin

Laminin is a non-collagenous glycoprotein and, together with perlecan, nidogen and collagen IV, is

one of the main components of the basement mem- brane [9]. It is composed of three disulphide-linked chains ( α, β and γ) with a characteristic cross shape.

A number of homologs of these chains have been discovered – five α chains, three β chains and three γ chains. Not all possible combinations of the three chains are used and so far 12 distinct laminin iso- forms have been identified. Among them, at least four may be found in human liver [9, 36]. Laminin is important not only in its structural role in the BM but also in its range of effects on cellular activities, namely cell adhesion, cell migration and cell differ-

Fig. 6.2. Changes in ECM and stellate cells during hepatic fibro- sis. A In normal sinusoids, stellate cells, located between sinu- soidal endothelium and hepatocytes, are quiescent and contain vitamin A droplets. Their foot processes encircle the sinusoid.

Microvilli depicted on hepatocytes indicate normal differentiat- ed function. The sinusoid endothelial cell fenestrae facilitate the flow of solutes between sinusoidal lumen and the hepatocytes.

Fig. 6.3. Binding of growth factors by ECM. The ECM can localize and store growth factors (see also Fig. 6.1), sequestering them for release by con- trolled proteolysis. (Reproduced from reference 50, with permission from the author and publisher)

B In chronic liver injury, a fibrillar matrix accumulates in the sub- endothelial space, produced primarily by activated stellate cells.

This matrix leads to loss of hepatocyte microvilli and reduced sinusoidal porosity. (Reprinted from Friedman SL, Arthur MJP.

Reversing hepatic fibrosis. Science and Medicine 2002;8(4):194–

205, with permission)

(4)

entiation. It mediates the cell–matrix interaction via binding to the integrin receptors [9, 11].

6.2.4 Fibronectin

Fibronectin (FN) is a multifunctional glycoprotein and plays crucial roles in various cellular functions.

It is a major component of normal and fibrotic he- patic ECM. Fibronectin molecules found in ECM are insoluble (cellular fibronectin, cFN) while plasma fibronectin, pFN, is soluble [50].

Fibronectin has a domain structure, consist- ing of three internally homologous repeats, termed types I, II and III [39]. The repeats are assembled into different functional domains that bind to vari- ous ligands such as collagen, heparin, fibrin and integrin. FN mRNA is post-transcriptionally modi- fied by alternative splicing at three variable regions of type III homology, EIIIA, EIIIB and EIIICS. Two mRNA isoforms are generated by either inclusion or exclusion of EIIIA and EIIIB, respectively. The II- ICS region has three subdomains (CS1, CS5 and the portion between these) and five isoforms may arise by exon subdivisions within this region. Alternative spliced variants have different biological properties.

For example, only cFN has EIIIA and EIIIB regions.

In vivo, FN forms fibrils stabilized by intermolecu- lar disulfide bridges. The polymerization process is driven by cell surface receptors, especially by in- tegrin α5β1.

Hepatocytes are the major source of pFN while cFN is produced by hepatocytes, activated HSC and sinusoidal endothelial cells [32, 58]. In rats, none of the FN isoforms is present in quiescent HSC from normal liver. In response to injury, sinusoidal en- dothelial cells express EIIIA fibronectin. This is a critical early event as the EIIIA segment is biologi- cally active, resulting in activation of HSC. The ac- tivated HSC in turn synthesize an EIIIA-containing FN themselves [32].

6.2.5

Matricellular Proteins

Matricellular proteins are a group of matrix proteins that modulate cell–matrix interaction and cell func- tion but do not contribute directly to the formation of structural elements [8]. Members of this group of distinct molecules include SPARC (secreted protein acidic and rich in cysteine, also termed osteonec- tin), thrombospondins (TSP1 and TSP2), osteopon- tin, tenascin-C, tenascin-X and CCN (CYR61, CTGF [connective tissue growth factor], Nov) family of

proteins. Studies in various cell types demonstrat- ed that these molecules are capable of sequestering growth factors (e.g. SPARC and PDGF [platelet-de- rived growth factor]), binding ions (e.g., osteonectin and Ca

2+

), inhibiting (TSP1) or clearing proteases (TSP2) and activating cytokines (e.g. TSP2 and TGF- β1). They regulate cell adhesion, migration, chemo- taxis, proliferation and apoptosis. Furthermore, complex effects are also exerted on ECM synthesis and collagen assembly.

