JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts Patricia L. Roche,1 Krista L. Filomeno,1 Rushita A. Bagchi,1 and Michael P. Czubryt*1 ABSTRACT Long regarded as a mere accessory cell for the cardiomyocyte, the cardiac fibroblast is now recognized as a critical determinant of cardiac function in health and disease. A recent renaissance in fibroblast-centered research has fostered a better understanding than ever before of the biology of fibroblasts and their contractile counterparts, myofibroblasts. While advanced methodological approaches, including transgenics, lineage fate mapping, and improved cell marker identification have helped to facilitate this new work, the primary driver is arguably the contribution of myofibroblasts to cardiac pathophysiology including fibrosis and arrhythmogenesis. Fibrosis is a natural sequel to numerous common cardiac pathologies including myocardial infarction and hypertension, and typically exacerbates cardiovascular disease and progression to heart failure, yet no therapies currently exist to specifically target fibrosis. The regulatory processes and intracellular signaling pathways governing fibroblast and myofibroblast behavior thus represent important points of inquiry for the development of antifibrotic treatments. While steady progress is being made in uncovering the signaling pathways specific for cardiac fibroblast function (including proliferation, phenotype conversion, and matrix synthesis), much of what is currently known of fibroblast signaling mechanisms is derived from noncardiac fibroblast populations. Given the heterogeneity of fibroblasts across tissues, this dearth of information further underscores the need for progress in cardiac fibroblast biological research. © 2015 American Physiological Society. Compr Physiol 5:721-760, 2015.

Introduction Throughout the body, fibroblasts play a valuable role in the synthesis, maintenance and breakdown of the extracellular matrix (ECM; see Table 1 for a list of abbreviations). Fibroblasts furthermore act as a source of intercellular signals to surrounding stromal cells, resulting in alterations in cell and tissue behavior. However, the mechanical requirements of different tissues vary widely, as do the physical forces impinging on these tissues, thus it is not surprising that the characteristics of fibroblasts can vary depending on their location within the body. Cardiac fibroblasts represent a unique cellular niche— they are exposed to constant periodic swings in physical forces transmitted via the ECM on a beat-to-beat basis; they signal to and receive signals from surrounding cardiomyocytes, a highly unique cell type within the body; and they can be significant contributors to cardiac pathology when their synthesis of ECM increases in the pathogenesis of cardiac fibrosis. The following review describes the current state of knowledge of the major signaling pathways active within cardiac fibroblasts and their phenotypic cousins, myofibroblasts, with due consideration of external signals including growth factors and physical force which act upon these cells. Key biological processes undertaken by fibroblasts, including matrix attachment, force sensing and mechanotransduction, migration, proliferation, and phenotype conversion are discussed in the context of the key signaling mechanisms responsible where known. The focus is on cardiac fibroblasts, but where deficiencies in the current literature are present, an

Volume 5, April 2015

overview of intracellular signaling in fibroblasts from other tissue sources is provided.

Cardiac Fibroblasts and the Extracellular Matrix Extracellular matrix production Fibroblasts are the most numerous individual cell type within the heart, accounting for over 50% of the cell population (a number that may vary widely between species), yet taking up a surprisingly low volume within the myocardium owing to their small physical size in comparison to cardiomyocytes (139, 399). During development, fibroblasts arise within the proepicardial organ from a precursor cell population dependent upon expression of the basic helix-loop-helix transcription factor TCF21: deletion of TCF21 abrogates cardiac fibroblast specification, resulting in late gestational lethality (2, 291). Fibroblasts are recognized as having a high degree of heterogeneity between organs and tissue types, with the expression of unique protein/mRNA signatures. This * Correspondence

to [email protected] Boniface Hospital Research Centre, University of Manitoba, Winnipeg, Manitoba, Canada 1 St.

P. L. Roche and K. L. Filomeno contributed equally to this work. Published online, April 2015 (comprehensivephysiology.com) DOI: 10.1002/cphy.c140044 Copyright © American Physiological Society.

721

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Table 1 ACE

Abbreviations

Comprehensive Physiology

Table 1

(Continued)

Angiotensin converting enzyme

Meox2

Mesenchyme homeobox 2

5’ AMP-activated protein kinase

MI

Myocardial infarction

AngII

Angiotensin II

MLC

Myosin light chain

ANP

Atrial natriuretic peptide

MLCK

MLC kinase

AP-1

Activator protein-1

MMP

Matrix metalloproteinase

αSMA

α-Smooth muscle actin

MRL

Mig-10/RIAM/lamellipodin

bFGF

Basic fibroblast growth factor/FGF2

MRTF-A

Myocardin-related transcription factor-A Membrane-type MMP Nuclear factor of activated T cells

AMPK

Brain natriuretic peptide

MT-MMP

Cas

Crk-associated substrate

NFAT

Cdc42

Cell division control protein 42 homolog

NF-κB

Nuclear factor κB

CDK

Cyclin-dependent kinase

Nox4

NADPH oxidase 4

CTGF

Connective tissue growth factor

p70S6K

p70S6 kinase p21-activated serine-threonine kinase

BNP

Diacylglycerol

PAK

DDR1

Discoidin domain receptor 1

PDGF

Platelet-derived growth factor

Dok1

Docking protein 1

PI3K

Phosphatidylinositol 3-kinase

ECM

Extracellular matrix

PIP2

Phosphatidylinositol(4,5)-bisphosphate

Epidermal growth factor receptor

PIP3

Phosphatidylinositol-3,4,5-trisphosphate

Extracellular signal-regulated kinase

PIX

PAK-interacting exchange factor

ET1

Endothelin 1

PKA

Protein kinase A

FA

Focal adhesion

PKC

Protein kinase C

Focal adhesion kinase

PKG

Protein kinase G

Protein 4.1, ezrin, radixin, and moesin homology

PLC

Phospholipase C

Filamin A

PTB

Phosphotyrosine binding

FOXO3a

Forkhead box O 3a

PYK2

Proline-rich tyrosine kinase 2

FSP1

Fibroblast-specific protein-1

Rb

Retinoblastoma protein

GEF

Guanine exchange factor

RIAM

Rap1-GTP-interacting adaptor molecule protein

G-protein-coupled receptor kinase-interacting protein 1

ROCK

Rho-associated kinase

Grb2

Growth factor receptor-bound protein 2

ROS

Reactive oxygen species

HA

Hyaluronic acid

RTK

Receptor tyrosine kinase

Has

Hyaluronan synthase

SF

Stress fiber

Hepatocyte growth factor

SFK

Src-family kinase

Hypoxia-inducible factor-1α

SLC

Small latent complex

Heparan sulfate proteoglycan

Sp1

Specificity protein 1

ICAP-1

Integrin cytoplasmic domain-associated protein 1

SRE

SRF response element

IGF-1

Insulin-like growth factor-1

SRF

Serum response factor

Interleukin-1

SYF

Src-Yes-Fyn

ILK

Integrin-linked kinase

Syn4

Syndecan-4

IP3

Inositol 1,4,5-trisphosphate

TAK1

TGFβ-activated kinase 1

JNK

c-Jun N-terminal kinase

TGFβ

Transforming growth factor β

LAP

Latency-associated protein

TGFβR1

TGFβ receptor 1

Leukemia-associated Rho guanine nucleotide exchange factor

TIMP

Tissue inhibitor of metalloproteinases

TNFα

Tumor necrosis factor α

LIM

Lin11, Isl-1, and Mec-3

TRPC

Transient receptor potential channel

LLC

Large latent complex

VASP

Vasodilator-stimulated protein

LTBP

Latent TGF-binding protein

VEGF

Vascular endothelial growth factor

MAPK

Mitogen-activated protein kinase

WASP

Wiskott-Aldrich Syndrome protein

mDia1

Mammalian diaphanous-1

WAVE

WASP-family verprolin-homologous protein

MEF

Mouse embryonic fibroblast

DAG

EGFR ERK

FAK FERM FLNa

GIT1

HGF HIF1α HSPG

IL-1

LARG

722

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

Intracellular Signaling of Cardiac Fibroblasts

heterogeneity extends to the heart as well (82). Atrial fibroblasts have been shown to express higher levels of genes implicated in matrix formation, proliferation, cellular structure, and metabolism; however, the biological significance of these variations is not fully understood (64). For example, plateletderived growth factor (PDGF) and its receptor are more highly expressed in atrial fibroblasts, and it has been speculated that these chamber-specific differences may be due to unique hemodynamic, neurohumoral, and structural characteristics between these two regions of the heart, resulting in increased fibrotic responses in the atrium (64). In the healthy adult myocardium, cardiac fibroblasts function primarily to maintain homeostasis via synthesis and secretion of ECM constituents and remodeling enzymes such as matrix metalloproteinases (MMPs; Fig. 1) (391, 425). Fibroblasts are characterized by an interior volume extensively occupied by endoplasmic reticulum and Golgi apparatus—features indicative of the elaborate synthetic and

Mechanical tension PDGF ED-A fibronectin

secretory machinery required for the key role these cells play in producing and depositing ECM components. The cardiac ECM provides a dynamic physical scaffold for the heart and plays a crucial role in initiating biochemical and mechanical cues for cell growth, development and differentiation. It is comprised of a myriad of constituents including fibrillar collagens, elastins, fibronectin, laminin, thrombospondin, proteoglycans, periostin, hyaluronan, and MMPs (45, 110, 149, 211). Synthesis of these ECM components by cardiac fibroblasts is regulated by a complex network of signaling pathways downstream of various effector molecules such as transforming growth factor β (TGFβ), PDGF, basic fibroblast growth factor (bFGF), and others, in both healthy and diseased states of the heart (66, 148). Cardiac fibroblasts also synthesize and secrete many bioactive molecules including interleukins, tumor necrosis factor α (TNFα), TGFβ, angiotensin II (AngII), endothelin 1 (ET1), natriuretic peptides, and vascular endothelial growth factor (VEGF) (395).

Mechanical tension TGF-β angiotensin II CTGF

Fibroblast

Proto-myofibroblast

Myofibroblast

ECM homeostasis Proliferation Migration Nascent adhesions

Collagen synthesis α-SMA expression ED-A fibronectin expression α-SMA-negative stress fibers Contractility Focal adhesions

Collagen cross-linking α-SMA-positive stress fibers Increased contractility Mature focal adhesions

Figure 1

Cardiac fibroblast activation and phenotype conversion. In the healthy myocardium, cardiac fibroblasts proliferate relatively slowly and maintain extracellular matrix homeostasis by continual synthesis of matrix constituents (e.g., collagens) and remodeling enzymes (e.g., matrix metalloproteinases/MMPs) at low basal levels. In response to stressors such as myocardial infarction, related proinflammatory and profibrotic cytokines, and/or increased matrix stiffness, cardiac fibroblasts become activated and initiate the wound healing process through matrix degradation (via increased expression of MMPs), proliferation, and migration to the site of injury, where they facilitate cardiac remodeling. In vitro work with fibroblasts has revealed a transient intermediate phenotype—the proto-myofibroblast, which shares features of both fibroblasts and myofibroblasts. Proto-myofibroblasts synthesize increased amounts of fibrillar collagens type I and III, and begin to express α-smooth muscle actin (αSMA) and the ED-A splice variant of fibronectin. Additionally, adhesions with the matrix are strengthened in proto-myofibroblasts as are intracellular actin filaments, imparting contractility. As of yet, it remains unclear whether proto-myofibroblasts exist as discrete entities in vivo. Fully phenotype converted cardiac myofibroblasts are characterized by reduced proliferation and migration, and in comparison to their precursors, are hypersynthetic for fibrillar collagens and other matrix components. Cardiac myofibroblasts also express increased levels of ED-A fibronectin and αSMA, the latter of which is also incorporated into stress fibers that are stronger and thicker than their αSMA-negative counterparts in proto-myofibroblasts. In conjunction with these fibers, the maturation and strengthening of focal adhesions imparts greater contractile ability upon cardiac myofibroblasts. Contraction of surrounding scar tissue and collagen crosslinking performed by myofibroblasts is necessary for scar maturation, though their persistence in the healed myocardium has also been implicated in the development of cardiac fibrosis.

Volume 5, April 2015

723

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Comprehensive Physiology

Fibrillar collagen production Fibrillar collagens are the major component of the cardiac ECM architecture and play a central role in maintaining the structural integrity of the heart (27). The collagen fibrils provide a scaffold for cell attachment, as well as sequestration of signaling molecules such as TGFβ, thereby contributing to biomechanical homeostasis of the myocardium. Of the more than 25 documented types of collagen, two are predominant in the heart: type I fibrillar collagen, which constitutes approximately 80% of myocardial collagen, and type III fibrillar collagen, which comprises about 10% (27, 124, 395). The remainder are a mixture of various fibrillar and nonfibrillar collagens such as type IV, V, and VI (37, 321, 449). Fibrillar collagens are produced and secreted into the interstitial space as procollagen molecules, which are then processed into smaller, mature collagen molecules by enzyme-mediated cleavage of the terminal propeptide domains. Mature collagen molecules can then be assembled and cross-linked to form structural collagen fibers. A disintegrin and metalloproteinase with thrombospondin motifs 2 (ADAMTS2) cleaves the amino (N)-terminal propeptide, while bone morphogenetic protein 1 cleaves the carboxy (C)terminal propeptide (106, 247). Upon cleavage by N- and C-proteinases, these collagen propeptides are released into the circulation (547). Serum levels of the 100 kDa C-terminal propeptide are positively correlated with diastolic dysfunction, and this is the most commonly used biomarker for quantitative determination of type I collagen synthesis (318). Increased serum levels of the 42 kDa N-terminal propeptide of type III collagen have been positively correlated with heart failure and mortality (103). The N-terminal propeptide is a prominent marker of type III collagen biosynthesis (428). Although measurements of these propeptides are indicative of overall fibrillar collagen synthesis, the organization and thickness of collagen type I and III fibrils can vary between not only healthy and diseased myocardium, but also within compartments of the heart itself such as the atria and ventricles (36).

Canonical TGF𝛽 signaling Collagen synthesis in the heart is regulated primarily through the TGFβ signaling cascade, initiated by activation of the heterotetrameric TGFβ receptor composed of type I and II subunit dimers (92,201). There is substantial evidence demonstrating that the canonical Smad effector protein signaling pathway, downstream of TGFβ, is responsible for increased collagen deposition in the cardiac ECM following injury (57). The binding of TGFβ to its receptor facilitates the phosphorylation of R-Smad2 and/or R-Smad3 by the receptor’s intrinsic serine-threonine kinase activity. Phosphorylated R-Smads can then associate with Co-Smad4, and the resulting RSmad2/3-Smad4 complex translocates to the nucleus to act as a transcription factor and drive the expression of target genes such as type I collagens (Fig. 2) (350,563). Smad3 is required for the induction of matrix gene expression. Expression of

724

TGF-β

TGF-β

Mechanical strain

LAP

LTBP-1

TGF-β receptor

Plasma membrane

Smad2/3

II I Smad2/3 P

Smad3 P

Smad4 Smad7

Scx

SMURF

Smad7

Smad2 P

sm Cytopla

SMURF

Nucleus

Scx

Smad3

Smad7

Smad4

P

Smad2 P

Ski Myofibroblast genes

COL1A2 Smad4

Scx

Smad3 P

Smad2 P

Figure 2

Canonical transforming growth factor-β (TGFβ) signaling and cardiac fibroblast gene expression. TGFβ is secreted by cardiac fibroblasts in an inactive form, bound noncovalently to latencyassociated peptide (LAP). LAP binds covalently to latent-transforming growth factor-β-bound protein-1 (LTBP-1) within the normal extracellular matrix, and these three components form the large latent complex, or LLC. Increased tension within the matrix causes a conformational change in the LLC that allows rapid release and activation of TGFβ. Such forces can be transmitted to the LLC either indirectly (via increased fibronectin fibril extension within the matrix), directly (through LLC-bound integrins on the surface of contractile myofibroblasts), or both. Active TGFβ can then bind to its receptor and induce heterotetramerization of TGFβ receptor subunits I and II. The serine/threonine kinase activity of TGFβ receptor subunit II is activated by ligand binding, in turn phosphorylating subunit I, which then phosphorylates receptor Smads 2 and 3 to allow recruitment of co-Smads such as Smad4. The Smad complex then translocates into the nucleus to activate transcription of target genes such as αSMA and fibrillar collagens. The transcription factor Ski attenuates this process in cardiac myofibroblasts, abrogating phenotype conversion. TGFβ-activated Smad3 increases expression of the transcription factor scleraxis (Scx), which acts synergistically with Smad3 to transactivate the collagen 1α2 gene. TGFβ receptor activation also activates a negative feedback loop via the expression of inhibitory Smad7, which suppresses Scx expression and forms a complex with Smurf in the nucleus that translocates to the cytoplasm to repress phosphorylation and activation of receptor Smads.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

fibrosis-related genes such as type I collagen and connective tissue growth factor (CTGF/CCN2) is interrupted in fibroblasts lacking the Smad3 gene (88, 137, 527). The inhibitory Smads, I-Smad6 and I-Smad7, prevent R-Smad activation through competitive binding for Smad2 and Smad3 to TGFβ receptor 1 (TGFβR1/ALK5) and increased receptor degradation (349, 362). The role of canonical signaling downstream of TGFβ has been extensively demonstrated in cardiac fibroblasts.

Noncanonical TGF𝛽 signaling Besides Smad-mediated regulation of collagen, TGFβ binding to its receptor also activates other noncanonical signaling cascades to regulate collagen synthesis, including extracellular signal-regulated kinase (ERK), c-Jun N-terminal kinase (JNK), TGFβ-activated kinase 1 (TAK1), and p38 mitogenactivated protein kinase (p38 MAPK) pathways (57, 278). Experimental data indicates that Ras, ERK, and JNK are rapidly activated within minutes of TGFβ treatment in several cell types, supporting the existence of additional Smadindependent TGFβ signaling pathways (169, 170, 343). The Leask laboratory has demonstrated that JNK activation in response to TGFβ in normal fibroblasts depends on focal adhesion kinase (FAK) (295). Recently they also reported that the absence of TAK1 impairs TGFβ-induced JNK phosphorylation: in contrast to wild-type fibroblasts, TAK1-null fibroblasts failed to exhibit JNK activation in response to TGFβ, and failed to upregulate the expression of the myofibroblast marker α-smooth muscle actin (αSMA) and CTGF (458). Activation of the Ras/MEK/ERK noncanonical TGFβ signaling cascade involves the heparan sulfate proteoglycan (HSPG) syndecan-4 (Syn4). In Syn4-null fibroblasts, TGFβ was unable to phosphorylate ERK or induce cell contraction (91). Syn4 can act directly as a coreceptor for TGFβ or indirectly by modulating the expression of genes required for TGFβ-induced activation of ERK. In dermal and cardiac fibroblasts, ERK activation is required for the expression of type I collagen and CTGF (282, 296). In turn, CTGF is an effector of TGFβ-induced myofibroblast phenotype conversion and ECM synthesis. Smad-independent responses to TGFβ appear to be mediated, at least partially, by activation of p38 MAPK. A TGFβ receptor subunit I mutant defective for Smad activation has been shown to activate p38 MAPK signaling in response to TGFβ (576). Betaglycan, another member of the HSPG family, has been shown to bind TGFβ by its core protein compartment (155). It controls access of TGFβ to TGFβ receptor subunit II, thereby affecting intracellular TGFβ activity (299). This HSPG is capable of activating the p38 MAPK pathway independently of the R-Smads and TGFβ ligand, consequently affecting target gene expression (424). These multiple and overlapping collagen synthesis pathways explain the potent profibrotic effect of TGFβ. However, the mechanisms of TGFβ-mediated activation of noncanonical pathways involving ERK, JNK, or MAPK, and the functional consequences of

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

activation of these pathways, are incompletely characterized in cardiac fibroblasts, in contrast to other cells such as dermal fibroblasts.