In liver, expression of SPARC is upregulated dur- ing fibrogenesis and hepatocarcinogenesis [34, 42].

HSC are the main source of SPARC as shown both in vivo and in vitro [7, 18]. Tenascin and osteopontin have both been found in normal and fibrotic livers and hepatocellular carcinoma (HCC) [33, 44, 56, 59].

Furthermore, osteopontin overexpression is associ- ated with larger tumours, high-grade and late-stage HCC, intrahepatic metastases and higher early re- currence rate [44, 60].

6.3

Changes in ECM

from Normal to Fibrotic Liver

Hepatic fibrosis is associated with a significant change in both the quantity and composition of the ECM. Total collagen content increases by three- to tenfold [48]. The peri-sinusoidal low-density ECM is gradually replaced by a high-density ECM with accu- mulation of fibrillar collagens (types I and III) and an electron-dense BM [24]. There is also an increase in glycoproteins, proteoglycans, glycosaminoglycans and a shift from BM-type proteoglycans (heparan sulfate) to interstitial-type (dermatan sulfates and chondroitin sulfates) [23]. Changes in ECM are as- sociated with disappearance of endothelial fenestra- tions, a process termed “sinusoid capillarization”.

In vitro studies showed that interstitial matrix in- duces loss of fenestrations whereas physiologically derived BM maintains them [40].

Complex interactions exist between the cel-

lular components of liver and the ECM [16]. With

the modification of the ECM microenvironment,

cellular functions and phenotypes are inevitably

affected. This is evidenced by the loss of microvilli

in hepatocytes in fibrotic liver and compromised

synthetic activity of hepatocytes when deprived of

BM matrix [6]. Meanwhile, the high-density matrix

activates HSC, which further perpetuates the proc-

ess of fibrogenesis [17]. On the other hand, quies-

cent HSC cultured on BM-like matrix derived from

Englebreth-Holm-Sarcoma remain non-prolifera-

tive and non-fibrogenic [17]. A recent study further

(5)

showed that BM-like matrix (Matrigel) can induce quiescence of activated HSC, suggesting that resto- ration of normal ECM in liver might downregulate fibrogenesis by restoring HSC to a quiescent state as well [19].

Specific components of the ECM have been iden- tified to regulate cellular activities. One example is the EIIIA segment of fibronectin as described above. Another molecule that has been intensely researched is microfibrillar collagen VI, the expres- sion of which is upregulated in liver fibrosis. Colla- gen VI stimulates DNA synthesis and inhibits apop- totic cell death in HSC in vitro [2, 50].

6.4

Pathways of Cell–Matrix Interaction

6.4.1

The Integrin Family

The ECM interacts with cells via matrix receptors on the cell membrane; the best-characterized of these being the integrin family of heterodimeric trans- membrane receptors [14, 55]. Integrin receptors are composed of α and β subunits, and at least 18 α and 8 β subunits are currently known. The various combi- nations result in over 20 functional integrin dimers with different specificities. Their globular head do- main binds to ligands, which include components of the ECM and cell adhesion molecules. Most integrin ligands contain an Arg-Gly-Asp (RGD) tripeptide sequence, which is necessary but insufficient for sig- naling. Intracellularly, integrins are connected via associated proteins to the actin cytoskeleton.

The integrins are important not only in their adhesive function but also in their roles in modu- lating signal transduction pathways downstream of other receptors. Modulation of signaling pathways takes place via a number of mechanisms. Cell ad- hesion may result in change in shape and tension of the cell and nucleus via the cytoskeleton, which in turn influences gene expression [35, 37]. Mean- while, integrins may affect signal transduction via parallel activation of pathways that synergize at the level of phosphorylation of proteins [12, 47]. Signals generated by other receptors are also enhanced due to clustering of ECM-bound integrins in the plane of membrane. The activated integrins recruit signal- ing molecules that form complexes called focal ad- hesion complexes [22]. Examples of such molecules are caveolin, paxillin and tyrosine kinases such as fyn and focal adhesion kinase (FAK). Further down- stream, the complexes are associated with other ki-

nases and adaptor molecules. Clustering of these proteins results in amplification of the signal trans- duction. Lastly, integrin may cluster and transacti- vate signaling pathways involving receptor tyrosine kinases. An example of such functional cooperation is that between the PDGF pathway and integrin sig- naling as demonstrated by clustering of ligand-ac- tivated PDGF- β receptors in areas corresponding to focal adhesion complexes [10]. A complex web of crosstalk thus exists between the integrin pathways and signaling mechanisms of the growth factor re- ceptors.

The normal adult human hepatocytes express low levels of three integrin dimers: a collagen and laminin receptor, α1β1; a fibronectin receptor, α5β1;

and a tenascin receptor, α9β1. On the other hand, re- ceptors α1β1, α2β1, α5β1 and α6β4 have been iden- tified in cultured HSC [11, 45]. In experimental liver fibrosis, there is upregulation of laminin-binding integrins α6β1 and α2β1, and fibronectin-binding α5β1 [11, 46, 62]. Functionally, integrin antagonism by soluble integrin recognition sequence pentapep- tide GRGDS in rat HSC disturbs actin stress fiber formation and tyrosine phosphorylation of FAK caused by adhesion to ECM. The pentapeptide also induces p53 expression and apoptosis in the same cell type [30, 31].

Recent studies have also elucidated the roles of integrin in hepatocarcinogenesis. In HepG2 cells, stable transfection of β1 integrin modulates respon- siveness to hepatocyte growth factor (HGF) with resultant increased proliferation [61]. On the other hand, osteopontin binding to α5β3 integrin aug- ments in vitro adhesion of HepG2 to hyaluronate and may contribute to the mechanism by which os- teopontin enhances metastatic behavior in hepato- cellular cancer cells [20].

6.4.2

Discoidin Domain Receptor-2

Another matrix receptor recently found to have a

potential role in liver fibrosis is discoidin domain

receptor-2 (DDR2) [43]. DDR2 is a tyrosine kinase

receptor that responds to ECM ligands but not to

soluble peptide factors. It is activated primarily by

collagen type I, and to a lesser extent by collagen

types II, III and V. DDR2 is induced during HSC ac-

tivation, and the phosphorylated receptor mediates

growth stimulation and matrixmetalloproteinase-2

(MMP-2; see 6.5 below) production in response to

type I collagen; thus DDR2 can stimulate degra-

dation of normal liver ECM via MMP-2, while it is

further upregulated by accumulating interstitial

collagen, thereby establishing a positive feedback

(6)

loop. The importance of DDR2 in liver fibrosis is substantiated by a recent study showing its presence at elevated levels in the small bile ducts of patients suffering from primary biliary cirrhosis [38].

6.4.3

Growth Factors in ECM

In addition to its structural role and direct inter- action with cells, ECM also regulates cell function indirectly via modulation of the availability and activity of growth factors including PDGF, TGF- β (see Part II, Chap. 11), CTGF, vascular endothelial growth factor, HGF and so on [54]. Proteoglycans such as decorin, biglycan, fibromodulin and gly- cosaminoglycans are the main ECM components that bind growth factors and cytokine. They can in- teract with growth factors either via their core pro- teins or via their glycosaminoglucan side chains.

For example, decorin and biglycan bind TGF- β by their protein cores but interact with HGF through the heparan sulfate [25, 41]. Other ECM components such as fibronectin and laminin bind tumor necro- sis factor- α (TNF-α), while collagen binds PDGF, HGF and interleukin-2 (IL-2) [1, 51–53]. Binding of survival factors by interstitial matrix may prevent apoptosis of hepatocytes in liver that have acquired DNA damage, thereby perpetuating the expansion of cells with mutations and genomic instability. This observation may explain why cancer is more likely in livers that are cirrhotic, particularly in patients with hepatitis C. As well as protecting such factors from proteolysis, the ECM controls their release through the actions of proteases and their inhibitors, result- ing in further modulation of their activities.

6.5

Matrixmetalloproteinase and its Inhibitors Since the ECM components are highly stabilized and cross-linked molecules, they can only be broken down by a specific family of enzymes, the matrix- metalloproteinases (MMP). The activity of MMP is balanced by a group of proteins, the tissue inhibitors of the MMP family (TIMP) [4]. The integral activity of the MMP and TIMP determines the rate of ECM degradation. The MMP family comprises 25 differ- ent calcium- and zinc-dependent enzymes divided into five broad categories: interstitial collagenases, gelatinases, stromelysins, membrane-type MMP (MT-MMP) and metalloelastase. This functional classification is somewhat arbitrary as there is over- lap in activities among categories.