Renin-angiotensin-aldosterone system signaling Regulation of collagen synthesis by cardiac fibroblasts has also been associated with the renin-angiotensin-aldosterone system. Adult rat cardiac fibroblasts treated with AngII or aldosterone demonstrated a significant increase in collagen synthesis, which is abolished by pharmacological blockade of angiotensin receptors (49). AngII acts via two receptors (AT1 and AT2) that belong to the G protein-coupled receptor superfamily (135). The majority of the known physiological effects of AngII occur following binding to the AT1 receptor, inducing well-defined G protein-linked signaling pathways including phospholipase C (PLC) activation, which in turn causes the release of Ca2+ and subsequent activation of calmodulin kinase and protein kinase C (PKC) (309). AngII induces upregulation of TGFβ1 expression in cardiac fibroblasts through AT1, as well as increased collagen expression via TGFβ and ERK signaling (74, 175, 570). It was recently reported that AngII-induced ERK activation in cardiac fibroblasts occurs via a pathway which includes the Gβγ subunit of Gi and tyrosine kinases (e.g., Src, Ras, and Raf) (593). AngII also causes Smad2 phosphorylation, nuclear translocation of phospho-Smad2/4 and increased binding of the Smad complex to DNA, an effect blocked by the AT1 blocker losartan (414). AngII-treated cardiac fibroblasts exhibit αSMA-positive stress fibers (SFs) and contractile function in the presence or absence of TGFβ receptor subunit 1, suggesting that AngII can also stimulate myofibroblast phenotype conversion independently of TGFβ (122). Downregulation of the transcription factor serum response factor (SRF) in cardiac fibroblasts via RNA interference attenuates AngII-mediated myofibroblast phenotype conversion (122). In turn, TGFβ1 stimulates AT1 receptor expression directly via Smads 2/3 and TGFβ receptor subunit I/ALK5, providing evidence of autocrine regulatory cross-talk between the TGFβ and AngII signaling pathways (316). The stimulatory effect of TGFβ on AT1 receptor expression involves p38 MAPK, JNK, and phosphatidylinositol 3-kinase (PI3K) signaling (316). The renin-angiotensin-aldosterone system is arguably the second most characterized mechanism of ECM regulation by cardiac fibroblasts after the TGFβ signaling pathway.

Non-collagen matrix components The ECM protein fibronectin plays key roles in the developing heart and in the injured myocardium. Neonatal hearts synthesize more fibronectin than adult hearts (42, 514). More specifically, the ED-A splice variant of fibronectin is implicated in phenotype conversion of fibroblasts, as well as the inflammatory response to cardiac injury. The insertion of the ED-A module into fibronectin causes a conformational

725

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Comprehensive Physiology

change which increases its cell adhesive properties (310). ED-A fibronectin regulates cell adhesion and proliferation by binding integrin α4/CD49d (290, 448). The accumulation of other ECM components, including type I collagen, depends upon polymerization of fibronectin within the ECM (474). In a model of myocardial infarction (MI) in mice lacking ED-A fibronectin, there was reduced synthesis and deposition of collagen in cardiac tissue during the remodeling process (16). ED-A fibronectin polymerization is required for TGFβ1 -induced fibroblast activation and phenotype conversion to myofibroblasts, preceding αSMA expression following TGFβ1 treatment of cultured cells, as well as during granulation tissue development in vivo (448). ED-A fibronectin induces proinflammatory gene expression by nuclear factorκB (NF-κB) activation and acts as an endogenous ligand for TLR2 and TLR4 (181,303). Fibronectin has also been identified as an in vivo substrate of MMP9 using proteomic analysis as demonstrated by Zamilpa et al. (581). ED-A fibronectin has been investigated primarily in fibroblasts of noncardiac origin, but has recently been recognized as an important marker of phenotype switching in cardiac myofibroblasts (425). Hyaluronan (hyaluronic acid, HA) is a critical component of the cardiac ECM due to its contribution to several functions such as providing a hydrated environment for cell migration and proliferation, interaction with ECM proteoglycans such as versican and aggrecan, and its association with surface receptors like CD44 to regulate cell behavior (510). Early studies showed that a HA-rich environment was necessary for proper development of cardiac valve precursor cushions (33). Hyaluronan synthase (Has) enzymes appear to play a critical role in the cardiac ECM, since mice with genetic disruption of Has2 exhibit deleterious cardiovascular defects (73). However, investigation into the role of HA and its regulation by cardiac fibroblasts is presently very limited, though HA secretion has been closely associated with maintenance of the noncardiac myofibroblast phenotype (324,546). Blockade of HA synthesis inhibits the upregulation of αSMA expression that occurs during myofibroblast phenotype conversion induced by TGFβ (545). Interaction of HA with its receptor CD44 triggers intracellular signaling via Ras, Rac, and PI3K Table 2

MMPs and matrix remodeling The cardiac ECM is a dynamic structure requiring a precise balance between synthesis and degradation of its components, particularly the collagens. Interference with this equilibrium can lead to significant disruption of matrix organization in the myocardium. This is evident in the failing heart, when collagen synthesis outpaces degradation, leading to excessive deposition of collagen, myocardial stiffness, and fibrosis. Conversely, excessive degradation may weaken the ECM and contribute to chamber dilatation. ECM degradation may occur as a result of physical damage, but is also dynamically regulated by the activity of a family of proteolytic enzymes, the MMPs (Table 2). Secreted by various cardiac cell types including fibroblasts, myocytes, and endothelial cells, as well as inflammatory cells such as macrophages, the MMPs are zinc-dependent remodeling endopeptidases capable of collectively degrading the majority of the ECM components (45, 294, 401, 530). These enzymes are synthesized as relatively inactive zymogens or pro-MMPs, which are activated by cleavage of an amino-terminal propeptide, exposing the enzyme’s catalytic domain. More than 25 MMPs have been cloned and characterized to date. Of these, MMP1, MMP2, MMP3, MMP8, MMP9, MMP12, MMP13, MMP28, and membrane-type MMPs (MT1-MMP/MMP14) are involved in myocardial remodeling (246, 302, 312, 475). MMP1 (collagenase I) targets collagen types I, II, and III, as well as basement membrane proteins. MMP1 activity is induced by stimuli such as interleukin-1 (IL-1), brain natriuretic peptide (BNP), PDGF, and hypoxia-reoxygenation, but is reduced in response to AngII treatment (49, 85, 375, 516, 571). MMP2 and MMP9, also known as gelatinases, can process collagen types I, IV, and V, while MMP2 can also degrade collagen type III (482). Furthermore, MMP2 and MMP9 play an active role in the development of fibrosis—both MMPs can release ECM-bound latent TGFβ, thereby inducing collagen synthesis via the canonical Smad and noncanonical MAPK

Matrix Metalloproteinases Secreted by Cardiac Fibroblasts, and Their Target Molecules

Group Collagenase

MMP

Name

Substrates

1

Interstitial collagenase

Fibrillar collagens, aggrecan, proteoglycans, versican

8

Neutrophil collagenase

13 Gelatinase

Stromelysins

Membrane-type MMP

726

(161, 233, 363). It may be through these pathways that HA contributes to myofibroblast development and maintenance.

Collagenase 3

2

Gelatinase A

9

Gelatinase B

3

Stromelysin 1

7

Matrilysin

14

MT1-MMP

Denatured fibrillar collagens, collagen type IV, fibronectin, laminin, proteoglycans Collagens I, III, IV, fibronectin, proteoglycans

Fibrillar collagens, perlecan, versican

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

signaling pathways (577). Cardiac fibroblasts constitutively secrete MMP2 and its expression is significantly upregulated by TGFβ, TNFα, IL-1β, BNP, mechanical stress, and oxidative stress (32, 54, 189, 220, 308, 337, 468, 469, 480, 516). IL-1 and BNP are also capable of inducing MMP3 (stromelysin1) secretion by cardiac fibroblasts (468, 516). Under basal conditions, cardiac fibroblasts express very low amounts of MMP9. However, this increases dramatically when these cells are exposed to proinflammatory cytokines and oxidative stress (396, 468, 469). AngII appears to elicit variable, species-dependent responses in MMP9 secretion from cardiac fibroblasts (375, 477). MMP8 and MMP13 also degrade type I, II, and III collagen, whereas MMP12 specifically targets elastin. The membrane-type MT1-MMP, which is likely associated with the cell surface (rather than being secreted into the matrix), is capable of cleaving ECM proteins such as fibronectin, laminin, and type I collagen (26, 314, 454). Overall, the activity of most MMPs is generally upregulated by oxidative stress, mechanical stretch, BNP, and proinflammatory cytokines with varying effects caused by AngII. These modulators have also been shown to induce transcription of tissue inhibitors of metalloproteinases (TIMPs) in the myocardium, which oppose and control the proteolytic activity of MMPs. TIMPs are the principal inhibitors of MMPs in the heart, and are produced primarily by cardiac fibroblasts (338,522). Of the four TIMPs that have been cloned and characterized to date, TIMP2, TIMP3, and TIMP4 are expressed in the healthy myocardium (530). The expression of TIMP1 in the healthy heart is quite low but it is significantly upregulated in diseased cardiac tissue (204,234,288). TIMP1 expression is regulated by TNFα and IL-1β, and TIMP2 expression is similarly regulated by these cytokines as well as BNP (130, 516). In the healthy myocardium, MMPs and TIMPs are coexpressed and strictly controlled to maintain homeostasis of the cardiac matrix, while dysregulation of this balanced system resulting in net changes in proteolytic activity contributes to myocardial remodeling and development of heart failure (476,523). A number of clinical studies have revealed that the activity of MMPs is increased while TIMPs are decreased in the failing heart (287, 503, 521).

Cell-Matrix Adhesions Adhesions formed between cardiac fibroblasts and their surrounding ECM allow these cells to “mechanosense” their environment, that is, to detect and respond to changes in matrix structure and stiffness. Changes in matrix composition can thus result in alterations in gene expression, providing a mechanism to maintain tissue homeostasis. Growth of cardiac fibroblasts and other mesenchymal cells is highly reliant upon their adhesion to ECM or a similar substrate, as cells in suspension or on nonadhesive substrates fail to proliferate, even in the presence of serum or growth factors (403). During contraction and migration—critical processes in cardiac wound healing and tissue remodeling—fibroblasts

Volume 5, April 2015

Intracellular Signaling of Cardiac Fibroblasts

also exert forces on the surrounding matrix. Transmission of inward and outward forces by fibroblasts relies on the form and function of cell-matrix adhesions. These cell-matrix adhesions are composed of multiple proteins including adaptor proteins talin, α-actinin, vinculin, zyxin, and paxillin; scaffolding proteins such as p130Cas; and associated FAK and Src Family Kinases (SFKs), which associate directly with β-integrins (97,158,221). These adhesions are linked directly to F-actin filaments (in fibroblasts) and Stress Fibers (SFs) (in myofibroblasts), and mediate force transmission to the cytoskeleton and associated signaling pathways. Focal complexes and adhesions are highly variable in their size and composition, and signaling through such structures has been implicated in a number of cellular processes including cellcycle regulation and migration (432).

Formation of adhesions In general, the process of formation of cell-matrix adhesion is initiated with the formation of weak, nascent adhesions during attachment (called focal complexes or contacts), followed by cell spreading, and eventually the formation of stronger Focal Adhesions (FAs) and SFs (163). After formation of focal complexes upon fibroblast interaction with its substrate, contacts can mature into stronger, larger FAs. Both types of adhesions contain integrins as the major transmembrane protein, which interact with ECM components during attachment to increase cell surface contact with the matrix and facilitate cell spreading.

Roles of Rho and Rac Focal complexes are less than 1 μm in diameter, are highly dynamic, and utilize integrin αvβ3 as their major integrin, which recognizes matrix proteins by their RGD motifs (97). The formation of focal complexes is dependent upon activation of the RhoA pathway, which has a large number of downstream effectors involved in not only adhesion formation, but also maturation and association with SFs (in myofibroblasts), among other cellular processes (418). A number of studies have shown that RhoA/Rho-associated kinase (ROCK)-mediated, myosin II-generated contractility of the actin cytoskeleton is required for the development and maturation of focal complexes in fibroblasts (67, 412, 418). However, early adhesion involves a transient downregulation of RhoA, as well as phosphorylation and activation of the RhoA inhibitor p190RhoGAP through a c-Src-dependent, integrin-mediated mechanism (17, 18, 128). Rac itself may contribute to RhoA inhibition, as it has been shown to activate p190RhoGAP, both directly and indirectly via reactive oxygen species (ROS) production (65, 355).

Role of FAK The formation of a FAK-p120RasGAP-p190RhoGAP complex during focal complex formation in fibroblasts implies that FAK also plays a role in downregulating RhoA activity

727

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

during adhesion formation (507). Additionally, Rac1 activation and localization to focal contacts is higher in FAKexpressing fibroblasts than in FAK-null cells (81). FAK autophosphorylation at Y397 creates a Src-binding site that facilitates further FAK phosphorylation by Src and full FAK activation (78,435,439). FAK Y397 autophosphorylation also facilitates the recruitment of adaptor proteins Shc and p130Cas to focal complexes, permitting the formation of a Src, p130Cas and Dock180 complex in fibroblasts. The formation of this FAK-Src-Cas-Dock180 complex is associated with Rac and JNK elevation, suggesting its involvement in transient RhoA downregulation during adhesion formation (215). Both Src and FAK directly mediate p130Cas and paxillin tyrosine phosphorylation, leading to the recruitment of adaptor proteins Crk and Nck, and signaling complex assembly at FAs (415, 437, 439, 518). Additionally, in FAK-null mouse embryonic fibroblasts (MEFs), the related tyrosine kinase Pyk2 was found to regulate p190RhoGAP and thus RhoA activity, suggesting a compensatory role of these two similar tyrosine kinases in adhesion formation (293).

Role of mDia1 Though RhoA and Rac signaling counteract each other in focal complex formation, mammalian diaphanous-1 (mDia1), involved in cytoskeletal actin polymerization, acts upstream of both factors during adhesion formation. Rho-dependent Rac activation is mediated by mDia1-induced phosphorylation of p130Cas, and is antagonized by ROCK (downstream of RhoA) (515). However, even in the absence of RhoA activity, constitutively active mDia1 is capable of inducing the formation of focal complexes in NIH-3T3 fibroblasts (418). Thus, during focal complex formation, the balance in RhoA signaling appears to be tipped in the direction of mDia1, reinforcing Rac activity and thus its own inhibition. Clearly, the formation of adhesions between fibroblasts and their surrounding matrix is a highly coordinated process involving complex regulatory networks of the Rac and RhoA pathways.

Maturation of focal complexes to focal adhesions Gradual Rac inhibition is accompanied by RhoA activation in the maturation of focal complexes to stronger and larger FAs (several micrometers in diameter) characteristic of cardiac myofibroblasts (184). The establishment of mechanically stable contacts between integrins and their ECM ligands precedes the maturation of focal complexes into FAs, a process which involves coclustering of different integrins. One such clustering involves the vitronectin receptor integrin αvβ3, found in focal complexes, with the fibronectin receptor integrin α5β1. It is during maturation that adaptor proteins including talin, vinculin, paxillin and α-actinin are recruited, allowing attachment of FAs to the SFs that develop during phenotype conversion of cardiac fibroblasts to myofibroblasts (97).

728

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

The importance of RhoA activation in FA maturation lies within the establishment of cytoskeletal tension and is characterized by the recruitment of talin, paxillin, vinculin, αactinin, and tensin, accompanied by an increase in membranecytoskeleton interactions (128, 176). As in focal complex formation, actomyosin contractility also appears to be involved in FA maturation, as blebbistatin treatment of MEFs results in reduced FAK autophosphorylation (Y397) and impaired FA maturation (587). Two major effector pathways are implicated in the development of RhoA-mediated tension force within the fibroblast: ROCK and mDia1. Further upstream, guanine exchange factors (GEFs) enhance RhoA activation by facilitating the exchange of GDP for GTP (44). Downstream of Fyn, a tyrosine kinase implicated in SF development, RhoA activation is stimulated by the GEFs p115RhoGEF and p190RhoGEF, as well as leukemiaassociated Rho guanine nucleotide exchange factor (LARG) (141, 213, 565). Phosphorylation and activation of LARG can be achieved by FAK (95). However, in MEFs subjected to force, LARG is activated even with blockade of either FAK or MAPK/ERK, but is attenuated in Src-Yes-Fyn (SYF)deficient cells. Thus, LARG activates RhoA in this pathway in response to phosphorylation by Fyn, and is potentially amplified by FAK activity (185). The transmembrane proteoglycan Syn4 may also aid in FA maturation in cardiac fibroblasts. Syn4 binds different domains in matrix proteins via heparan sulfate side chains, and interacts with the actin cytoskeleton via paxillin and α-actinin (76, 129). Syn4 has been shown to localize with integrins in fibroblast FAs, and even in the absence of integrin binding is capable of recruiting adaptor proteins (30). Mechanical deformation of Syn4-binding sites induces activation of the MAPK/ERK pathway, and thus may contribute to FA maturation in concert with FAK (30). In the myocardium, force also appears to induce RhoA activation by a parallel FAK-dependent Ras-Raf-MAPK-ERK pathway during integrin-mediated adhesion maturation (588). In the absence of FAK, force-induced activation of Ras and ERK is completely abolished (185). In addition, force induces MEK activation and causes phosphorylation of GEF-H1, inducing RhoA activation (185). Inhibition of either FAK or MEK prevents activation of GEF-H1, which has been shown to be phosphorylated by ERK (171, 232). This additional pathway (FAK-Ras-Raf-MAPK-ERK-GEFH1-RhoA) does not appear to be dependent upon Src, as Src inhibition did not affect activation of FAK or Ras (185). Thus both pathways likely converge in force-induced RhoA-mediated adhesion maturation, at least in MEFs, and may play redundant/ compensatory roles in cardiac fibroblasts as well.