HSC are the main source of MMP in the liver.

In early primary culture, HSC express MMP-3 (a stromelysin) and MMP-2 but not TIMP [27, 29, 57].

Although this matrix-degrading phenotype could be a result of the isolation process, it might also re- flect the early phase of “pathological” matrix degra- dation in vivo after injury, during which the normal subendothelial matrix and BM are damaged. With prolonged culture of HSC, MMP-1/MMP-13 are downregulated while expression of TIMP-1, TIMP-2, MMP-2 and MMP-14 is enhanced [4]. In vivo stud- ies in rodent and human specimens concurred with such findings by demonstrating increased TIMP expression associated with fibrosis [5, 27]. Accord- ing to one proposed model, increased MMP-2 and MMP-14 production by HSC causes degradation of the pericellular matrix. This results in an altered HSC–matrix interaction and further HSC activa- tion [4]. Meanwhile, as the expression of TIMP is increased, degradation of newly synthesized colla- gen is inhibited, with consequent accumulation of ECM.

Resolution of fibrosis, on the other hand, is as- sociated with degradation of fibrillar ECM and res- toration of normal hepatic architecture. In carbon tetrachloride-treated rats, liver fibrosis regresses after cessation of treatment. The process is associ- ated with marked reduction in TIMP activity and an approximately fivefold increase in hepatic colla- genase activity [28]. MMP-1/MMP-13 derived either from HSC or Kupffer cells, together with MMP-14 and MMP-2 contribute to the fibrolysis, leading to restoration of normal histology.

Development of HCC in a fibrotic liver is asso- ciated with increased expression of MMP-2, TIMPs and MMP-14. ECM components including collagen 1, collagen IV and laminin are also overexpressed.

Interestingly, non-encapsulated HCC, which is as- sociated with poorer prognosis, has the highest expression of these molecules. This may imply an enhanced aggressiveness in such tumors with im- pairment of capsule formation.

6.6 Conclusion

The hepatic ECM can no longer be seen as an inert

structural element of the liver. Dynamic changes in

its composition and quantity take place in response

to external stimuli. This plasticity and responsive-

ness serve important physiologic functions, as

typified by the wound-healing response of fibro-

genesis. HSC are the predominant cells responsible

for producing the components of ECM as well as

(7)

the enzymes that break down the ECM. Injurious stimuli directly or indirectly through hepatocytes or Kupffer cells activate HSC via various signaling pathways, resulting in the synthesis of these compo- nents. Yet interaction between ECM and their sur- rounding cells is bidirectional. The biological activ- ities contained in the ECM components, in addition to the growth factors sequestered, regulate cellular functions in a multitude of ways. Further research and better understanding of this complex web of interactions and its molecular basis may eventu- ally facilitate the development of new therapies for chronic liver diseases and carcinoma.

Selected Reading

Schuppan D et al. Matrix as a modulator of hepatic fibrogenesis.

Semin Liver Dis 2001;21:351–372. (A detailed and up-to-date review of different matrix components and their possible bi- ological functions in liver. It also contains a series of figures depicting different structures of matrix proteins.)

Bedossa P, Paradis V. Liver extracellular matrix in health and dis- ease. J Pathol 2003;200:504–515. (A succinct overview of composition of extracellular matrix in liver, and cellular and molecular basis of liver fibrogenesis.)

Benyon RC, Arthur MJ. Extracellular matrix degradation and the role of hepatic stellate cells. Semin Liver Dis 2001;21:373–

384. (An excellent review on the role and interaction of met- alloproteinases and their inhibitors in liver fibrosis providing important insight on possible mechanisms of fibrolysis.) Friedman SL. Mechanisms of Disease: mechanisms of hepatic

fibrosis and therapeutic implications. Nat Clin Pract Gastro- enterol Hepatol. 2004;1:98–105. (A comprehensive and up- to-date series on the mechanism, natural history, diagnostic and therapeutic approches of liver fibrosis.)

References

1. Alon R, Cahalon L, Hershkoviz R et al. TNF-alpha binds to the N-terminal domain of fibronectin and augments the beta 1- integrin-mediated adhesion of CD4+ T lymphocytes to the glycoprotein. J Immunol 1994;15:1304–1313.

2. Atkinson JC, Ruhl M, Becker J et al. Collagen VI regulates normal and transformed mesenchymal cell proliferation in vitro. Exp Cell Res 1996;228:283–291.

3. Bedossa P, Paradis V. Liver extracellular matrix in health and disease. J Pathol 2003;200:504–515.

4. Benyon RC, Arthur MJ. Extracellular matrix degradation and the role of hepatic stellate cells. Semin Liver Dis 2001;21:373–

384.

5. Benyon RC, Iredale JP, Goddard S et al. Expression of tissue inhibitor of metalloproteinases 1 and 2 is increased in fibrot- ic human liver. Gastroenterology 1996;110:821–831.

6. Bissell DM, Caron JM, Babiss LE, Friedman JM. Transcrip- tional regulation of the albumin gene in cultured rat hepa- tocytes. Role of basement-membrane matrix. Mol Biol Med 1990;7:187–197.

7. Blazejewski S, Le Bail B, Boussarie L et al. Osteonectin (SPARC) expression in human liver and in cultured human liver myofibroblasts. Am J Pathol 1997;151:651–657.

8. Bornstein P, Sage EH. Matricellular proteins: extracellular modulators of cell function. Curr Opin Cell Biol 2002;14:608–

616.

9. Bosman FT, Stamenkovic I. Functional structure and com- position of the extracellular matrix. J Pathol 2003;200:423–

428.

10. Carloni V, Pinzani M, Giusti S et al. Tyrosine phosphorylation of focal adhesion kinase by PDGF is dependent on ras in hu- man hepatic stellate cells. Hepatology 2000;31:131–140.

11. Carloni V, Romanelli RG, Pinzani M et al. Expression and function of integrin receptors for collagen and laminin in cultured human hepatic stellate cells. Gastroenterology 1996;110:1127–1136.

12. Chen Q, Lin TH, Der CJ, Juliano RL. Integrin-mediated acti- vation of MEK and mitogen-activated protein kinase is in- dependent of Ras [corrected]. J Biol Chem 1996;271:18122–

18127.

13. Czaja MJ, Geerts A, Xu J et al. Monocyte chemoattractant protein 1 (MCP-1) expression occurs in toxic rat liver injury and human liver disease. J Leukoc Biol 1994;55:120–126.

14. Danen EH, Sonnenberg A. Integrins in regulation of tissue development and function. J Pathol 2003;200:471–480.

15. David G. Integral membrane heparan sulfate proteoglycans.

FASEB J 1993;7:1023–1030.

16. Friedman SL. Mac the knife? Macrophages – the double- edged Sword of hepatic fibrosis. J Clin Invest. 2005;115 (in press).

17. Friedman SL, Roll FJ, Boyles J et al. Maintenance of differ- entiated phenotype of cultured rat hepatic lipocytes by basement membrane matrix. J Biol Chem 1989;264:10756–

10762.

18. Frizell E, Liu SL, Abraham A et al. Expression of SPARC in nor- mal and fibrotic livers. Hepatology 1995;21:847–854.

19. Gaca MD, Zhou X, Issa R et al. Basement membrane-like ma- trix inhibits proliferation and collagen synthesis by activat- ed rat hepatic stellate cells: evidence for matrix-dependent deactivation of stellate cells. Matrix Biol 2003;22:229–239.

20. Gao C, Guo H, Downey L et al. Osteopontin-dependent CD44v6 expression and cell adhesion in HepG2 cells. Car- cinogenesis 2003;24:1871–1878.

21. Geerts A. History, heterogeneity, developmental biology, and functions of quiescent hepatic stellate cells. Semin Liver Dis 2001;21:311–335.

22. Geiger B, Bershadsky A, Pankov R, Yamada KM. Transmem- brane crosstalk between the extracellular matrix–cytoskel- eton crosstalk. Nat Rev Mol Cell Biol 2001;2:793–805.

(8)

23. Gressner AM, Bachem MG. Cellular sources of noncollagen- ous matrix proteins: role of fat-storing cells in fibrogenesis.

Semin Liver Dis 1990;10:30–46.