Regulation of cell-matrix adhesions Intracellular regulators The structure and function of cell-matrix adhesions is regulated by various factors, including extracellular forces (e.g., substrate rigidity and mechanical strain), integrin-ligand

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

binding and activation, and the cytoplasmic environment of cardiac fibroblasts. By far, the cytoplasmic component of cell-matrix adhesions is the most complex, comprised of a vast number of proteins. Two major groups of cytoplasmic proteins interact directly with β-integrin domains in FAs, namely, structural and regulatory proteins. Structural proteins such as talin, α-actinin, and Filamin-A (FLNa) participate in FA assembly and adhesion-mediated fibroblast migration, whereas regulatory proteins (FAK, vinculin, and paxillin) modulate adhesion formation and dynamics. FLNa and talin are actin-binding proteins that associate with the cytoplasmic tail of β1-integrins (250). Though some studies indicate that FLNa inhibits integrin-mediated adhesion by competing with talin for integrin tail binding, more recent evidence suggests that FLNa may be required for β1 integrin activation (via interaction with vimentin and PKCε) and adhesion to collagen substrate by human gingival fibroblasts and mouse NIH-3T3 fibroblasts (249, 251, 252). In cultured NIH-3T3 cells, FLNa functions to assist in the preservation of talin-integrin interactions in FAs, and also appears to play a mechano-protective role by preventing force-induced apoptosis through the Rac/Pak/p38 MAPK signaling pathway (392, 459). Though talin is not required for initial spreading of fibroblasts, it is critical for adhesion stabilization. Talin depletion of fibroblasts results in cell lifting and rounding (586). Recent work in MEFs has shown that α-actinin and talin may in fact compete for binding to β3 integrins in focal complexes, though they interact cooperatively in binding β1 integrins (413). α-Actinin provides a link between integrins and the actin cytoskeleton and is involved in cross-linking of actin filaments, in turn regulating FA maturation, as well as the development of intracellular tension (369). Increased tension in adhesion-associated α-actinin coincides with FA growth and maturation, downstream of RhoA activation. Inhibition of Rho-ROCK signaling results in decreased α-actinin tension and dissociation of FAs in Madin-Darby canine kidney cells (573). α-Actinin expression is necessary for FA maturation— depletion of α-actinin in MEFs decreases FA maturation and subsequent mechanotransduction (413). Knockdown of the type 4 α-actinin isoform ACTN4 in murine lung fibroblasts also impairs FA maturation and cell motility, though cell spreading is increased (450). Thus, it appears α-actinin may be directly involved in regulating FA maturation, though its involvement in initial adhesion formation may be negligible, and studies of its effect in cardiac fibroblasts are currently lacking. Vinculin is a cytoskeletal protein composed of three primary regions—the N-terminal head, a proline-rich and flexible neck region with hinge action, and a C-terminal tail (150). Vinculin in its inactive form exists in a folded conformation, and upon force-induced activation and recruitment to FAs, switches to an open conformation (24, 185). Conformational activation and recruitment of vinculin to FAs via interaction with various FA proteins is dependent upon sustained actomyosin-mediated forces (77). The head

Volume 5, April 2015

Intracellular Signaling of Cardiac Fibroblasts

region of vinculin regulates integrin dynamics and clustering in FAs, and participates in FA formation and maturation, independently of vinculin tail region interactions (146, 218). In contrast, the full-length (and activated) molecule is necessary for the development of vinculin-dependent tensile force within MEFs. The head region of vinculin interacts with talin and α-actinin, allowing attachment to membrane integrins (77). Interaction of vinculin with talin is critical to the formation and maturation (increased size and strength) of FAs in fibroblasts (210, 218). The vinculin neck region interacts with signaling proteins including vinexin, vasodilator-stimulated protein (VASP), ponsin, and the Arp2/3 complex (77). VASP localizes to FAs where it binds with Arp2/3, talin, and Rap1-GTP-interacting adaptor molecule protein (RIAM), a member of the Mig10/RIAM/lamellipodin (MRL) family of proteins (285, 446). Though the role of VASP and Arp2/3 in fibroblast adhesion dynamics has not yet been examined, VASP appears to modulate F-actin nucleation by Arp2/3 and enhance actin turnover rates (via integrin β1 activation), potentially contributing to adhesion-cytoskeleton interactions (446,560). RIAM is a critical downstream mediator of Rap1-dependent fibroblast adhesion, and via interaction by its N-terminus with talin, also promotes talin-induced integrin activation (285). Knockdown of RIAM reduces adhesion, integrin activation, and cellular F-actin (192, 267, 272). The tail region of vinculin interacts with phosphatidylinositol(4,5)-bisphosphate (PIP2 ), paxillin, and actin to link FAs to the actin cytoskeleton (77, 218, 591). Linking of F-actin with FAs causes an inward pull, inducing conformational changes in vinculin and its subsequent activation, which further increases adhesion growth and stabilization (77). Paxillin is a cytoskeletal protein localized to FAs which is heavily tyrosine-phosphorylated during fibroblast phenotype conversion and adhesion formation. Paxillin is not necessary for the formation of FAs, and appears to be recruited either downstream or independently of vinculin (218). In fibroblasts, paxillin recruitment to FAs occurs independently of interactions with vinculin and/or FAK and appears to rely on its Lin11, Isl-1 and Mec-3 (LIM) domains (52). In vitro, paxillin binds to vinculin as well as c-Src (via its SH3 domain) and Crk (via its SH2 domain). The N-terminal half of paxillin also contains a putative binding region for the FA tyrosine kinase pp125Fak, though this interaction has not been observed in cardiac fibroblasts (520).

Integrins The major transmembrane components in adhesions are integrins, which mediate fibroblast adherence to ECM components such as fibronectin, collagen, and vitronectin. Additionally, integrins may participate in the formation of cell-cell adhesions through binding of adhesion molecules like vascular cell adhesion molecule and intercellular adhesion molecule (386). Integrins are heterodimeric transmembrane receptors

729

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

essential to cell-matrix adhesions that function primarily in mechanotransduction (179, 222, 564). They possess a large extracellular domain to which various ECM components bind, a single transmembrane segment, and a short cytoplasmic tail that interacts with cytoplasmic adapter proteins that link integrins to the actin cytoskeleton (176, 219, 580). Integrin heterodimers are composed of one α and one β subunit, resulting in at least 24 possible combinations. Each integrin dimer is capable of binding various extracellular ligands, and each ligand is typically able to bind various integrins, resulting in tremendous diversity and potential for functional redundancy (83, 224, 377, 445, 466). Cardiac fibroblasts have been shown to express specific integrins including α5β1 (which binds fibronectin and osteopontin), αvβ1, αvβ3, and αvβ5 (which binds fibronectin, osteopontin, and vitronectin as well as latent TGFβ) (331). Additionally, expression of α1, α2 (which are collagen-specific), α5 (which specifically binds fibronectin), β1, and β2 integrin subunits has been demonstrated in primary cardiac fibroblasts (59, 315). Integrins function bidirectionally across the plasma membrane. In the absence of external stress, integrins exist in a “bent,” low affinity conformation, and require activation to take on an extended high-affinity conformation (572). Integrin activation can occur by the binding of ECM ligands (“outside-in” signaling), or by the binding of its cytoplasmic tail to intracellular activators (“inside-out” signaling) (453). Binding of the major cytoskeletal protein talin to βintegrin tails has been shown by numerous lines of evidence to be a common final step in activation of β1, β2, and β3 integrins—at least in vitro (15, 68, 71, 351, 465, 489). Interaction of talins with β-integrins occurs through binding of a high-affinity phosphotyrosine-binding (PTB) domain found in the FERM (protein 4.1, ezrin, radixin, and moesin homology) domain of the N-terminal head of talin to the first NPXY motif in the β-integrin cytoplasmic tail (12, 70, 387, 489). Other regions of the N-terminal head have been shown to enhance integrin activation by the PTB domain (46). Additional proteins containing PTB domains have been shown to bind integrins without activation (69). Activation by talin requires an additional interaction of the N-terminal head of talin with the membrane-proximal region of the integrin β3 cytoplasmic tail (69, 524, 551). This additional interaction mediates formation of a stable helix that spans the transmembrane region of β-integrin and forms a salt bridge with an aspartate residue on the β3-integrin cytoplasmic tail, disrupting the interaction of this residue with an arginine on the α-integrin component of the heterodimer (13,551). The alteration of β-integrin tilt angle at its transmembrane segment may explain why talin is critical for integrin activation (342). Studies in various nonfibroblast cells have identified other FERM domain proteins called kindlins as coactivators in talinmediated integrin activation (231, 336). In dermal fibroblasts, kindlin-2 was essential for activation of β1 integrin, as well as stabilization and maturation of FAs and SFs, but it is unclear whether a similar mechanism occurs in cardiac fibroblasts (199).

730

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

The question remains as to how talins (and perhaps kindlins) are recruited to integrins, but there is evidence that FAK may fulfill this role. In wild-type MEFs, talin, and paxillin colocalize with integrins in nascent adhesions (276). However, in FAK-null MEFs, talin was absent from paxillinpositive nascent adhesions, yet present in mature adhesions. Interestingly, β1 integrin was observed even in talin-null adhesions in these cells, and in mammary epithelial cells depleted of talin 1 (539). Additionally, talin colocalized with FAK in early adhesions in MEFs with a mutation (Y783A) in the talin-binding region of β1-integrin tails. Together, these results suggest that FAK may recruit talin to β-integrin tails in nascent adhesions, and is also capable of shuttling talin to adhesion sites in the absence of integrin binding. However, these studies also imply that talin is not required for β1-integrin activation, though this may not be the case with other β-integrin types, and involvement of talin in integrin activation may also be cell type specific. The mechanism by which FAK itself is recruited to integrins remains unknown, but may involve p190RhoGEF, which binds FAK (327). Another potential candidate in talin recruitment to integrins involves the Ras GTPase subfamily member Rap1, which has been shown to stimulate integrin activation in epithelial and endothelial cells (40, 43, 264). MRL proteins have been shown to act as scaffolds that link Rap1 to talin and facilitate the recruitment of talin to β3 integrin tails at the plasma membrane of CHO cells (285). Short N-terminal sequences in the MRL proteins RIAM and lamellipodin both bind directly to talin, and in the case of RIAM this interaction is required for activation of αIIbβ3 integrin (285). In T cells, RIAM was shown to act downstream of Rap1 in β1 and β3 integrin activation, and was necessary for Rap1-induced cell adhesion. RIAM also interacts with profilin and VASP, proteins involved in regulation of actin dynamics. Not surprisingly, RIAM knockdown not only displaced Rap1 from the plasma membrane of these cells, but also impaired adhesion and cytoskeletal actin polymerization (272). Regulation of talin binding to β-integrin tails and subsequent integrin activation is regulated by several potential mechanisms. PTB domain-containing proteins, which have also been shown to bind the NPXY motif of β-integrin tails, may compete with talin for binding and inhibit integrin activation through “inside-out” signaling. In vitro binding assays have identified potential competitors, including Numb (a Notch signaling inhibitor), EPS8 (a regulator of Rac signaling), tensin (a FA protein), integrin cytoplasmic domainassociated protein 1 (ICAP-1), and docking protein 1 (Dok1, a downstream target of integrin signaling), which bind to various classes of β-integrins (69). Though in most cases the binding affinity is low, and thus not competitive with talin, in some cases binding affinity can be increased by tyrosine phosphorylation of these competitors by enzymes like SFK (12, 370). For example, in the absence of SFK, Dok1 binding is low and does not cause a decrease in integrin activation (12, 370). However in the presence of SFK, Dok1-binding affinity for the NPXY domain is increased and competition occurs.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

Other mechanosensitive pathways have been linked to force-induced integrin activation, such as activation of FLNa by mechanical force application on beads bound to β1integrins. Force application by this method was found to induce association of p38 MAPK with the transcription factor specificity protein 1 (Sp1) and subsequent Sp1 bindingmediated FLNa activation (117). In addition to integrin activation, association of the talin N-terminal head with β-integrin cytoplasmic tails is also implicated in integrin clustering, an important step in mechanosensory outside-in signaling (202, 421). Clustering of integrins at the cell membrane of fibroblasts is linked to the localization of FAK at the cell periphery and its activation by autophosphorylation at Y397, leading to downstream activation of Rac1 and the formation of lamellipodia (378). In mouse and human gingival fibroblasts, as well as NIH-3T3 fibroblasts, discoidin domain receptor 1 (DDR1) regulates interaction of β1 integrins with fibrillar extracellular collagen (386, 479). Overexpression of DDR1 resulted in increased number, length, and size of FAs, as well as increased β1integrin activation (479). Clearly, the regulation of FA formation, maturation, and association with the actin cytoskeleton involves a complex temporal and spatial interplay of a variety of intracellular and transmembrane players. Additionally, extracellular forces sensed via mechanotransduction play an important role in cell-matrix adhesion dynamics, cytoskeletal organization, and gene expression.

Force Transduction in Cardiac Fibroblasts Mechanosensing of extracellular forces As in fibroblasts of the skin and tendons, cardiac fibroblasts respond to mechanical cues and physical forces deriving from the cardiac ECM. In a fibroblast firmly attached to its substrate, the external and internal forces are in equilibrium; however, a disturbance of this balance in either direction leads to changes in cell form and function. Transmission of forces between fibroblasts and the ECM occurs through adhesion contacts and attached cytoskeletal components. Active “pulling” on such components by the cell provides information about the mechanical state of the surrounding matrix, and in turn, mechanosensing alters the dynamics of cell adhesions and attached cytoskeleton components. In the normal myocardium, cardiac fibroblasts are exposed to a low degree of tissue stiffness, typically in the range of 4 to 18.5 kPa (34, 35, 180, 554). In the fibrotic heart, however, the increase in interstitial collagen results in increased tissue stiffness, in the range of 20 to 100 kPa (208). In response to increased stiffness, cardiac fibroblasts convert to the myofibroblast phenotype, increasing αSMA expression and forming actin SFs that terminate at FAs, which increase in size, number, and strength (208). This phenotype conversion event is capable of occurring in response to rigid substrates without any additional factors

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

present, thus a mechanism must exist to permit cells to identify physical changes in their extracellular environment and initiate the appropriate intracellular responses. A number of different mechanisms have been proposed in fibroblast force mechanotransduction. One candidate is the activation of mechanosensitive cell membrane channels such as transient receptor potential cation channels (TRPCs), which induce signaling by facilitating the movement of ions such as calcium into the cell, resulting in activation of calcineurin/NFAT signaling (381). Another candidate mechanism is the activation of a critical step in RhoA signaling in a force-dependent manner. FAK may play the role of the critical step in this mechanism, as its activation is induced by integrin-mediated force in skin fibrosis, activating inflammatory signals such as monocyte chemoattractant protein-1 (559). FAK is also induced by cyclic strain in MEFs (378). Tensile strain on matrix adhesions in MEFs results in autophosphorylation (on Y397) and activation of FAK, which is blocked by RhoA/ROCK inhibitors (511). Additionally, it has been hypothesized that mechanotransduction may occur via direct transmission of force (either external or cell-generated) to transcriptional regulators such as Yap and Taz. These factors are elements of the growthmodulating Hippo pathway, and have been found to be involved in stiffness-mediated transcriptional changes. In epithelial cells on rigid substrates, Yap/Taz is localized to the nucleus, whereas on softer substrates, Yap/Taz is found in the cytoplasm or reduced altogether (228). In this model, changes in the shape of the nucleus and alterations of chromatin organization result in transmission (rather than transduction) of mechanical signals through a mechanically-integrated cytoskeleton. The linker of nucleoskeleton and cytoskeleton complex (LINC) is capable of transmitting such forces to the nuclear matrix or the lamin network underlying the nuclear membrane, resulting in structural changes in the nucleus and chromatin that could alter chromatin folding, protein association, and thus regulate transcription by direct force transmission (228). However, the latter mechanism has not yet been examined in the context of the cardiac fibroblast.

Force-activated ion channels Tension applied to the plasma membrane of cells on stiff substrates may increase cation influx (e.g., calcium) through stretch-activated channels, disrupting intracellular equilibrium. For example, the enzyme proline-rich tyrosine kinase 2 (PYK2) is only weakly activated by integrin activation in response to fibronectin attachment (whereas this causes strong FAK activation), but is highly sensitive to intracellular calcium signals (292). PYK2 and FAK have similar domain structures (their domains include FERM, kinase, proline-rich, and FAT), possess common autophosphorylation sites, and act as scaffolding proteins in the transmission of signals from G protein-coupled receptors to downstream signaling pathways such as MAPK (506). Despite these similarities however, FAK and PYK2 possess distinct signaling roles due to differential

731

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

binding of targets to their FERM or FAT domains, respectively. In addition, FAK is localized to FAs whereas PYK2 displays a perinuclear distribution (258). Additionally, cardiac fibroblast TRPCs including TRPM7, TRPC3, TRPC6, and TRPV4 regulate Ca2+ influx necessary for myofibroblast phenotype conversion (122, 140, 194, 502, 579). Although all four TRP channels exhibit mechanosensitivity in other cell types, only TRPV4 has been specifically studied in the context of mechanosensing by cardiac fibroblasts (359,380). TRPV4 inhibition in cardiac fibroblasts attenuated the combined phenotype conversion effects of substrate stiffness and TGFβ treatment, as measured by incorporation of αSMA into SFs (3).

Comprehensive Physiology

The most well-defined responses to increased substrate stiffness include increased actomyosin contraction, activation of RhoA, increased tyrosine phosphorylation of various cytoplasmic proteins, and activation of FAK. The modality through which such signals are produced involves physical forces being exerted on cell-matrix adhesions consisting of transmembrane integrins, adaptor/scaffold proteins (talin, paxillin, vinculin, zyxin, and p130Cas), associated kinases (Src kinase and FAK) and the actin cytoskeleton (F-actin, α-actinin, and myosin II). Intracellular signals transmitted through these complexes result in alterations in cell morphology, migration, contractility, phenotype conversion, and other functions (315,383). Integrins are critical components of the mechanotransduction apparatus in cardiac fibroblasts, and their activation is involved in stimulating downstream kinases (such as FAK or integrin-linked kinase/ILK), small GTPases (Rho and Rac), and signaling pathways such as those involving Akt, Raf, PI3K, and the MAPK/ERK pathways (Fig. 3). Expression of ILK is necessary for force-induced stimulation of RhoA signaling (307).

complexes (40, 272, 431, 587). Conversely, ROCK activity downstream of RhoA antagonizes this pathway (515). Inhibition of ROCK by the compound Y27632 suppresses tyrosine phosphorylation of FAK and paxillin, but not p130Cas (515). Though p130Cas has been placed downstream of kinase activation, it appears to have mechanosensory capabilities. In response to stretch of either the cytoskeleton or individual Cas molecules, the CasSD becomes exposed, facilitating its phosphorylation, Cas activation, and downstream signaling events (431, 495). Exposure of the CasSD involves separation of the C-terminus, which binds SFK, and the N-terminus, which contains an SH3 domain with which proline-rich sequences of its critical binding partner FAK interact (162, 348, 422). Colocalization of FAK and Cas, as well as high levels of Cas phosphorylation at adhesion sites, is lost in fibroblasts with a Cas binding-deficient version of FAK (712-715 P>A) (587). However, since blocking FAK kinase activity and Y397 autophosphorylation does not affect CasSD Y165 phosphorylation, it appears as though FAK may act as a docking molecule for Cas in adhesions, but is not necessary for activation of its downstream signaling pathways. Activation of Cas appears to be adhesion dependent, as suspended fibroblasts treated with RGD peptide (a soluble integrin ligand) do not exhibit increased CasSD phosphorylation (587). During initial spreading of MEFs on fibronectin-coated substrate, p130Cas phosphorylation was found to be catalyzed by F-actin polymerization and enhanced by integrin engagement and anchorage to fibronectin. Independently of its kinase activity, FAK linked p130Cas to an N-WASP (ubiquitous WASP homolog)/actin polymerization complex. When p130Cas expression was silenced via shRNA, MEFs showed reduced cell spreading on fibronectin and cell-matrix adhesions failed to mature (587). In addition to cell-matrix adhesions, N-cadherins are able to act as force and stiffness sensors in fibroblasts (174, 259). N-cadherins associate with α-/β-catenin complexes that are capable of binding vinculin, α-actinin, and F-actin, and direct activation of cyclin D1 by the β-catenin component (260,499, 575).