24. Hahn E, Wick G, Pencev D, Timpl R. Distribution of basement membrane proteins in normal and fibrotic human liver: col- lagen type IV, laminin, and fibronectin. Gut 1980;21:63–71.

25. Hildebrand A, Romaris M, Rasmussen LM et al. Interaction of the small interstitial proteoglycans biglycan, decorin and fi- bromodulin with transforming growth factor beta. Biochem J 1994;302(Pt 2):527–534.

26. Hogemann B, Edel G, Schwarz K et al. Expression of biglycan, decorin and proteoglycan-100/CSF-1 in normal and fibrotic human liver. Pathol Res Pract 1997;193:747–751.

27. Iredale JP, Benyon RC, Arthur MJ et al. Tissue inhibitor of met- alloproteinase-1 messenger RNA expression is enhanced relative to interstitial collagenase messenger RNA in experi- mental liver injury and fibrosis. Hepatology 1996;24:176–

184.

28. Iredale JP, Benyon RC, Pickering J et al. Mechanisms of spon- taneous resolution of rat liver fibrosis. Hepatic stellate cell apoptosis and reduced hepatic expression of metalloprotei- nase inhibitors. J Clin Invest 1998;102:538–549.

29. Iredale JP, Goddard S, Murphy G et al. Tissue inhibitor of metalloproteinase-I and interstitial collagenase expression in autoimmune chronic active hepatitis and activated hu- man hepatic lipocytes. Clin Sci (Lond) 1995;89:75–81.

30. Iwamoto H, Sakai H, Nawata H. Inhibition of integrin sign- aling with Arg-Gly-Asp motifs in rat hepatic stellate cells. J Hepatol 1998;29:752–759.

31. Iwamoto H, Sakai H, Tada S et al. Induction of apoptosis in rat hepatic stellate cells by disruption of integrin-mediated cell adhesion. J Lab Clin Med 1999;134:83–89.

32. Jarnagin WR, Rockey DC, Koteliansky VE et al. Expression of variant fibronectins in wound healing: cellular source and biological activity of the EIIIA segment in rat hepatic fibro- genesis. J Cell Biol 1994;127:2037–2048.

33. Kawashima R, Mochida S, Matsui A et al. Expression of os- teopontin in Kupffer cells and hepatic macrophages and stellate cells in rat liver after carbon tetrachloride intoxi- cation: a possible factor for macrophage migration into hepatic necrotic areas. Biochem Biophys Res Commun 1999;256:527–531.

34. Le Bail B, Faouzi S, Boussarie L et al. Osteonectin/SPARC is overexpressed in human hepatocellular carcinoma. J Pathol 1999;189:46–52.

35. Lelievre SA, Weaver VM, Nickerson JA et al. Tissue phenotype depends on reciprocal interactions between the extracel- lular matrix and the structural organization of the nucleus.

Proc Natl Acad Sci USA 1998;95:14711–14716.

36. Lietard J, Loreal O, Theret N et al. Laminin isoforms in non- tumoral and tumoral human livers. Expression of alpha1, alpha2, beta1, beta2 and gamma1 chain mRNA and an alpha chain homologous to the alpha2 chain. J Hepatol 1998;28:691–699.

37. Maniotis AJ, Chen CS, Ingber DE. Demonstration of mechan- ical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure. Proc Natl Acad Sci USA 1997;94:849–854.

38. Mao TK, Kimura Y, Kenny TP et al. Elevated expression of ty- rosine kinase DDR2 in primary biliary cirrhosis. Autoimmu- nity 2002;35:521–529.

39. Matsui S, Takahashi T, Oyanagi Y et al. Expression, localiza- tion and alternative splicing pattern of fibronectin messen- ger RNA in fibrotic human liver and hepatocellular carcino- ma. J Hepatol 1997;27:843–853.

40. McGuire RF, Bissell DM, Boyles J, Roll FJ. Role of extracellular matrix in regulating fenestrations of sinusoidal endothelial cells isolated from normal rat liver. Hepatology 1992;15:989–

997.

41. Mizuno K, Inoue H, Hagiya M et al. Hairpin loop and second kringle domain are essential sites for heparin binding and biological activity of hepatocyte growth factor. J Biol Chem 1994;269:1131–1136.

42. Nakatani K, Seki S, Kawada N et al. Expression of SPARC by activated hepatic stellate cells and its correlation with the stages of fibrogenesis in human chronic hepatitis. Virchows Arch 2002;441:466–474.

43. Olaso E, Ikeda K, Eng FJ et al. DDR2 receptor promotes MMP- 2-mediated proliferation and invasion by hepatic stellate cells. J Clin Invest 2001;108:1369–1378.

44. Pan HW, Ou YH, Peng SY et al. Overexpression of osteopontin is associated with intrahepatic metastasis, early recurrence, and poorer prognosis of surgically resected hepatocellular carcinoma. Cancer 2003;98:119–127.

45. Pinzani M, Marra F, Carloni V. Signal transduction in hepatic stellate cells. Liver 1998;18:2–13.

46. Racine-Samson L, Rockey DC, Bissell DM. The role of alpha1beta1 integrin in wound contraction. A quantitative analysis of liver myofibroblasts in vivo and in primary cul- ture. J Biol Chem 1997;272:30911–30917.

47. Renshaw MW, Ren XD, Schwartz MA. Growth factor ac- tivation of MAP kinase requires cell adhesion. EMBO J 1997;16:5592–5599.

48. Rojkind M, Giambrone MA, Biempica L. Collagen types in normal and cirrhotic liver. Gastroenterology 1979;76:710–

719.

49. Salmivirta M, Heino J, Jalkanen M. Basic fibroblast growth factor-syndecan complex at cell surface or immobilized to matrix promotes cell growth. J Biol Chem 1992;267:17606–

17610.

50. Schuppan D, Ruehl M, Somasundaram R, Hahn EG. Matrix as a modulator of hepatic fibrogenesis. Semin Liver Dis 2001;21:351–372.

51. Schuppan D, Schmid M, Somasundaram R et al. Collagens in the liver extracellular matrix bind hepatocyte growth factor.

Gastroenterology 1998;114:139–152.

52. Somasundaram R, Ruehl M, Tiling N et al. Collagens serve as an extracellular store of bioactive interleukin 2. J Biol Chem 2000;275:38170–38175.

(9)

53. 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 1996;271:26884–26891.

54. Taipale J, Keski-Oja J. Growth factors in the extracellular ma- trix. FASEB J 1997;11:51–59.

55. van der Flier A, Sonnenberg A. Function and interactions of integrins. Cell Tissue Res 2001;305:285–298.

56. Van Eyken P, Geerts A, De Bleser P et al. Localization and cel- lular source of the extracellular matrix protein tenascin in normal and fibrotic rat liver. Hepatology 1992;15:909–916.

57. Vyas SK, Leyland H, Gentry J, Arthur MJ. Rat hepatic lipocytes synthesize and secrete transin (stromelysin) in early primary culture. Gastroenterology 1995;109:889–898.

58. Xu G, Niki T, Virtanen I et al. Gene expression and synthe- sis of fibronectin isoforms in rat hepatic stellate cells. Com- parison with liver parenchymal cells and skin fibroblasts. J Pathol 1997;183:90–98.

59. Yamada S, Ichida T, Matsuda Y et al. Tenascin expression in human chronic liver disease and in hepatocellular carcino- ma. Liver 1992;12:10–16.

60. Ye QH, Qin LX, Forgues M et al. Predicting hepatitis B virus- positive metastatic hepatocellular carcinomas using gene expression profiling and supervised machine learning. Nat Med 2003;9:416–423.

61. Zhang H, Ozaki I, Mizuta T et al. Mechanism of beta 1-integrin- mediated hepatoma cell growth involves p27 and S-phase kinase-associated protein 2. Hepatology 2003;38:305–313.

62. Zhou X, Zhang Y, Zhang J et al. Expression of fibronectin re- ceptor, integrin alpha 5 beta 1 of hepatic stellate cells in rat liver fibrosis. Chin Med J (Engl) 2000;113:272–276.

63. Zhu ZW, Friess H, Wang L et al. Enhanced glypican-3 expres- sion differentiates the majority of hepatocellular carcinomas from benign hepatic disorders. Gut 2001;48:558–564.