P130Cas

Focal adhesion kinase

The Crk-associated substrate (Cas) family consists of multidomain scaffolding proteins activated by phosphorylation. Src and Ab1 family kinases target a motif consisting of 15 YxxP repeats in the central substrate domain (CasSD) of p130Cas (125). Phosphorylation at the CasSD by FAK produces docking sites for SH2 domain proteins such as Crk (162, 257, 422, 533). In turn, Crk recruits GEFs to Rac1 and Rap1, activators of lamellipodium protrusion and cell-matrix adhesion formation (62, 192, 431, 451, 495). Additionally, p130Cas mediates adhesion-induced integrin signaling in an SFK-dependent manner (360, 452, 587). Phosphorylation of p130Cas induces downstream activation of the small GTPase Rap1, which then activates RIAM, subsequently causing activation of talin and integrins, and the formation of focal

FAK is the principal downstream target of integrin activation in cardiac fibroblast mechanosensing (315). Integrindependent FAK phosphorylation at Y397 promotes SH2 domain binding of SFK and other SH2 domain-containing proteins, such as PLC and growth factor receptor-bound protein 2 (Grb2) (436, 585). Downstream of FAK activation and SFK binding, adhesion dynamics are affected by activation of the Ras and ERK/MAPK pathways, as well as myosin light chain kinase (MLCK) (331). Additionally, force-induced transcriptional activation of αSMA expression [via myocardin-related transcription factor-A (MRTF-A)] in NIH-3T3 fibroblasts required FAK, gelsolin (a key regulator of actin filament assembly), and type-I phosphatidylinositol 4-phosphate 5 kinase-γ (involved in uncapping gelsolin from

Effector Signaling Pathways in Mechanosensing

732

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

Intracellular Signaling of Cardiac Fibroblasts

ECM ligands (FN, OPN, VTN)

Activated integrin heterodimer

Inactive integrin heterodimer Plasma membrane

α β

Talin

αβ

FAK Rap1

Talin

WASP

FAK

SFK

RIAM Dok1 and others

p130Cas

SFK RhoA F-act in

Ras

mDia1

ROCK

Raf

MAPK

MRTF-A

MRTF-A

F-actin formation and contraction

ERK

sm Cytopla

SRF

Nucleus

α SMA MRTF-A

SRF

c-fos, egr-1

Sap

Figure 3

Mechanosensing through integrins. Integrin activation and downstream signaling occurs in response to both outside-in and inside-out signaling, both of which evoke conformational changes in integrin structure. The inside-out pathway involves binding of talin to the β-integrin NXPY motif. Various mechanisms have been proposed in the regulation of talin binding-mediated integrin activation through competing factors such as the Src-family kinase (SFK)-activated docking protein 1 (Dok1), among others. Outside-in signaling requires interaction of extracellular matrix (ECM) ligands with integrins, a mechanism that is sensitive to the degree of extracellular tension or force. Ligands expressed by cardiac fibroblasts and found in the ECM include fibronectin (FN), osteopontin (OPN), and vitronectin (VTN). Similar to inside-out signaling, recruitment of talin to integrins is a key step. Putative candidates in talin recruitment in cardiac fibroblasts include focal adhesion kinase (FAK) and Rap1-GTP-interacting adaptor molecule (RIAM) downstream of Rap1 activation. Force-activated integrins mediate phosphorylation of SFK and FAK, the latter of which forms a complex linking the mechanosensory protein p130Cas to Wiskott-Aldrich Syndrome protein (WASP) and cytoskeletal F-actin filaments. Force-induced FAK activation upregulates the small GTPase RhoA, activating downstream mammalian diaphanous-1 (mDia1) and Rho-associated kinase (ROCK), inducing subsequent actin filament polymerization and myosin II-mediated contraction. ROCK activation also induces the release of myocardinrelated transcription factor-A (MRTF-A), which forms a complex with serum response factor (SRF) and directly activates α-smooth muscle actin (αSMA) transcription. Activation of FAK and SFK also results in downstream activation of the Ras-Raf-MAPK-ERK pathway, inducing SRF activation, which in turn activates the transcription factor Sap to upregulate downstream target genes.

Volume 5, April 2015

733

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

actin filaments), all of which colocalized at sites of force application (79).

RhoA/ROCK pathway Mechanosensing between cardiac fibroblasts and their ECM requires “pulling” on adhesions and therefore is reliant upon actomyosin contractility of the cell’s cytoskeleton, which is largely regulated by the RhoA pathway. RhoA is activated in cardiac fibroblasts and numerous other mesenchymal cell types by mechanical stretch or when cells are cultured on rigid substrates (379, 562). Though initially decreased in response to integrin-mediated adhesion to a rigid substrate, sustained adhesion activates several GEFs, which in turn induce an increase in RhoA activity (18, 141, 293, 405). Applying tension directly to integrins (isometric tension) in cells including cardiac fibroblasts causes RhoA activation, confirming its mechanosensory role (185, 588). Numerous lines of evidence suggest that RhoA activation and subsequent actin polymerization is mediated by two pathways (see section on Stress fibers, below), that is, the ROCK-cofilin-LIM Kinase pathway, and via the effector mDia1 (228, 588).

Cell cycle genes The downstream effects of mechanosensing are not limited to changes in cell motility, shape, or adhesion, but may also affect cell survival by altering the expression of cell cycle genes. In floating collagen gels, human foreskin fibroblasts responded to decreased stiffness of their environment by increasing their expression of p27Kip1, while activation of the MAPK pathway mediator ERK was reduced, as was cyclin D1. In MEFs exposed to increased substrate stiffness, autophosphorylation of FAK was increased, leading to Rac activation and Rac-dependent cyclin D1 transcriptional activation (256). Cyclin D1 is implicated in the suppression of adhesion and induction of fibroblast migration. Cyclin D1null MEFs exhibit increased adhesion and reduced motility, and these cells show increased RhoA and ROCK activity, accompanied by increased phosphorylation of LIM Kinase, cofilin, and MLCK (289).

Immediate early genes SRF has been shown to be particularly sensitive to mechanical changes in the ECM (561). SRF is a transcription factor that binds CArG (C AT-rich G) box elements (or SRF response elements, SREs) of its target genes. The SRE contains two sites that are critical to transcriptional activation, including the ternary complex factor (TCF) site, which confers fibroblast responses to varying degrees of cell-matrix adhesions. With increased substrate rigidity, expression of SRF increases, stimulating JNK-mediated phosphorylation of the transcription factor Sap (561). Sap then binds the TCF, initiating transcription of SRF-regulated genes such as Fos and Egr-1. Conversely, on low-rigidity substrate, SRF

734

Comprehensive Physiology

expression is inhibited via p38-dependent phosphorylation of the transcriptional repressor Net, which binds the TCF region of SRF target genes and inhibits their transcriptional activation by this pathway (561).

MRTF-A Myocardin and its related transcription factors MRTF-A and -B are expressed in numerous cell types and act during mechanosensing to convey stimulatory signals from RhoA and the actin cytoskeleton (329). In the nonstressed fibroblast, MRTF-A associates with and sequesters G-actin monomers, preventing assembly into F-actin filaments (329). In response to mechanical stress and RhoA/ROCK activation, MRTFA is released from actin filaments and actin polymerization proceeds (20, 177, 217, 301, 307, 427, 588). Additionally, MRTF-A is translocated into the nucleus where it binds SRE CArG boxes of mechanoresponsive genes including SRF and αSMA, thereby regulating major components of fibroblast phenotype conversion (19, 80, 329, 561). Upon activation of SRF, MRTF-A forms a complex with SRF, which in turn binds and transactivates the αSMA promoter (217, 329, 588). In NIH-3T3 cells subjected to mechanical force, FAK was necessary for MRTF-A-induced αSMA activation (Fig. 3) (79). Evidence indicating αSMA is regulated by various forceinduced pathways suggests it plays a central role in mechanotransduction and subsequent morphological and functional changes in cardiac fibroblasts (537).

Cardiac Injury and Repair Following injury such as MI, extensive myocardial cell death triggers a wound healing process consisting of three phases: an initial inflammatory response, a proliferative phase marked by fibroblast proliferation and migration to the site of injury, and a remodeling phase that largely overlaps with the proliferative phase, and which is marked by extensive cardiac ECM remodeling, fibroblast-to-myofibroblast phenotype conversion and de novo ECM synthesis (104, 116). These responses occur over the minutes, hours and days following the injury, and serve to mitigate further exacerbation of cardiac damage as well as to prevent aneurysm. The beginning of the inflammatory phase is marked by activation of the innate immune system in response to cardiomyocyte necrosis in the injured area. A myriad of injury signals, including cytokines, growth factors, cellular debris, and matrix fragments activate the inflammasome, a multiprotein complex located within the cytosol of resident cardiac fibroblasts. The inflammasome contains apoptosis-associated speck-like protein containing a CARD, which recruits caspase-1. In turn, caspase-1 cleaves pro-IL-1β to produce IL1β as well as IL-18. This initial cytokine production by fibroblasts facilitates the infiltration of neutrophils, leukocytes, and mononuclear cells that mediate the inflammatory response, of which IL-1β is a major mediator (164, 165, 244, 317, 319).

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

In cardiac fibroblasts, IL-1β inhibits fibroblast phenotype conversion to avoid premature wound healing (433). The main mechanism by which IL-1β produces this effect appears to be through inhibition of TGFβ signaling, abrogating TGFβ-induced αSMA expression. Two mechanisms of IL-1β repression of TGFβ signaling in cardiac fibroblasts have been elucidated: upregulation of BAMBI, a transmembrane molecule that inhibits TGFβ receptor heterotetramerization; and downregulation of endoglin, a TGFβ coreceptor (367, 433). A recent study reported that BAMBI overexpression protected mice from pressure overload by restraining TGFβ signaling, while BAMBI deletion resulted in an increase in TGFβ-mediated profibrotic effects in isolated cardiac fibroblasts (529). IL-1β also stimulates cardiac fibroblasts to produce MMP3 and MMP8, facilitating clearance of existing ECM and infiltration of inflammatory cells, while inhibiting cardiac fibroblast proliferation. Through negative regulation of cyclins and cyclin-dependent kinases (CDKs), IL-1β prevents fibrotic responses from occurring before the dead or injured tissue has been cleared from the site of cardiac injury (265, 374, 433). At the end of the inflammation phase, macrophages phagocytose neutrophils and other inflammatory mediators and secrete TGFβ and IL-10 (89,447). Expression of the IL-1β receptor IL-1R1 is attenuated, helping to bring the inflammatory phase to an end (433). The proliferative phase, which is initiated after the inflammatory phase has begun but prior to its resolution, is characterized by proliferation of fibroblasts and migration of these cells to the site of injury. The remodeling phase may occur shortly thereafter or concomitantly, and features conversion of fibroblasts into myofibroblasts, the primary matrix-producing cells within the infarct scar (Fig. 1) (105, 555). Leukotrienes, IL-1, CT-1, TGFβ, and FGF all participate in attracting fibroblasts to the infarcted region (38,137,164,167,330). There, they are responsible for inducing synthesis and secretion of ECM and contractile proteins for scar formation (105, 166). Macrophage-produced IL-23 has been found to be an important mediator in wound healing, and its receptor is upregulated within the infarcted area of the myocardium (528). In cardiac fibroblasts, IL-23 signals through STAT3 to mediate interstitial matrix deposition and capillary growth (205, 430). STAT3 activation also aids in fibroblast phenotype conversion in a process mediated by oncostatin M (347). IL-23 knockout animals exhibit decreased collagen synthesis, αSMA expression and activation of STAT3 and ERK1/2. The importance of this cytokine in mediating scar formation is underscored by the observation that IL-23 deficiency following MI ultimately results in fatal ventricular rupture (430). Following MI, scar tissue is created from cross-linked collagen deposited by myofibroblasts during the proliferative phase of wound healing, and thins during subsequent maturation. There is a substantial increase in collagen type I and type III production and accumulation in the infarct scar (105). AngII, which increases in response to various types of cardiac injury, plays an active role in profibrotic signaling within the myofibroblast where it signals through AT1 receptors to

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

induce TGFβ1 expression and to enhance TIMP-1 production, promoting collagen accumulation and scar formation within the affected area (74, 102). After the mature scar is formed, growth factors and profibrotic regulatory proteins are cleared from the area and some myofibroblasts undergo apoptosis, resulting in their clearance from the infarct scar area (100, 164, 493). However, it has been shown that myofibroblasts can persist in the infarct scar within the myocardium for years, contributing substantially to cardiac fibrosis (555,583). Different types of fibrosis can occur within the heart, depending on the affected area and the specific pathological circumstances. Interstitial and perivascular fibrosis, the latter of which occurs in the adventitia of the coronary arteries, do not involve cardiomyocyte necrosis and can be triggered in the absence of cardiac events such as MI (48, 548). Conversely, focal fibrosis, which is sometimes referred to as reparative or replacement fibrosis, occurs as a result of cardiomyocyte necrosis and is usually preceded by a cardiac event (434, 549). The remodeling performed by myofibroblasts that resist apoptosis and persist for prolonged periods of time contributes to detrimental changes in systolic and diastolic function and decreases overall compliance of the myocardium (51, 108, 550). Increased active tension with a decrease in relaxation time has also been found in the hearts of spontaneously hypertensive rats with cardiac fibrosis (108). These changes in heart function contribute to impaired contraction and ultimately heart failure (322).

Cardiac Fibroblast Migration Migration of cardiac fibroblasts to sites of myocardial injury is a critical step in the early reparative remodeling process. External cues such as mechanical strain and biochemical stress mediators stimulate this coordinated cell movement, which involves the assembly and disassembly of the actin cytoskeleton, accompanied by turnover of cell-matrix adhesions. The dynamic nature of the actin cytoskeleton links changes in cell architecture at the leading edge of the cell to the formation, maturation, and degradation of cell-matrix adhesions. Factors involved in regulating directed cardiac fibroblast migration include the topography and stiffness of the ECM, cell polarity, receptor activation including integrins and their clustering, as well as cytoskeletal contraction. Migration of cardiac fibroblasts is induced primarily by injury-related proinflammatory cytokines (Fig. 4). The matrix component fibronectin and its truncated form (called migration stimulating factor) are capable of inducing fibroblast spreading and migration (154, 442). AngII appears to induce cardiac fibroblast migration through a NADPH oxidase 4 (Nox4)-ROS-dependent ERK pathway leading to activation of transcription factors NF-κB, Activator Protein-1 (AP-1) and Sp1, which induce MMP activation and downregulation of the adhesion protein REversion-inducing-Cysteine-rich protein with Kazal motifs/RECK (460, 491, 542). IL-8 stimulates a similar pathway via Nox4 and ROS with a

735

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Immune cells

Matrix remodeling and migration

Adipocytes Leptin

Comprehensive Physiology

IL-17

Adiponectin Pro-MMP2

ObR

IL-8 and IL-18 AngII

MMP2 Interleukin receptor

AdipoR MT1-MMP

MMPs

AngII receptor

AMPK Nox4 APPL1

P AMPK

Akt

Nox4

RhoA ROS MAPK

ROCK

ROS

MT1-MMP

ERK

Cofilin

ERK

F-actin

TRAF3IP2 Sp1

NF-kB Actin remodeling

Activation of MMPs

AP-1

NF-kB

AP-1

sm Cytopla s

Nucleu

RECK AP-1

Sp1

Decreased adhesion

Figure 4

Signaling pathways governing cardiac fibroblast migration. Fibroblast migration requires both intracellular actin remodeling to drive motility, and the extracellular release of remodeling enzymes to facilitate cell movement through the surrounding stroma. A host of extracellular ligands act on cardiac fibroblasts to induce migration, acting via distinct pathways to govern both activities. Adipocyte-released cytokines leptin and adiponectin act on their respective receptors in cardiac fibroblasts to induce migration. Leptin-induced signaling activates the RhoA-RhoGTPase-associated kinase (ROCK)-cofilin pathway. Cofilin severs actin filaments, inducing actin remodeling and reducing F-actin content. Leptin signaling also causes activation and relocation of type-I transmembrane matrix metalloproteinase (MT1-MMP) to the cell surface, where it can cleave and activate pro-MMP2, causing degradation of extracellular matrix components and increased migration. Adiponectin also activates MT1-MMP downstream of the adaptor protein, phosphotyrosine interaction, PH domain and leucine zipper containing-1 (APPL1)-AMP-activated protein kinase (AMPK) signaling cascade. Binding of angiotensin II (AngII) and interleukin-8 and -18 to their respective transmembrane receptors causes activation of NADPH oxidase-4 (Nox4), resulting in increased reactive oxygen species (ROS) production. AngII-induced ROS production activates the transcription factors Nuclear Factor κB (NF-κB), activator protein-1 (AP-1), and specificity protein-1 (Sp1) through extracellular signalregulated kinase (ERK) signaling. Through this pathway, NF-κB and AP-1 induce activation of various MMPs, and both Sp1 and AP-1 repress activity of the adhesion protein reversion-inducing-cysteine-rich protein with kazal motifs (RECK). Thus, AngII signaling induces cardiac fibroblast migration by increasing matrix degradation and reducing cell adhesion. Interleukin-8 and -18 also utilize Nox4-ROS signaling, but with a tumor necrosis factor receptor-associated-3-interacting protein-2 (TRAF3IP2) intermediate. Activation of AP-1 and NF-κB downstream of these interleukins induces MMP activity, increasing cardiac fibroblast migration. It is likely that additional pathways identified in noncardiac fibroblasts are also involved in cardiac fibroblast migration, however they remain to be characterized.

TRAF3-interacting protein-2/TRAF3IP2 intermediate (147). IL-18 also utilizes Nox4 signaling, resulting in NF-κB/AP1-induced MMP activation (525). In kidney myofibroblast activation, RhoA/ROCK signaling has been shown to act

736

upstream of Nox4 and ROS production (311). IL-17 also stimulates MMP activation and cardiac fibroblast migration, though this appears to occur through Akt/miR-101/MKP1-dependent activation of the p38 MAPK/ERK pathway,

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

causing downstream activation of NF-κB, AP-1, and CCAAT enhancer-binding protein (111, 526). Interleukins have also been shown to stimulate migration in fibroblasts from other tissues through these and other nuclear proteins (248, 345). In a rat model of MI, expression of the cytokine cardiotrophin-1 was observed to advance from the border zone alone at 1 day post-MI to the infarct zone within 4 days, concomitant with rapid phosphorylation of Janus kinase, Jak2, STAT1, STAT3, p42/44 MAPK, and Akt (168). Adipocyte cytokines are also capable of stimulating cardiac fibroblast migration. The metabolic regulator adiponectin induces cardiac fibroblast migration facilitated by enhanced MMP activity through Adaptor protein, Phosphotyrosine interaction, PH domain and Leucine zipper-containing 1/APPL1mediated phosphorylation of 5’ AMP-activated protein kinase (AMPK). AMPK activation results in secretion and activation of MMP14 (MT1-MMP) (119). Similarly, leptin induces cardiac fibroblast migration by increasing expression and cell surface localization of MMP14 via RhoA/ROCK-dependent reorganization of the actin cytoskeleton involving the actin disassembly protein cofilin (443, 444).

Cell polarization In the absence of stimuli, cardiac fibroblasts on tissue culture plates exhibit a stellate shape with an evenly distributed actin meshwork and cell-matrix adhesions. For efficient migration, cells must become polarized, developing a leading edge (lamellipodia extending in the direction of migration) and a trailing edge (retracted by retrograde actin flow). In addition to chemical factors, fibroblast polarization is induced by stiff substrates in a protein tyrosine kinase-regulated manner (398). In preparation for movement, the polarization process involves asymmetric distribution of various signaling proteins and actin cytoskeleton reorganization. Besides actin filaments, microtubules comprised of polymerized tubulin monomers play an important role in fibroblast polarization. In fibroblasts, the activity of cell division control protein 42 homolog (Cdc42) is highest at the tip of the leading edge, and Rac1 activity is highest immediately behind the leading edge (172). Rac and Cdc42 regulate the formation of FA complexes at the lamellipodia, and in fibroblasts Cdc42 has been implicated in rapid activation of Rac1, followed by RhoA activation (358). Rac1 is necessary for the formation of lamellipodia, and Cdc42 is similarly necessary for the formation of filopodia, the initial protrusions in directed fibroblast migration (266,410,481). Interaction between Cdc42 and Rac is involved in the integration of these membrane protrusions and coordinated movement (358). Cdc42 and its downstream target N-WASP are both necessary to establish and maintain fibroblast cell polarity, the former acting specifically in migrating primary rat embryonic fibroblasts to regulate lamellipodial activity and reorientation of the Golgi apparatus (325, 357). N-WASP is an activator of actin polymerization and regulates actin-based processes, including cell protrusion, at least

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

in part by activation of the Arp2/3 complex (331). FAK, NWASP, and the mechanosensory protein p130Cas form a complex at the distal edge of protruding lamellipodia in spreading fibroblasts (587). The activity of Src-family kinases, including Src, Yes, and Fyn, is induced by fibronectin-activated integrins in the plasma membrane. Subsequently, Src-family kinases phosphorylate p130Cas within the Cas-FAK-N-WASP complex, inducing downstream activation of Rac1 and actin polymerization within the lamellipodium (451). Microtubule stabilization in migrating fibroblasts occurs, at least in the case of lysophosphatidic acid-induced motility, via RhoA and its downstream effector mDia1, though these proteins are not restricted to the lamellipodia, indicating the need for additional regulatory factors (109, 372, 373). The Arp2/3 complex facilitates actin polymerization at the leading edge of migrating fibroblasts. Studies with fibroblasts lacking the ARPC3 gene, which encodes the p21 subunit of this complex, have provided evidence of a critical role for Arp2/3 in polarization and lamellipodia protrusion (487). ARPC3−/− fibroblasts are unable to extend lamellipodia, generating leading edges primarily composed of filopodia exhibiting concentrated mDia1 at their tips. Additionally, these cells lack the ability to coordinate protrusion and directional migration, indicating a role for Arp2/3 (and potentially mDia1) in establishing and maintaining cell polarity. In fibroblasts expressing a dominant negative form of FAK, cell spreading is impaired and cells lack a continuous lamellipodium. The dominant negative FAK also causes reduced phosphorylation events in FAK at Y397 and p130Cas at Y165 (587). Studies in various fibroblast lines have provided evidence that FAK, but not Cas, Src, Fyn, or Yes, is necessary for microtubule stabilization and acts upstream of RhoA and mDia1 to support this function (373). Although some cells utilize mechanisms such as blebbing for extending their membranes, the primary mode by which mesenchymal cells migrate involves the formation of membrane extensions including lamellipodia and filopodia (also referred to as membrane ruffling). Membrane ruffling in NIH-3T3 fibroblasts is inhibited by a dominant-negative mDia1 mutant (515). Inhibition of p130Cas phosphorylation induces membrane ruffling, though there is evidence for FAK C-terminal binding of p130Cas via a SH3 domain, promoting migration through coordinated Rac activation in the lamellipodia (331,515). This suggests FAK may act to sequester and prevent activation of p130Cas, which is involved in stabilization of adhesions and may be suppressed during cardiac fibroblast migration. Part of the polarization process involves the development of lipid rafts at what will be the leading edge of the cell, enabling Rac1 recruitment to the lamellipodia. The development of lipid rafts is regulated by FAK and/or integrin signaling (126). Using the lipid raft marker ganglioside GM1 , Palazzo et al. demonstrated a lack of lipid rafts in fibroblasts null for FAK, but not for SYF or Cas (373). Thus it appears that an integrin-FAK pathway enables the formation of a lipidrich leading edge, as well as signaling by RhoA and downstream mDia1, resulting in microtubule stabilization during

737

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

polarization of migrating fibroblasts. Recent work by Carisey et al. has also indicated vinculin, interacting through its tail region with the actin cytoskeleton, is necessary for polarization in migrating MEFs (77).

Lamellipodia protrusion and adhesion turnover Lamellipodia are sheet-like structures composed of an actin meshwork that terminates on focal complexes, and filopodia are finger-like protrusions with thin parallel actin bundles (358). Just behind the lamellipodia lies the lamellum, containing SF-like actin filaments and transverse arcs which, along with microtubule growth, provide the driving force for lamellipodial protrusion (471, 558). Downregulation of the Rho-kinase pathway is involved in adhesion turnover during fibroblast migration. Blocking the Rho pathway in NIH-3T3 fibroblasts causes dissolution of existing FAs and stimulates the formation of new complexes, in addition to stimulating membrane ruffling (418). These processes are associated with advancement of the cell front during migration. The GTPase Rac1 antagonizes RhoA activity in most situations, and its activation appears to be a critical component in adhesion turnover during cardiac fibroblast migration. Rac1 activation is induced by microtubule growth, and its activity is required for lamellipodial protrusion (558). Additionally, Rac can activate Rho, inducing the formation of new adhesion sites and the assembly of contractile filaments within the advancing lamellipodium (358). The cellular pathways by which Rac activity makes a switch from RhoA suppression to RhoA activation remain unknown, but likely involve signaling crosstalk. The p21-activated serine-threonine kinase (PAK)interacting exchange factor PIX is a GEF enriched in Cdc42/Rac1-induced adhesions, acting upon Rac1 to enhance its activity (81, 498). PIX associates with and is required for recruitment of PAK to sites of adhesions (313). In CHO cells, the localization of PAK at the plasma membrane of migrating cells is critical to cytoskeletal reorganization involved in directed motility (53). In fibroblasts, there is evidence of Cdc42-induced activation of a PAK/PIX complex, leading to Rac1 activation and adhesion turnover (313). Additionally, PAK interacts with the adaptor protein noncatalytic region of tyrosine kinase adaptor protein 1 (Nck1) (313,589). PIX also binds to the G-protein-coupled receptor kinaseinteracting protein GIT1, which can promote adhesion disassembly even in the absence of actin-myosin contractility (589). The C-terminal domain of GIT1 interacts directly with paxillin, and GIT1 overexpression results in delocalization of paxillin from cell-matrix adhesions (589). In turn, paxillin has been shown to mediate actin redistribution by recruiting the PAK/PIX complex, via interactions with p95PKL (paxillin-kinase linker), to nascent focal complexes (519). Moreover, the process of PAK/PIX-mediated Rac1 activation and adhesion turnover during fibroblast migration appears to also involve FAK. GIT has been shown to directly bind FAK via a conserved Spa2 homology domain, and FAK (but not

738

Comprehensive Physiology

SFKs) directly associates with and tyrosine phosphorylates PIX (81, 589).

Focal adhesion kinase in migration FAK is a tyrosine kinase expressed in many cell types including cardiac fibroblasts, and is recruited to and involved in formation of cell-matrix adhesions (331). However, its primary role is in the disassembly (or “turnover”) of FAs, a process required for cell motility and the production of tensile forces (462). FAK-null fibroblasts demonstrate excessive adhesion formation and reduced migration in response to chemotactic stimuli (223). Activation of FAK is induced by autophosphorylation at Y397, in response to either integrin clustering or force applied to matrix adhesions (47, 378). In the latter, RhoA/ROCK signaling appears necessary, as its blockade prevents phosphorylation of Y397 and thus prevents FAK activation (511). FAK is composed of an N-terminal FERM domain, a central kinase domain, numerous proline-rich regions, and a C-terminal FA-targeting (FAT) domain (331). The FERM domain has been shown to link signals from receptor tyrosine kinases (RTKs) such as epidermal growth factor receptor (EGFR) and PDGF Receptor, and its overexpression blocks activation of FAK and inhibits cell migration (196, 461, 485). Two proline-rich regions in the C-terminal domain of FAK contain Src-homology 3 (SH3) domain binding sites that allow its interaction with adaptor proteins (331). Though SH3-mediated binding of GTPase regulator associated with FAK (GRAF) and Arf GTPase-activating protein containing SH3, ankyrin repeat and pleckstrin homology (PH) domains1 (ASAP1) proteins to the FAK C-terminal domain plays a role in mediating actin cytoskeleton dynamics and FA assembly, the downstream interactions of these two proteins remain unknown (378). The C-terminal domain also contains the FAT region, which mediates colocalization of FAK and integrins at FAs (331). Initially, it was presumed that FAK bound directly to the cytoplasmic tail of integrins, however more recent evidence indicates that FAK associates with cell-matrix adhesions indirectly through complex formation with paxillin and talin (331). The ability of integrins to induce FAK phosphorylation is dependent upon interaction between the FAT domain of FAK and talin (84). FAK phosphorylation (Y925) induces recruitment of Grb2 and promotes activation of the Ras/Raf/MEK/ERK pathway (438,440). In turn, Ras regulates turnover of adhesions and actin filaments essential for fibroblast motility (357). ERK and MEK are downstream modulators of Ras activity and blockade of Ras in wounded rat embryonic fibroblasts results in persistent cell-matrix adhesions and lack of motility, indicating that Ras is required for adhesion turnover during fibroblast migration (357).

Retraction of the trailing edge With the protrusion of the lamellipodia and the formation of new attachments at the leading edge, tension is generated

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

within the cell, mediated at least in part by myosin II-induced contractility (6,67,334,411). The formation of integrin-based nascent adhesions and F-actin polymerization at the lamellipodium is necessary to cause retraction of the trailing edge of cardiac fibroblasts. The actin network in lamellipodia and lamella of migrating cardiac fibroblasts is far from stationary with respect to their substrate. In a process termed “retrograde flow,” the pressure of actin polymerization at the leading edge and contractility within the cell results in a net movement of the actin cytoskeleton away from the leading edge of the cell (6, 113). In the trailing edge, actin filament disassembly provides the majority of the force required for retraction (113). Combined retrograde actin flow and myosin II activity appear to be important in the development of new adhesions in fibroblast lamella regions (10). In addition to F-actin, adhesion proteins such as zyxin and VASP also undergo retrograde flow at regions of cell-matrix adhesion in the lamella of migrating fibroblasts (107, 188). The lamellipodium is characterized by more rapid actin turnover and retrograde flow (which involves the Arp2/3 complex and downstream actin depolymerizing factor cofilin) than the more central lamellum, which contains more mature adhesions and harbors contractility components such as tropomyosin and myosin II (6, 216, 261, 394). Activation of Rac at the leading edge induces lamellipodia formation, and Rac directly activates WASP-family verprolin-homogolous protein (WAVE), which in turn causes Arp2/3 activation (326, 371, 494). WASP and WAVE proteins bind to the actin monomer-recruiting protein profilin; however, studies of these interactions have not yet been completed in cardiac fibroblasts (376). The inhibitory protein Arpin also acts at the lamellipodium downstream of Rac to inhibit Arp2/3 and control directionality during fibroblast migration (120). Thus, a balance appears to exist between Rac-Arpin-Arp2/3 inhibition signaling and Rac-WAVE-Arp2/3 activation signaling, though how these pathways are regulated in the cardiac fibroblast remains unknown. Vinculin, a protein found in focal complexes and adhesions, also appears to regulate retrograde actin flow by coupling the actin cytoskeleton to cell-matrix adhesions and acting as a molecular clutch. Vinculin expression in MEFs slows F-actin flow and promotes the development of nascent adhesions, while inhibiting their maturation, within the lamellipodium (501). In cells with disrupted vinculin-actin binding, establishment of the lamellipodium-lamellum border and generation of traction forces are inhibited.

Cardiac Fibroblast Proliferation The numerous players involved in cell proliferation and progression of the cell cycle have been well characterized in general, and are beyond the scope of this review. However, excessive proliferation of cardiac fibroblasts, coupled with failure to remove fibroblasts and/or myofibroblasts from healing sites of injury, contributes to fibrosis; thus a brief overview

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

is appropriate. In eukaryotes, the cell cycle consists of four phases separated by a system of checkpoints to govern cell progression to the next phase. If the cell is damaged, for example, it will undergo cell cycle arrest at a specific checkpoint, and will not proceed further through the cycle (152,193,354). Along with the system of checkpoints, the cell cycle is also mediated by serine/threonine kinases called CDKs, which are activated through the binding of cyclin proteins and participate in signaling mechanisms that guide the cell through to the next phase of the cycle. CDKs are also regulated by phosphorylation, subcellular localization within the nucleus, and CDK inhibitors (339-341, 456). The cell cycle begins in the first gap phase (G1 )—the longest phase within the cycle. While the cell is in G1 , the activated CDK4/Cyclin D complex phosphorylates the retinoblastoma protein (Rb) to form phospho-Rb. Late in G1 , the CDK2/Cyclin E complex further phosphorylates phosphoRb, which is then able to activate E2F transcription factors to induce the expression of genes required for the next phase (151, 193, 300, 352, 353, 455, 456, 553). The cell then enters the S phase and DNA synthesis commences. In S phase, the CDK2/Cyclin A complex is activated, leading the cell to the second gap phase G2 (193, 483, 567). During the G2 phase, the Cdc2/Cyclin B complex becomes active and permits progression to the mitotic phase and cell division (193, 255). While this general mechanism holds true for most eukaryotic cells, different cells may employ different regulatory mechanisms. In cardiac fibroblasts, several proteins and factors have been found to play regulatory roles in the cell cycle and cellular proliferation. For example, ERK 1/2 is activated by growth factor-mediated intracellular signaling, and mediates proliferation, differentiation and apoptosis of cardiac fibroblasts (323, 368, 568). ERK 1/2 has been shown to enhance proliferation by indirect downregulation of the CDK inhibitor p27 in response to mitogen signals (127). ERK 1/2 upregulates S-phase kinase-associated protein 2 (Skp2), which negatively regulates p27 through post-translational mechanisms, and also downregulates forkhead box O 3a (FOXO3a), which normally enhances transcription of p27. Through this reciprocal regulation of Skp2 and FOXO3a, ERK1/2 promotes cardiac fibroblast proliferation (399). AngII induces proliferation of adult cardiac fibroblasts via activation of ERK1/2. AngII binds to the AT1 receptor, activating PLCγ which then cleaves membrane bound phosphatidylinositol-3,4-bisphosphate to produce inositol 1,4,5-trisphosphate (IP3 ) and diacylglycerol (DAG). IP3 facilitates the mobilization of intracellular calcium stores while DAG activates PKCδ (96,112,365,419). AngII signals specifically through the PKCδ isoform. Both of these pathways operate in parallel to activate ERK1/2. Inhibition of both pathways simultaneously is needed to significantly attenuate ERK1/2 activation, illustrating the importance of this convergence (365). AngII-induced ERK1/2 activation is not PI3Kdependent and does not involve EGFR transactivation or c-Src within the signaling pathways. This further illustrates how different adult cardiac fibroblasts are from other cell types

739

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

including neonatal cardiac fibroblasts, which utilize EGFR transactivation in their AngII-induced proliferation pathway (365). PDGF is an important inducer of proliferation in cardiac fibroblasts. The PDGF-BB isoform stimulates the MAPK signaling cascade through a PKC-dependent pathway to regulate DNA synthesis and thus proliferation. PKC can either activate Raf1 directly through phosphorylation or indirectly via activation of Ras. Raf1 in turn phosphorylates ERK1/2 to induce proliferation (39, 382). Despite the widespread heterogeneity of fibroblasts across tissues, experiments to confirm the details of known cell signaling mechanisms specifically in cardiac fibroblasts are relatively uncommon. However, both PDGF-AA and PDGF-BB have been shown to induce early genes such as c-myc and c-fos, indicators of proliferation induction, through the activation of MAPK (464). A number of growth factors such as PDGF and FGF bind to RTKs to induce proliferation through Akt in myriad cell types. RTKs recruit the p85 and p110 PI3K subunits to form the active enzyme, which in turn catalyzes the conversion of PIP2 to phosphatidylinositol-3,4,5-trisphosphate (PIP3 ) (230, 298,472). PIP3 recruits Akt to the plasma membrane enabling phosphorylation by PDK1, resulting in both Akt activation and translocation into the nucleus (5, 566). In the nucleus, p70S6 Kinase (p70S6K) is activated downstream of phosphoAkt. p70S6K then phosphorylates the S6 protein in the 40S ribosomal subunit to mediate cell growth and proliferation (Fig. 5) (58, 464). PI3K may also associate with the AT1 receptor, allowing AngII to promote Akt phosphorylation as well (538). Hepatocyte growth factor (HGF) is one of several factors that act as an antifibrotic cytokine within the heart. HGF binds within the Sema domain of its receptor c-Met. Ligand binding induces dimerization of two receptor molecules, leading to its activation and phosphorylation (235, 263, 478). In cardiac fibroblasts, HGF appears to negatively regulate proliferation through induction of the c-Met-Akt-TGIF pathway: TGIF, a Smad transcriptional corepressor downstream of c-Met, inhibits proliferation by blocking the nuclear translocation of phospho-Smad2/3 to attenuate TGFβ signaling (297,574). Loss-of-function studies have supported this mechanism in cardiac fibroblasts. Knockdown of HGF induces expression of Proliferating Cell Nuclear Antigen, Ki67 and Cyclin D1, which increases the overall TGFβ-induced proliferative response (574). HGF downregulation also has an inhibitory effect on the phosphorylation of c-Met and Akt (574). Cardiac fibroblasts can also respond in an autocrine fashion to exogenous signals to induce proliferation. For example, evidence of the cAMP-adenosine pathway has been found in cardiac fibroblasts through the use of specific enzyme inhibitors (142). This pathway is induced by exogenous factors such as hormones, resulting in stimulation of adenylyl cyclase to produce cAMP which can itself signal or be converted intracellularly by phosphodiesterase to AMP, which is then converted into adenosine by 5′ -nucleotidase. Adenosine can be transported into the extracellular space via facilitated

740

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

transport mechanisms, where it activates A2 receptors in an autocrine fashion (41, 142). The overall effect of A2 receptor activation is antimitogenic, attenuating cellular growth and proliferation (143, 144). This pathway may also take place more efficiently in the extracellular space catalyzed by ectophosphodiesterase and ecto-5′ -nucleotidase, since adenylate kinase competes for AMP within the cytosol (142). Negative regulators of proliferation in cardiac fibroblasts may serve a cardioprotective role at times of myocardial injury or hypertrophy, protecting against pathological persistence of fibroblasts within the heart (142). Conversely, dysregulation of this finely tuned, intricate web of proliferative signals in cardiac fibroblasts may contribute to fibrotic disease progression. It is also important to consider, however, that fibroblast proliferation represents an important step of the reparative process in the infarct region after MI. Thus, attempts to regulate fibrosis by targeting fibroblast proliferation must consider the nature of the underlying disease and reparative processes. Even in the case of MI, modulation of fibroblast proliferation may be viable since distal fibrosis typically occurs after the infarct scar has formed (116).