(10)

Specific Signaling Pathways II

Chapter 7

Signaling Pathways in Liver Diseases: IL-6/gp130/Stat3

77

Rebecca Taub

Chapter 8

Vascular Endothelial Growth Factor Signaling

91

David Semela, Jean-François Dufour Chapter 9

Insulin Pathways

105

Miran Kim, Jack R. Wands Chapter 10

Nature and Function of Hepatic Tumor Necrosis Factor-

α Signaling 115

Jörn M. Schattenberg, Mark J. Czaja

Chapter 11

The Fas/FasL Signaling Pathway

129

Maria Eugenia Guicciardi, Gregory J. Gores Chapter 12

TGF-

β and the Smad Pathway in Liver Fibrogenesis 139

Axel. M. Gressner, Steven Dooley, Ralf Weiskirchen Chapter 13

Interferon Signaling

151

Massimo Levrero Chapter 14

CD14 and Toll Receptor

165

Allan Tsung, David A. Geller Chapter 15

The Wnt/

β-Catenin Pathway 173

Satdarshan P.S. Monga, George K. Michalopoulos Chapter 16

Notch Signaling in Liver Disease

193

Sarbjit Nijjar, Alastair Strain Chapter 17

Extracellular ATP: Important Developments in Purinergic Signaling

201

David Gatof, J. Gregory Fitz

(11)

Chapter 18

Calcium Signaling

211

Lawrence D. Gaspers, Nicola Pierobon, Andrew P. Thomas Chapter 19

MAP Kinase Pathways

in the Control of Hepatocyte Growth, Metabolism and Survival

223

Paul Dent

Chapter 20

PI3K, PTEN and Akt

239

Thomas F. Franke, Daniel C. Berwick Chapter 21

TOR Signaling and Cell Growth Control

259

Lisa M. Ballou, Richard Z. Lin

Chapter 22

Peroxisome Proliferator Activated Receptors

267

Raphaël Genolet, Liliane Michalik, Walter Wahli Chapter 23

Transcriptional Response to cAMP in the Liver

281

Maria Agnese Della Fazia, Giuseppe Servillo, Paolo Sassone-Corsi Chapter 24

Heme Oxygenase System

291

Sei-ichiro Tsuchihashi, Jerzy W. Kupiec-Weglinski Chapter 25

Nitric Oxide

299

Jose M. Prince, Timothy R. Billiar Chapter 26

Hypoxia-Inducible Factor-1 Signaling System

311

Deborah Stroka, Daniel Candinas

Chapter 27

Cyclins and CDKs in Liver Diseases

325

Siu Tim Cheung, Ronnie T. Poon Chapter 28

Telomeres and Telomerase:

Distinctive Roles in Liver Regeneration, Cirrhosis and Carcinogenesis

333

Ande Satyanarayana, K. Lenhard Rudolph

Chapter 29 NF-

κB 341

Tom Lüdde, Christian Trautwein Chapter 30

Ceramide: Cell Regulation from a Lipid Perspective

353

Jeffrey A. Jones, Yusuf A. Hannun

(12)

Chapter 31

Apoptosis and Mitochondria

367

Jose C. Fernández-Checa, Carmen Garcia-Ruiz Chapter 32

Ubiquitin-Proteasome Pathway in the Pathogenesis of Liver Disease

377

Samuel W. French, Fawzia Bardag-Gorce

Chapter 33

Biological Clock in the Liver

391

Hotishi Okamura

Riferimenti

Documenti correlati

Ad un gruppo composto in prevalenza da scialpinisti (più qualche ciaspolatore) è stato proposto un questionario on-line che sottopone loro alcune

Photosynthetic performance, the expression of genes involved in light signaling, and in the biosynthesis of isoprenoids and phenylpropanoids were analyzed in green (TIG) and red

    Dati  produttivi  delle  piante,  qualitativi  dei  frutti  alla  raccolta  e  nutrizionali   Produzione,  g/pianta Peso   Consistenz Resistenza  alla medio   a

Dato il suo ruolo nella regolazione di sopravvivenza, proliferazione e migrazione delle cellule endoteliali e delle cellule della muscolatura liscia dei vasi,

Fedrizzi, Influence of oxygen ovoilobility during skin-contoct mocerotion on the formotion of precursors of 3-mercoptohexon-l-ol in Múller-Thurgqu ond Souvignon Blonc

In conclusion, our pilot study shows for the first time with NGS of the 16S ribosomal subunit that COS and P supple- mentation routinely performed during IVF treatment