Fibroblast-to-Myofibroblast Phenotype Conversion Fibroblasts are not only responsible for the secretion and maintenance of ECM components, but also in regulating the transmission of mechanical and electrical stimuli within the healthy working heart (72, 425). Upon myocardial injury, in response to stress including hypoxia or cell stretch, or following stimulation by profibrotic growth factors such as TGFβ, fibroblasts undergo a phenotype conversion to myofibroblasts (Fig. 1) (116). While this conversion is frequently referred to in the literature as a differentiation process, this nomenclature is difficult to rationalize since fibroblasts already clearly exhibit signs of differentiation such as SFs, and it would be difficult to argue that fibroblasts are undifferentiated. Rather, this conversion may better be considered as a “fine-tuning” of the cell phenotype to meet the changing demands of the surrounding tissue. There are multiple factors and interconnecting pathways that modulate fibroblast-to-myofibroblast phenotype conversion. The dysregulation of any one of these regulatory pathways may result in either the overaccumulation of collagens or impaired healing of the injured heart, making phenotype conversion a double-edged sword. Although myofibroblasts are required to heal the injured myocardium following MI, they are also the primary drivers of tissue fibrosis. Thus, these pathways have been the subject of extensive research, and while our picture of the phenotype conversion process is more complete than ever, significant gaps in our understanding remain. For example, there is debate as to whether an intermediate cell type, called the proto-myofibroblast, arises during fibroblast phenotype conversion in vivo, although a number of laboratories have reported the existence of these cells in vitro

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

Intracellular Signaling of Cardiac Fibroblasts

Growth factors (PDGF, FGF)

Receptor tyrosine kinase (RTK) Plasma membrane

P PIP3

PIP2

Akt Pl3Kp85

PDK1

Pl3Kp110 P Cyclin D1

P Akt sm Cytopla

P s Nucleu

Cyclin D1

P Akt

GSK3

Cyclin D1 P70S6 Kinase

40S subunit

P S6 protein

Cell growth and proliferation

Figure 5 Akt-mediated regulation of cardiac fibroblast proliferation. Growth factors such as PDGF and FGF bind to their cognate receptors, resulting in associated receptor tyrosine kinase (RTK) activation. Upon ligand binding, PI3K subunits p85 and p110 are recruited to the receptor to form the activated PI3K enzyme, catalyzing the conversion of PIP2 to PIP3 . PIP3 recruits Akt to the plasma membrane, where it is phosphorylated by PDK1 thereby promoting its translocation into the nucleus. p70S6 Kinase is activated downstream of phospho-Akt and in turn phosphorylates the S6 protein within the 40S ribosomal subunit, positively regulating cell growth and proliferation. Phospho-Akt further promotes cell proliferation by phosphorylating and thereby deactivating GSK3. GSK3 flags Cyclin D1 for translocation into the cytoplasm of the cell, make it unavailable to regulate the cell cycle and resulting in cell-cycle arrest.

(139, 209, 508). The proto-myofibroblast is characterized by increased expression of FA proteins, and the first stages of SF formation. Proto-myofibroblast SFs are composed of βand γ-actins but specifically lack αSMA, despite the fact that αSMA expression is increased in these cells (139, 508).

Volume 5, April 2015

Additionally, proto-myofibroblasts show TGFβ-induced activation of Rho-GTPase and ROCK, facilitating polymerization of G-actin into F-actin. However, while proto-myofibroblasts can be readily detected in vitro, the many technical challenges of studying cell adhesions in vivo leave the existence of

741

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

proto-myofibroblasts during tissue pathogenesis in doubt (209). Myofibroblasts represent an activated form of fibroblast, and are the primary mediators of cardiac wound healing (116). In contrast to fibroblasts, myofibroblasts exhibit mature SFs containing actin and myosin, which unlike protomyofibroblast SFs, also have αSMA incorporated. These SFs exert tension on the ECM and facilitate the production of contractile force by the myofibroblast, yet at the same time oppose retractile force, aiding the physical remodeling of the ECM (72, 121, 145, 425, 508). Phenotype conversion of fibroblasts occurs very quickly and is mainly induced by mechanical tension, deposition of the ED-A splice variant of fibronectin, and TGFβ stimulation, although many other factors have been found to affect this process both positively and negatively (see below) (416, 448, 484). Indeed, the conversion process is easily activated in vitro, providing a challenge for the study of primary fibroblasts once isolated from cardiac tissue. Tissue culture plates that are most commonly used for cell culture are highly rigid, with stiffness several orders of magnitude higher (within the MPa range) than even highly fibrotic tissue (207). As a result, by the time isolated fibroblasts have been expanded in culture prior to experimentation, they have already begun expressing αSMA, and thus have already begun converting to proto-myofibroblasts or myofibroblasts. This can occur even before the very first cell passage, and makes studying these phenotypes challenging (209, 425).

Positive regulation of phenotype conversion In cardiac fibroblasts, TGFβ1 stimulation induces an elevation in collagen production and αSMA expression in a dosedependent manner (66, 132, 388). TGFβ1 induces cardiac fibroblast phenotype conversion to myofibroblasts through canonical Smad signaling (157). There is also evidence to suggest that TGFβ1 signaling converges with the AngII pathway to induce the myofibroblast phenotype. It has been shown that TGFβ1 induces expression of the angiotensin converting enzyme (ACE) protein, which is essential for the production of AngII (389). Pharmacological suppression of AngII by resveratrol or losartan results in the prevention of TGFβ1 induced fibroblast phenotype conversion and the reduction of Smad2 and Smad4 expression, respectively (136, 364). AngII has also been shown to induce TGFβ1 expression in cardiac fibroblasts (74, 284). This cross-talk may behave as a positive feedback mechanism, in which TGFβ1 induces more TGFβ1 production in an autocrine fashion through mediation of ACE. Although the effects of these cytokines on cardiac fibroblasts are well characterized, the precise molecular mechanisms remain to be elucidated. TGFβ has many origins including secretion both from resident cardiac cells and infiltrating inflammatory cells, as well as inactive stores within the tissue itself, emphasizing the importance of this cytokine in many organs including the heart. Generally, TGFβ is synthesized and secreted in a homodimeric protein pro form, consisting of a TGFβ

742

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

molecule covalently linked to latency-associated protein (LAP). The resulting small latent complex (SLC) prevents TGFβ from associating with its receptor (11). TGFβ can also be secreted as a large latent complex (LLC), composed of a SLC linked to latent TGF-binding protein (LTBP) by a disulfide bond (420). LTBPs bind to ECM proteins such as fibrillin-1, vitronectin and fibronectin, creating a TGFβ store within the ECM (11, 441, 490, 497). Latent TGFβ is released as an active cytokine following proteolytic cleavage by enzymes that degrade ECM, including the MMP family (11, 118, 305, 577). Glycosidases release active TGFβ by removing carbohydrates from the SLC, resulting in a conformational change of the complex and TGFβ release (332). Thrombospondin-1 also releases TGFβ from the SLC or LLC through the disruption of the bond between LAP and TGFβ (346). Aside from latent TGFβ already present within the ECM and interstitium of the tissue, TGFβ may also be secreted by activated macrophages during wound healing (21). Neonatal cardiac fibroblasts have also been shown to secrete TGFβ1 and TGFβ2 in response to norepinephrine and AngII stimulation, working in an autocrine and paracrine fashion (160). The endothelin family of biopeptides consists of endothelins 1-3 (ET1-3). The majority of ET1 is generated by endothelial cells and is produced in the heart, kidney, pituitary, and central nervous system (225, 226). ET1 is a potent vasoconstrictor and is known to increase blood pressure and vascular tone (4). ET1 is also involved in wound healing postmyocardial injury, and leads to phenotype conversion of cardiac fibroblasts to myofibroblasts (356). One laboratory has reported that overexpression of TRPC6 or constitutively activated NFAT in cardiac fibroblasts attenuated the effect of ET1 and reduced fibroblast-to-myofibroblast conversion (270). In sharp contrast, however, more recent studies have implicated TRPC6 as a profibrotic factor: deletion of the TRPC6 gene impaired both cardiac and dermal wound healing, and TRPC6 was sufficient to increase type I collagen expression and fibroblast phenotype conversion via induction of calcineurin, the Ca2+ -dependent phosphatase that activates NFAT via dephosphorylation (122). Haploinsufficiency of the TGFβ coreceptor endoglin attenuated activation of the calcineurin/TRPC6 pathway and reduced expression of αSMA (236). Other TRP channels such as TRPC3 and TRPM7 have also been implicated in the phenotype conversion of fibroblasts to myofibroblasts and in atrial fibrosis, and the common mode of action of these channels appears to be facilitating entry of Ca2+ to activate calcineurin/NFAT signaling (502, 578). The remodeled matrix itself also contributes to fibroblast phenotype conversion. As the ECM is remodeled and disrupted, fibroblast protection from mechanical stress is reduced, resulting in fibroblast phenotype conversion through the Rho/ROCK signaling cascade (508, 588). Through this pathway, mechanical force may induce αSMA expression and ultimately induce fibroblast to myofibroblast phenotype conversion.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

There is growing evidence that microRNAs can play a role in fibroblast phenotype conversion. MicroRNAs suppress protein expression by either facilitating mRNA degradation or by inhibiting translation (29, 50). One such example is miR145, which has been reported to induce the myofibroblast phenotype similar to TGFβ. miR-145 overexpression upregulated the myofibroblast markers fibronectin, vimentin, and αSMA as well as connexin 43, and downregulated the fibroblast marker DDR2 in cultured fibroblasts, indicative of a shift toward the myofibroblast phenotype. The cells also formed highly organized SFs and F-actin bundles, providing contractility, as well as enhanced deposition of mature collagen type I (543). It has been shown in vascular smooth muscle cells that miR-145 regulates cellular phenotype through the negative regulation of KLF5, which itself is a negative regulator of myocardin (94). Wang et al. reported that this pathway exists in cardiac fibroblasts. At 3 days following MI, miR-145 was downregulated, alleviating its effect on KLF5 expression and resulting in decreased myocardin expression. At 7 days post-MI, miR-145 expression returns to basal levels, reducing KLF5 expression and in turn upregulating myocardin (94,543). In contrast, however, recent work has suggested that myocardin is not endogenously expressed by cardiac fibroblasts (470). This discrepancy could be explained by the fact that myocardin and MRTFs are very closely related and share homology in multiple functional domains [as reviewed by Pipes et al. (393)], thus it may be possible for MRTFs to be detected using supposedly myocardin-specific primers. Both proteins are associated with SRF (535). Myocardin is primarily localized within the nucleus whereas MRTFs are bound to G-actin via their N-terminal RPEL domain (269, 470). As described above, ROCK signaling reorganizes actins within the cytoplasm and releases the MRTFs, making them available to translocate into the nucleus where they enhance smooth muscle gene expression (304, 329, 588). Small et al. found a 50% reduction in scar size within the hearts of MRTF-A null mice (470). This was accompanied by attenuated upregulation of SM22, αSMA, Col1α1, Col1α2, Col3α1, and elastin post-MI, consistent with a diminished myofibroblast phenotype. However, the MRTF-A null mice did manage to create sufficient scar to prevent cardiac rupture. This research group also found that inhibition of ROCK signaling with Y27632 partially attenuated MRTF-A translocation and TGFβ1 -mediated collagen synthesis, providing evidence of a new potential treatment option to prevent excessive fibroblast differentiation without risking cardiac rupture.

Negative regulation of phenotype conversion TGFβ signaling is negatively regulated by inhibitory I-Smad7, which both interferes with phosphorylation of RSmad2/3 by the TGFβ receptor and facilitates degradation of the type I TGFβ receptor subunit through the recruitment of Smurf2 (242). The proto-oncoprotein Ski has also been shown to negatively regulate cardiac fibroblast phenotype conversion

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

(115). Ski inhibits Smad signaling by binding to DNA-bound R-Smad2 and repressing the transcription of Smad-activated genes (488). Though research to date has specifically focused on R-Smad2, it has been hypothesized that Ski negatively regulates both R-Smads (Fig. 2) (115). Through interference with Smad-mediated gene regulation, Ski downregulates the production of Zeb2 (Smad interacting protein-1) (114). Zeb2 acts as a negative regulator of mesenchyme homeobox 2 (Meox2), which has been shown to inhibit conversion to the myofibroblast phenotype by downregulating αSMA and ED-A fibronectin expression (90, 114). The attenuation of Smad signaling and Zeb2 expression by Ski is alleviated following its translocation to the cytoplasm, resulting in induction of the profibrotic gene expression program (114, 115). It has been reported that Ski is upregulated within the cardiac infarct scar, but is sequestered in the cytoplasm, leaving Zeb2 to repress Meox2 and resulting in upregulation of αSMA, ED-A fibronectin, and the myofibroblast marker smooth muscle embryonic myosin heavy chain (114, 115). Atrial natriuretic peptide (ANP)/cGMP/protein kinase G (PKG) signaling may serve a cardio-protective role by interfering with TGFβ signaling (286). ANP induces upregulation of cGMP, which in turn activates PKG (404). Activated PKG phosphorylates Smad3, but this occurs at different sites than those phosphorylated by the TGFβ receptor kinase, thus inhibiting the translocation of the Smad-complex to the nucleus and subsequently inhibiting collagen synthesis, fibroblast proliferation, and phenotype conversion (286). Elevated BNP and nitric oxide levels similarly attenuate TGFβ expression through the activation of cGMP (281, 417). Another factor regulating fibroblast phenotype is bFGF, which besides being a potent mitogen has been shown to prevent cardiac fibroblast phenotype conversion and to act as an antifibrotic factor. In cardiac fibroblasts from hypertensive rat hearts, bFGF downregulated αSMA and collagen types I and III expression (254). It also downregulated expression of MMP2, MMP9, and TIMP2 while upregulating TIMP1. This effectively attenuated both remodeling of the ECM and collagen production.

The Cardiac Myofibroblast Myofibroblasts can arise from a number of different sources. Besides conversion from resident fibroblasts within the myocardial interstitium, myofibroblasts may also arise from differentiation or phenotype conversion from other cells such as circulating myeloid or bone marrow-derived cells, pericytes, fibrocytes, and epithelial cells from the epicardium (1, 335, 402, 582). A recent study of kidney fibrosis found that approximately 50% of myofibroblasts arose from resident fibroblasts, with 35% from differentiation of bone marrow-derived cells and the remainder from epithelial- or endothelial-to-mesenchymal transition (283). It is unclear, however, whether myofibroblasts in the fibrotic or healing heart arise from similar sources, or in similar ratios.

743

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Comprehensive Physiology

Stress fibers SFs are the major contractile structure in numerous nonmuscle cell types, with the same basic macromolecular architecture as the contractile apparatus in smooth muscle cells. SFs are comprised of large bundles of actin filaments that often span a cell’s diameter, and are anchored at one or both ends by FAs (178). They contain actin-myosin-based contractile proteins such as tropomyosin and filamin. Cross-linking between SFs is achieved via periodically distributed α-actinin, alternating with myosin II. The presence of myosin filaments is responsible for the contractility of SFs observed in fibroblastic cells of many tissue sources, and can be used to regulate isometric tension within the cell (60, 240, 268, 385). Studies have demonstrated that central SFs are mostly regulated by the RhoA pathway, whereas peripheral SFs appear to be regulated primarily by MLCK activity (237, 238, 240). Two types of contraction are observed in cultured cardiac fibroblasts: that which is calcium-dependent, and regulated via a calmodulin/MLCK signaling system; and calcium-independent contraction, regulated by ROCK2 activation (8, 9). Generally, calcium-dependent contraction tends to be short-lived and rapid, whereas RhoA-mediated contraction is sustained by flexible control of myosin phosphatase activity (through ROCK; Fig. 6) (513). Cardiac myofibroblasts are characterized by the formation of SFs that incorporate αSMA. Migrating fibroblasts and those cultured on soft substrates do not develop SFs, and they are rarely observed in situ, although they have been identified in wounded dermis (60, 173, 203, 384, 508). On immobilized collagen gels, isometric tension induces the development of SFs in fibroblasts. However, when gels are released, contraction occurs and SFs are disassembled (183, 191, 333, 509). Three different manifestations of SFs have been identified in myofibroblasts, that is, ventral SFs, dorsal SFs, and transverse arcs. Most common usage of the term “SF” refers to ventral SFs, which are the most commonly observed and which are anchored at both ends by FAs, often spanning the length of the cell. Dorsal, or radial, SFs are anchored at only one end by a focal complex or adhesion, and are likely precursors to ventral SFs. In migrating cells, dorsal SFs form at the leading edge. In contrast to these two types of SFs, transverse arcs are not anchored by FAs and are generally much smaller and convex, developing at the base of lamellipodia in spreading and migrating NIH-3T3 fibroblasts (200,473,541). Although they are not considered SFs by traditional definitions, transverse arcs may also give rise to ventral SFs (214). SFs anchored by one or both ends by FAs are composed of F-actin cross-linked by a periodic pattern alternating between α-actinin and myosin II (61). Both RhoA and ROCK have been found to bind SFs, and Rho-associated GTPases control the formation and disassembly of SFs (331, 408, 409). In line with the development of SFs, culture of fibroblasts on rigid substrates induces RhoA activation, in turn activating downstream effectors such as ROCK/ROCK2 (379, 562). Activation of ROCK2 mediates

744

Mechanical strain

Calcium ion influx

Ca2+

Integrin

Plasma membrane

TRPC

Cytoplasm

Ca2+

RhoA Ca2+

Calmodulin

ROCK MP MLCK

P

F-actin

MLC

Slow, sustained contraction

MLC

MLC

Rapid, transient contraction

Figure 6

Myofibroblast contractile response. Contraction of stress fibers occurs via calcium-dependent and -independent signaling pathways in cardiac fibroblasts. Calcium-dependent contraction is typically rapid and occurs primarily in peripheral stress fibers. Through receptors such as transient receptor potential cation (TRPC) channels, calciumdependent contraction occurs in response to an influx of Ca2+ ions, which form a complex with and activate calmodulin, resulting in myosin light chain kinase (MLCK) activation and subsequent phosphorylation of myosin light chain (MLC) bound to stress fibers. However, contraction is not sustained due to the activity of myosin phosphatase (MP) within the cell, which dephosphorylates and inactivates actin-bound MLC. Conversely, calcium-independent contraction is relatively slow and acts primarily on central stress fibers in sustained contractile events, usually in response to mechanical strain. Calcium-independent contraction occurs primarily through integrin activation and downstream RhoARho-GTPase-associated kinase (ROCK) signaling, which inhibits MP to maintain phosphorylation of MLC, allowing for prolonged contractile events.

the organization of SFs and FAs and its inhibition with Y-27632 causes disassembly of centrally located ventral SFs (7-9, 237, 241, 409). The RhoA/ROCK2 complex is localized on SFs via myosin phosphatase-RhoA interacting protein/ M-RIP (407). ROCK2 then phosphorylates MLC in a calcium-independent manner either directly or indirectly. ROCK2-dependent phosphorylation of MLC is essential for

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

fibroblast ventral SF formation, and occurs by T697 phosphorylation of the 130 kDa myosin-binding subunit of myosin phosphatase (159, 227, 238, 245, 253, 344, 406, 512, 556). Thus, myosin phosphatase activity upon MLC is inhibited, which also is necessary for sustained contraction. Additionally, RhoA and ROCK activation induces LIM kinase activation, which inhibits cofilin phosphorylation, rendering it inactive (306). Because cofilin mediates F-actin severing, ROCK activation of LIM kinase also contributes to increased F-actin content and RhoA-induced ventral SF assembly (14, 25, 63, 275, 500). Cofilin severing of actin filaments (and subsequent SF disassembly) is also inhibited by cellular tension—less cofilin is bound to actin filaments with tension than in fibroblasts “at rest” (198). In dorsal SFs, RhoA activation of the effector molecule mDia1 drives actin polymerization (214, 361). Nonetheless, both arms of the RhoA pathway (i.e., ROCK and mDia1) are necessary for normal SF patterning in myofibroblasts. Watanabe et al. determined that in the absence of RhoA and ROCK, mDia1-mediated SF formation resulted in sheet-like filaments with no discernible “bundles” of F-actin. In the absence of both RhoA and mDia1, ROCK-induced SFs were increased in size and appeared in a star conformation rather than longitudinally along the cell (544). The ability of SFs to cause cardiac myofibroblast contraction lies in their architecture. Rather than being composed of unidirectional F-actin filaments, SFs have unipolar filaments that are oppositely oriented at each myosin bridge. That is, a unipolar filament is formed with actin polymerization and SF formation, but requires cleavage and capping, followed by actin nucleation at the other side of the “cap” to generate filaments polarized in the opposite direction, which are eventually joined by myosin II (63). Though the proteins involved in these cleavage and capping functions have not been identified in cardiac fibroblasts, it is possible that the actin-severing protein formin may fulfill cleavage, capping, and elongation functions (195). Recruitment of myosin II is likely dependent on tropomyosin, an actin-binding protein recruited to various actin filament populations (186, 187). In the absence of tropomyosin, SFs spontaneously disassemble, possibly through elevation of cAMP, which has also been shown to cause SF disassembly (23, 55, 190, 273). cAMP elevation results in phosphorylation of RhoA [dependent upon protein kinase A (PKA) activity] which causes RhoA sequestration by enhancing its binding to Rho guanine nucleotide dissociation inhibitor (153,274,429). Additionally, cAMP elevation induces PKA phosphorylation of MLC kinase, in turn inhibiting its kinase activity in phosphorylating MLC (273).

Contraction Unlike contraction in muscle cells, which can be nearly instantaneous and involves actin-myosin cross-bridge cycling, contraction of cardiac myofibroblasts is slower, often more sustained, and depends upon the phosphorylation state of the

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

light chain of myosin II bound to SFs. In myofibroblasts with a myosin II mutation resulting in disrupted motor activity, actin binding capability remained although the cells were unable to generate SF-induced contractile force (99). SFs isolated from fibroblasts with Triton X-100 detergent extraction contract in response to Mg2+ -ATP treatment in the presence of calcium ions, a process caused by phosphorylation of MLC via calmodulin and MLCK signaling, the same mechanism by which smooth muscle contraction occurs (239, 240). Treatment of fibroblasts with RhoA activators (e.g., lysophosphatidic acid and bombesin) induces SF reorganization and contraction (408, 409). ROCK activation is calcium independent and relies on upstream RhoA activation (238, 513). ROCK interacts with SFs by binding directly to myosin II (243). ROCK activation induces SF contraction via phosphorylation of MLC, both directly by phosphorylation (S854/T697) and indirectly through inhibition of myosin phosphatase (Fig. 6) (8, 9, 238, 513).

Cardiac myofibroblasts and migration A growing body of evidence produced by studying fibroblast to myofibroblast phenotype conversion indicates that myofibroblasts are less motile than their precursors (425). This is likely attributable to the increased size and strength of FAs in myofibroblasts. Decreased migration of cardiac myofibroblasts logically follows increased adhesion, synthesis of matrix components, reduced MMP activity, and increased contractility. Although the reduction in motility observed in myofibroblasts can likely be attributed primarily to increased adhesions and synthesis, there is the potential for the involvement of distinct signaling pathways, though these remain undetermined. Wnt-Frizzled signaling is implicated in phenotype conversion, and reduces migration in telomerase-immortalized cardiac fibroblasts (271). However, dishevelled-1, a signaling component of the Wnt-Frizzled pathway, appears to be associated with migration of cardiac fibroblasts during post-MI wound healing, appearing in the border zone of the infarct 1 day post-MI and in both the border zone and infarct scar after 4 days (86). Additionally, inhibition of MMPs results in increased density of the matrix and may also downregulate migration. For example, TIMP-4 is capable of inhibiting myofibroblast motility (517). Conversely, some studies report increased migration by myofibroblasts compared to fibroblasts. On gelatin-coated substrates, TGFβ induced migration of myofibroblasts, and Smad3-null MEFs have shown a decrease in migratory ability (137, 592). Another study found that TGFβ1 induced MMP activity in cardiac fibroblasts, facilitating their motility, through amplification of the activity of the convertase furin, which acts to activate TGFβ1 , and downstream MMP4 activation (480). Thus, although the majority of the evidence is supportive of decreased myofibroblast motility, controversy remains regarding not only the migratory capacity

745

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

of these cells, but also the definition of “fibroblast” and “myofibroblast” in vitro. Two issues should be considered in resolving this paradox: first, the heterogeneity of cells in culture derived from primary tissues introduces the possibility that, at the time of assay, some cells are myofibroblasts while others, presumably with higher motility, retain their fibroblast character, and it is these latter cells that actually migrate (e.g., in scratch assays). The second possibility is that fibroblasts migrate initially, followed by conversion to the myofibroblast phenotype. Careful experimental design, such as the performance of mobility assays followed immediately by fixation and immunostaining for fibroblast- and myofibroblast-enriched markers, may help to address this issue.

Increased matrix production by myofibroblasts Myofibroblasts are characterized by hyper-secretion of ECM components such as type I collagen, periostin, fibronectin, and ED-A fibronectin (448, 508). The healthy myocardium is largely devoid of myofibroblasts; however, these cells become abundant within a few days after myocardial injury such as MI (486). Myofibroblasts are also defined by the de novo expression of smooth muscle cell genes including caldesmon, SM22, and αSMA, the latter being incorporated into SFs and increasing cell contractility as noted above (131, 209, 277, 400, 426). Excessive production and remodeling of ECM components by myofibroblasts directly leads to cardiac fibrosis. As mentioned previously, TGFβ plays a prime role in the phenotypic conversion of fibroblasts to myofibroblasts (206, 280, 534). TGFβ stimulation of either cell type induces ECM gene expression and deposition of ECM components, simultaneously with alterations in expression of MMPs and TIMPs. This paradigm is not limited to the heart but extends to most tissues in the body (122, 390, 534, 584). The canonical TGFβ-Smad2/3 signaling pathway directly transactivates ECM genes including Col1α1 and Col1α2, as well as MMPs and TIMPS (22, 156, 527). Hearts from Smad3 null mice exhibit attenuated collagen production and deposition, and reduced fibrotic remodeling following MI (137). Additionally, the transcription factor scleraxis has been shown to tightly coordinate the regulation of type I collagen expression in synergy with R-Smad3 in primary isolated cardiac fibroblasts and myofibroblasts (22, 156). The noncanonical signaling pathway initiated by TGFβ binding to TGFβ receptor subunit II induces recruitment of TAK1 and TAK1 binding protein/TAB followed by activation of MAPK signaling cascades including JNK and p38 MAPK (123). Targeted inhibition of noncanonical TGFβ signaling using TGFβ receptor subunit II KO mice resulted in reduced fibrosis and remodeling in the pressureoverloaded myocardium (262). Inhibition of p38 signaling can also block TGFβ-induced myofibroblast traits including expression of type I collagen, fibronectin, αSMA-positive SF formation, and collagen gel contraction (122). Collectively, these studies identify TGFβ signaling pathways as critical drivers of myofibroblast-mediated fibrosis.

746

Comprehensive Physiology

AngII is elevated in fibrotic pressure overloaded hearts and appears to positively affect myofibroblast activity and fibrosis including increased ECM production (197, 212). AngII treatment of cardiac fibroblasts leads to induction of TGFβ expression through the AT1 receptor causing upregulation of downstream fibrotic genes including type I collagen, an effect that could be blocked by the administration of losartan (74, 284). Activation of the Smad2/3 and MAPK signaling pathways occurs in response to AngII treatment in cardiac myofibroblasts (175,570). AngII thus appears to be a crucial regulator of the myofibroblast phenotype through multiple mechanisms. As described above, significant crosstalk also appears to exist between TGFβ and AngII signaling pathways, resulting in reinforcing and synergistic effects on fibroblast phenotype and function. ET1 production during cardiac injury induces myofibroblast conversion from primary cardiac fibroblasts in vitro (356). ET1 also strongly induced ECM production in fibroblasts (279). AngII is capable of increasing ET1 expression in cardiac fibroblasts, thus driving them to a myofibroblast phenotype (93). This effect can be blocked using losartan. TGFβ also works together with ET1 in driving the myofibroblast phenotype (457). Thus it appears that ET1 is a downstream synergistic mediator of TGFβ and AngII signaling, promoting fibroblast activation, maintenance of the myofibroblast phenotype and consequent tissue fibrosis (278, 280).

Senescence and cell death In dermal wounding, myofibroblasts arise as part of the healing process, but are lost through apoptosis once healing is completed (116). In contrast, cardiac myofibroblasts may persist for years, thus the discovery of an “off-switch” to remove myofibroblasts without adversely affecting proper maintenance of the ECM and the structural integrity of the heart would represent a therapeutic boon. To this end, understanding the factors governing myofibroblast senescence and death is of key concern. Cellular senescence refers to permanent cell cycle arrest, which occurs in most higher eukaryotic cells as they reach the end of their life cycle (75, 328). When a cell reaches senescence naturally at the end of its life cycle, it loses the ability to respond to mitogenic factors (28). Replicative senescence appears to serve as an excellent model for age-related senescence, since cultured cardiac fibroblasts at higher passages show similar signs of cellular senescence as primary fibroblasts taken from the hearts of older animals (134, 540). Senescence of cardiac fibroblasts remains a poorly understood area. Some groups have suggested that senescence is a way of limiting the fibrotic response, while others contend that senescent fibroblasts add to the problem, reporting that fibroblasts become larger, cease growth, and secrete more collagen (496, 540, 590). PDGF-AA is a strong mitogenic factor for cardiac fibroblasts (463). In older fibroblasts, PDGF-AA is less efficient at stimulating a mitogenic response and also showed attenuated

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

expression of Akt, a required component in the signaling cascade for PDGF to stimulate proliferation; conversely, Akt overexpression in older fibroblasts partially restored responsiveness to PDGF (134). In other cell lines it has been shown that PIP3 recruits the serine/threonine kinase Akt to the plasma membrane, where it can be phosphorylated by PDK1 (5). Phospho-Akt in turn phosphorylates GSK3, thereby deactivating it and preventing GSK3-mediated phosphorylation of cyclinD1 which would result in cyclinD1 translocation into the cytoplasm, making it unavailable to participate in cell cycle signaling cascades. The phosphorylation of cyclinD1 thus ultimately results in cell cycle arrest (133). This pathway may be one of many involved in aging-related cellular senescence due to attenuated Akt expression. As fibroblasts age, F-actin fibers become disorganized and there is an upregulation of 4-hydroxynonenal, an aging marker, and β-catenin, suggestive of dysregulation in the Wnt/β-Catenin pathway (540, 557). This is accompanied by downregulation of p70S6K, Cdc42, Mdm2, LOX-1, and MSR1 (540). LOX-1, the receptor for oxidized low-density lipoprotein, is partially responsible for the downregulation of these other proteins (87, 540). The overexpression of LOX-1 in older fibroblasts increased the expression of Mdm2, pAkt, Cdc42 and p70S6K, although protein levels were still lower than normal basal levels (540). Cdc42 promotes cell cycle progression into the S phase (366). Changes in the expression of microRNAs may also play a role in age-related cellular senescence in cardiac fibroblasts. As the heart ages, miR-22 is upregulated in cardiac fibroblasts (229). miR-22 negatively regulates expression of osteoglycin, which leads to the induction of senescence through the upregulation of p16. miR-21 is upregulated in cardiac fibroblasts during heart failure and induces hypertrophy and tissue fibrosis through the induction of bFGF/FGF2 secretion and inhibition of apoptosis (504, 505). Bcl-2, an antiapoptotic factor, is positively regulated by miR-21 (138, 320). miR-21 has also been shown to induce activation of the ERK/MAPK pathway, thereby enhancing proliferation of cardiac fibroblasts (505). This poses a possible mechanism for apoptosis resistance in cardiac myofibroblasts and the persistence of the profibrotic response following myocardial injury. Although many targets for miR21 have been proposed, the specific protein(s) it targets to impact FGF2 secretion, apoptosis and proliferation remains to be elucidated (138). Though cellular senescence occurs naturally in cardiac fibroblasts, it also occurs in response to myocardial injury. The tumor suppressor p53 regulates many genes including those involved in apoptosis, senescence and cell-cycle arrest (532). p53 activity increases following myocardial injury, resulting in limitation of collagen deposition and tissue fibrosis (532, 590). In cardiac fibroblasts, it has been proposed that p53 directly decreases Col1α1 expression, inhibits proliferation through the activation of p21 and attenuates TGFβinduced CTGF expression through unknown mechanisms (496). Fibroblast Specific Protein-1 (FSP1) regulates collagen

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

production and proliferation via the inhibition of p53 (496). FSP1 undergoes a conformational change when it binds to calcium and becomes activated. It dimerizes and modulates the binding of p53 to DNA by directly binding the protein (31, 182, 423). FSP1 is upregulated in cardiac fibroblasts of hypertensive rats, and FSP1 knockout has been shown to significantly reduce the fibrotic response following myocardial injury (496). This mechanism may help explain resistance to cellular senescence in cardiac fibroblasts that results in maladaptive healing after injury. During ischemic conditions within the heart, ROS are widely produced concomitant with elevated intracellular calcium levels. Reperfusion following ischemia results in protein degradation and ultimately ischemia/reperfusion injury, leading to cell death via necrosis or apoptosis (56, 98, 397, 467). pERK1/2 levels decrease during ischemia but are replenished following reperfusion whereas pAkt levels remain at basal levels during ischemia and are upregulated during reperfusion (531). It has been proposed that insulin-like growth factor-1 (IGF-1) is produced by cardiac fibroblasts and could play a cardio-protective role following myocardial injury (492). IGF-1 resists pERK1/2 downregulation during ischemia, further upregulates pAkt during ischemia and reperfusion and also partially prevents cell death during reperfusion. Apoptosis induction is inhibited through the activation of both the MEK-ERK and PI3K/Akt pathways simultaneously (531). Apoptosis may also be induced by severe hypoxia (101). Hypoxia inducible factor-1α (HIF1α) protects against excessive apoptosis in response to hypoxia. HIF1α is the subunit of heterodimeric HIF1 that regulates gene expression and helps restore oxygen homeostasis (536, 552). The proapoptotic effects of hypoxia include the induction of Bax expression, increased intracellular calcium levels, caspase-3 activation, and downregulation of antiapoptotic Bcl-2 (569). Simultaneously, hypoxic conditions result in the upregulation of HIF1α within the nucleus, which not only inhibits the proapoptotic effects of hypoxia but also promotes increased fibroblast viability (569). Thus, although proapoptotic and antiapoptotic signals are being transduced simultaneously, it is the net effect of these two signals that will ultimately determine the cell fate.

Summary and Conclusions Cardiac fibroblasts and myofibroblasts are central players in the health and diseases of the heart, yet remain poorly understood in comparison to their myocyte neighbors. Concerted efforts over the past decade, by an ever-increasing tally of research laboratories, have finally begun to reveal the innermost workings of these intriguing cells (Table 3). This work has been driven forward not only by intellectual curiosity, but also by the burgeoning need to develop novel therapeutics against unexploited targets in the treatment of cardiovascular diseases. Attempts to therapeutically target pleiotrophic

747

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

Table 3

Comprehensive Physiology

Major Effectors and Signaling Pathways of Key Cardiac Fibroblast/Myofibroblast Biological Functions

Biological function

Cellular process

Major regulatory effectors and pathways

Extracellular matrix synthesis and degradation

Extracellular matrix production

AngII (RAAS), bFGF, CTGF, PDGF, TGFβ

Matrix remodeling (MMPs, TIMPs)

AngII, BNP, IL-1β, PDGF, TGFβ, TNFα

Phenotype conversion of fibroblasts to myofibroblasts

Cell-matrix adhesions

Migration

Myofibroblast hypersynthesis of ECM

AngII, ET1, TGFβ

Positive regulators

AngII, ET1, miR-145, MRTF-A, Rho-ROCK, TGFβ,TRPC3, TRPC6, TRPM7

Negative regulators

ANP/BNP, bFGF, I-Smad7, Meox2, NO, Ski, Smurf2

Adhesion formation

FAK, mDia1, Rac, RhoA

Adhesion maturation

Ras-Raf-MAPK-ERK, RhoA, Syn 4

Mechanosensing/mechanotransduction

FAK, integrins, p130Cas, RhoA-ROCK, Yap/Taz

Cell polarization

FAK, Rac1, RhoA

Adhesion turnover

Rac1

Retraction of trailing edge

Rac, vinculin

Cardiac injury response

Reparative processes

AngII, Caspase-1, CT-1, FGF, IL-1β, IL-10, IL-18, IL-23, MMP3, MMP8, TIMP1, TGFβ

Cell cycle and proliferation

Proliferation

Adenylyl cyclase, AngII, CDK2, CDK4, Cyclin A, Cyclin B, Cyclin D, Cyclin E, ERK1/2, FGF, FSP1, HGF, miR-21, PDGF-AA, PDGF-BB, PI3K, Skp2

Senescence and cell death

4-Hydroxynonenal, β-catenin, HIF1α, LOX-1, miR-22, p53, PDGF-AA, PIP3, ROS

Stress fiber formation

Rho-associated GTPases, ROCK

Contraction

MLCK, RhoA-ROCK

Contraction

regulators of fibrosis, including growth factors like TGFβ and AngII or general signaling pathways like ERK, are unlikely to be fruitful given the likelihood of off-target effects such as has been observed with oncogenesis and TGFβ inhibition. Instead, a greater focus on targeting one or a combination of factors that appear to be much more specific to the processes of ECM formation and fibrosis may improve specificity of treatment, such as the potential for targeting the function of scleraxis, or for augmenting or blocking the function of specific miRNAs to attenuate expression of downstream target genes like collagens. Approximately one-third of individuals in the developed world will die as a result of cardiovascular diseases, and cardiac fibrosis will be a contributory factor for many of these patients given its association with such common pathologies as MI, hypertension, diabetes, valve disease, cardiomyopathies, and simple aging, particularly as our population collectively gets older. The development of new noninvasive technologies such as late gadolinium enhancement of magnetic resonance imaging for identifying and measuring fibrosis in cardiac patients promises to provide a much clearer picture of exactly how big the clinical problem is. Precise dissection of the intracellular signaling pathways governing fibroblast/myofibroblast proliferation, migration, and ECM synthesis offers the promise of identifying the means to

748

specifically attenuate one or more of these processes, leading to the development of therapeutic relief just as it is most needed.

Acknowledgements This work was supported by a Manitoba Health Research Council Manitoba Partnership Program grant and an Open Operating Grant from the Canadian Institutes of Health Research to MPC (MOP 106671). K. L. Filomeno was supported through a grant from the Heart & Stroke Foundation of Canada. P. L. Roche was the recipient of a graduate studentship from MHRC. R. A. Bagchi was the recipient of a doctoral fellowship from CIHR/MHRC.

References 1. 2.

Abe R, Donnelly SC, Peng T, Bucala R, Metz CN. Peripheral blood fibrocytes: Differentiation pathway and migration to wound sites. J Immunol 166: 7556-7562, 2001. Acharya A, Baek ST, Huang G, Eskiocak B, Goetsch S, Sung CY, Banfi S, Sauer MF, Olsen GS, Duffield JS, Olson EN, Tallquist MD. The bHLH transcription factor Tcf21 is required for lineage-specific EMT of cardiac fibroblast progenitors. Development 139: 2139-2149, 2012.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

3.

4. 5.

6.

7. 8.

9.

10.

11. 12.

13.

14. 15. 16.

17. 18. 19.

20. 21.

22. 23.

24.

25. 26.

Adapala RK, Thoppil RJ, Luther DJ, Paruchuri S, Meszaros JG, Chilian WM, Thodeti CK. TRPV4 channels mediate cardiac fibroblast differentiation by integrating mechanical and soluble signals. J Mol Cell Cardiol 54: 45-52, 2013. Agapitov AV, Haynes WG. Role of endothelin in cardiovascular disease. J Renin Angiotensin Aldosterone Syst 3: 1-15, 2002. Alessi DR, James SR, Downes CP, Holmes AB, Gaffney PR, Reese CB, Cohen P. Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Balpha. Curr Biol 7: 261-269, 1997. Alexandrova AY, Arnold K, Schaub S, Vasiliev JM, Meister JJ, Bershadsky AD, Verkhovsky AB. Comparative dynamics of retrograde actin flow and focal adhesions: Formation of nascent adhesions triggers transition from fast to slow flow. PloS One 3: e3234, 2008. Amano M, Chihara K, Kimura K, Fukata Y, Nakamura N, Matsuura Y, Kaibuchi K. Formation of actin stress fibers and focal adhesions enhanced by Rho-kinase. Science 275: 1308-1311, 1997. Amano M, Ito M, Kimura K, Fukata Y, Chihara K, Nakano T, Matsuura Y, Kaibuchi K. Phosphorylation and activation of myosin by Rho-associated kinase (Rho-kinase). J Biol Chem 271: 20246-20249, 1996. Amano M, Mukai H, Ono Y, Chihara K, Matsui T, Hamajima Y, Okawa K, Iwamatsu A, Kaibuchi K. Identification of a putative target for Rho as the serine-threonine kinase protein kinase N. Science 271: 648-650, 1996. Anderson TW, Vaughan AN, Cramer LP. Retrograde flow and myosin II activity within the leading cell edge deliver F-actin to the lamella to seed the formation of graded polarity actomyosin II filament bundles in migrating fibroblasts. Mol Biol Cell 19: 5006-5018, 2008. Annes JP, Munger JS, Rifkin DB. Making sense of latent TGFbeta activation. J Cell Sci 116: 217-224, 2003. Anthis NJ, Haling JR, Oxley CL, Memo M, Wegener KL, Lim CJ, Ginsberg MH, Campbell ID. Beta integrin tyrosine phosphorylation is a conserved mechanism for regulating talin-induced integrin activation. J Biol Chem 284: 36700-36710, 2009. Anthis NJ, Wegener KL, Ye F, Kim C, Goult BT, Lowe ED, Vakonakis I, Bate N, Critchley DR, Ginsberg MH, Campbell ID. The structure of an integrin/talin complex reveals the basis of inside-out signal transduction. EMBO J 28: 3623-3632, 2009. Arber S, Barbayannis FA, Hanser H, Schneider C, Stanyon CA, Bernard O, Caroni P. Regulation of actin dynamics through phosphorylation of cofilin by LIM-kinase. Nature 393: 805-809, 1998. Arnaout MA, Goodman SL, Xiong JP. Structure and mechanics of integrin-based cell adhesion. Curr Opin Cell Biol 19: 495-507, 2007. Arslan F, Smeets MB, Riem Vis PW, Karper JC, Quax PH, Bongartz LG, Peters JH, Hoefer IE, Doevendans PA, Pasterkamp G, de Kleijn DP. Lack of fibronectin-EDA promotes survival and prevents adverse remodeling and heart function deterioration after myocardial infarction. Circ Res 108: 582-592, 2011. Arthur WT, Burridge K. RhoA inactivation by p190RhoGAP regulates cell spreading and migration by promoting membrane protrusion and polarity. Mol Biol Cell 12: 2711-2720, 2001. Arthur WT, Petch LA, Burridge K. Integrin engagement suppresses RhoA activity via a c-Src-dependent mechanism. Curr Biol 10: 719722, 2000. Asparuhova MB, Ferralli J, Chiquet M, Chiquet-Ehrismann R. The transcriptional regulator megakaryoblastic leukemia-1 mediates serum response factor-independent activation of tenascin-C transcription by mechanical stress. FASEB J 25: 3477-3488, 2011. Asparuhova MB, Gelman L, Chiquet M. Role of the actin cytoskeleton in tuning cellular responses to external mechanical stress. Scand J Med Sci Sports 19: 490-499, 2009. Assoian RK, Fleurdelys BE, Stevenson HC, Miller PJ, Madtes DK, Raines EW, Ross R, Sporn MB. Expression and secretion of type beta transforming growth factor by activated human macrophages. Proc Natl Acad Sci U S A 84: 6020-6024, 1987. Bagchi RA, Czubryt MP. Synergistic roles of scleraxis and Smads in the regulation of collagen 1alpha2 gene expression. Biochim Biophys Acta 1823: 1936-1944, 2012. Bakin AV, Safina A, Rinehart C, Daroqui C, Darbary H, Helfman DM. A critical role of tropomyosins in TGF-beta regulation of the actin cytoskeleton and cell motility in epithelial cells. Mol Biol Cell 15: 4682-4694, 2004. Bakolitsa C, Cohen DM, Bankston LA, Bobkov AA, Cadwell GW, Jennings L, Critchley DR, Craig SW, Liddington RC. Structural basis for vinculin activation at sites of cell adhesion. Nature 430: 583-586, 2004. Bamburg JR, McGough A, Ono S. Putting a new twist on actin: ADF/cofilins modulate actin dynamics. Trends Cell Biol 9: 364-370, 1999. Barbolina MV, Stack MS. Membrane type 1-matrix metalloproteinase: Substrate diversity in pericellular proteolysis. Sem Cell Dev Biol 19: 24-33, 2008.

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

27. Bashey RI, Martinez-Hernandez A, Jimenez SA. Isolation, characterization, and localization of cardiac collagen type VI. Associations with other extracellular matrix components. Circ Res 70: 1006-1017, 1992. 28. Bauer EA, Silverman N, Busiek DF, Kronberger A, Deuel TF. Diminished response of Werner’s syndrome fibroblasts to growth factors PDGF and FGF. Science 234: 1240-1243, 1986. 29. Bauersachs J, Thum T. Biogenesis and regulation of cardiovascular microRNAs. Circ Res 109: 334-347, 2011. 30. Bellin RM, Kubicek JD, Frigault MJ, Kamien AJ, Steward RL, Jr., Barnes HM, Digiacomo MB, Duncan LJ, Edgerly CK, Morse EM, Park CY, Fredberg JJ, Cheng CM, LeDuc PR. Defining the role of syndecan4 in mechanotransduction using surface-modification approaches. Proc Natl Acad Sci U S A 106: 22102-22107, 2009. 31. Berge G, Maelandsmo GM. Evaluation of potential interactions between the metastasis-associated protein S100A4 and the tumor suppressor protein p53. Amino Acids 41: 863-873, 2011. 32. Bergman MR, Cheng S, Honbo N, Piacentini L, Karliner JS, Lovett DH. A functional activating protein 1 (AP-1) site regulates matrix metalloproteinase 2 (MMP-2) transcription by cardiac cells through interactions with JunB-Fra1 and JunB-FosB heterodimers. Biochem J 369: 485-496, 2003. 33. Bernanke DH, Markwald RR. Effects of hyaluronic acid on cardiac cushion tissue cells in collagen matrix cultures. Tex Rep Biol Med 39: 271-285, 1979. 34. Berry MF, Engler AJ, Woo YJ, Pirolli TJ, Bish LT, Jayasankar V, Morine KJ, Gardner TJ, Discher DE, Sweeney HL. Mesenchymal stem cell injection after myocardial infarction improves myocardial compliance. Am J Physiol Heart Circ Physiol 290: H2196-2203, 2006. 35. Bhana B, Iyer RK, Chen WL, Zhao R, Sider KL, Likhitpanichkul M, Simmons CA, Radisic M. Influence of substrate stiffness on the phenotype of heart cells. Biotechnol Bioeng 105: 1148-1160, 2010. 36. Bing OH, Ngo HQ, Humphries DE, Robinson KG, Lucey EC, Carver W, Brooks WW, Conrad CH, Hayes JA, Goldstein RH. Localization of alpha1(I) collagen mRNA in myocardium from the spontaneously hypertensive rat during the transition from compensated hypertrophy to failure. J Mol Cell Cardiol 29: 2335-2344, 1997. 37. Bishop JE, Laurent GJ. Collagen turnover and its regulation in the normal and hypertrophying heart. Eur Heart J 16(Suppl C): 38-44, 1995. 38. Blomer N, Pachel C, Hofmann U, Nordbeck P, Bauer W, Mathes D, Frey A, Bayer B, Vogel B, Ertl G, Bauersachs J, Frantz S. 5Lipoxygenase facilitates healing after myocardial infarction. Basic Res Cardiol 108: 367, 2013. 39. Blumer KJ, Johnson GL. Diversity in function and regulation of MAP kinase pathways. Trends Biochem Sci 19: 236-240, 1994. 40. Boettner B, Van Aelst L. Control of cell adhesion dynamics by Rap1 signaling. Curr Opin Cell Biol 21: 684-693, 2009. 41. Booz GW, Baker KM. Molecular signalling mechanisms controlling growth and function of cardiac fibroblasts. Cardiovasc Res 30: 537543, 1995. 42. Borg TK, Rubin K, Lundgren E, Borg K, Obrink B. Recognition of extracellular matrix components by neonatal and adult cardiac myocytes. Dev Biol 104: 86-96, 1984. 43. Bos JL. Linking Rap to cell adhesion. Curr Opin Cell Biol 17: 123-128, 2005. 44. Bos JL, Rehmann H, Wittinghofer A. GEFs and GAPs: Critical elements in the control of small G proteins. Cell 129: 865-877, 2007. 45. Bosman FT, Stamenkovic I. Functional structure and composition of the extracellular matrix. The Journal of pathology 200: 423-428, 2003. 46. Bouaouina M, Lad Y, Calderwood DA. The N-terminal domains of talin cooperate with the phosphotyrosine binding-like domain to activate beta1 and beta3 integrins. J Biol Chem 283: 6118-6125, 2008. 47. Boutahar N, Guignandon A, Vico L, Lafage-Proust MH. Mechanical strain on osteoblasts activates autophosphorylation of focal adhesion kinase and proline-rich tyrosine kinase 2 tyrosine sites involved in ERK activation. J Biol Chem 279: 30588-30599, 2004. 48. Brilla CG, Weber KT. Reactive and reparative myocardial fibrosis in arterial hypertension in the rat. Cardiovasc Res 26: 671-677, 1992. 49. Brilla CG, Zhou G, Matsubara L, Weber KT. Collagen metabolism in cultured adult rat cardiac fibroblasts: Response to angiotensin II and aldosterone. J Mol Cell Cardiol 26: 809-820, 1994. 50. Brodersen P, Voinnet O. Revisiting the principles of microRNA target recognition and mode of action. Nature Rev Mol Cell Biol 10: 141-148, 2009. 51. Brooks WW, Healey NA, Sen S, Conrad CH, Bing OH. Oxygen cost of stress development in hypertrophied and failing hearts from the spontaneously hypertensive rat. Hypertension 21: 56-64, 1993. 52. Brown MC, Perrotta JA, Turner CE. Identification of LIM3 as the principal determinant of paxillin focal adhesion localization and characterization of a novel motif on paxillin directing vinculin and focal adhesion kinase binding. J Cell Biol 135: 1109-1123, 1996.

749

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

53. 54. 55.

56. 57. 58. 59. 60. 61. 62. 63. 64.

65. 66. 67.

68. 69.

70. 71.

72. 73.

74. 75. 76. 77.

78.

79.

750

Brown MC, West KA, Turner CE. Paxillin-dependent paxillin kinase linker and p21-activated kinase localization to focal adhesions involves a multistep activation pathway. Mol Biol Cell 13: 1550-1565, 2002. Brown RD, Jones GM, Laird RE, Hudson P, Long CS. Cytokines regulate matrix metalloproteinases and migration in cardiac fibroblasts. Biochem Bioph Res Co 362: 200-205, 2007. Bryce NS, Schevzov G, Ferguson V, Percival JM, Lin JJ, Matsumura F, Bamburg JR, Jeffrey PL, Hardeman EC, Gunning P, Weinberger RP. Specification of actin filament function and molecular composition by tropomyosin isoforms. Mol Biol Cell 14: 1002-1016, 2003. Buja LM. Modulation of the myocardial response to ischemia. Lab Invest 78: 1345-1373, 1998. Bujak M, Frangogiannis NG. The role of TGF-beta signaling in myocardial infarction and cardiac remodeling. Cardiovasc Res 74: 184-195, 2007. Burgering BM, Coffer PJ. Protein kinase B (c-Akt) in phosphatidylinositol-3-OH kinase signal transduction. Nature 376: 599-602, 1995. Burgess ML, Terracio L, Hirozane T, Borg TK. Differential integrin expression by cardiac fibroblasts from hypertensive and exercisetrained rat hearts. Cardiovasc Pathol 11: 78-87, 2002. Burridge K. Are stress fibres contractile? Nature 294: 691-692, 1981. Burridge K, Chrzanowska-Wodnicka M, Zhong C. Focal adhesion assembly. Trends Cell Biol 7: 342-347, 1997. Burridge K, Wennerberg K. Rho and Rac take center stage. Cell 116: 167-179, 2004. Burridge K, Wittchen ES. The tension mounts: Stress fibers as forcegenerating mechanotransducers. J Cell Biol 200: 9-19, 2013. Burstein B, Libby E, Calderone A, Nattel S. Differential behaviors of atrial versus ventricular fibroblasts: A potential role for platelet-derived growth factor in atrial-ventricular remodeling differences. Circulation 117: 1630-1641, 2008. Bustos RI, Forget MA, Settleman JE, Hansen SH. Coordination of Rho and Rac GTPase function via p190B RhoGAP. Curr Biol 18: 16061611, 2008. Butt RP, Laurent GJ, Bishop JE. Collagen production and replication by cardiac fibroblasts is enhanced in response to diverse classes of growth factors. Eur J Cell Biol 68: 330-335, 1995. Cai Y, Biais N, Giannone G, Tanase M, Jiang G, Hofman JM, Wiggins CH, Silberzan P, Buguin A, Ladoux B, Sheetz MP. Nonmuscle myosin IIA-dependent force inhibits cell spreading and drives F-actin flow. Biophys J 91: 3907-3920, 2006. Calderwood DA. Talin controls integrin activation. Biochem Soc Tran 32: 434-437, 2004. Calderwood DA, Fujioka Y, de Pereda JM, Garcia-Alvarez B, Nakamoto T, Margolis B, McGlade CJ, Liddington RC, Ginsberg MH. Integrin beta cytoplasmic domain interactions with phosphotyrosinebinding domains: A structural prototype for diversity in integrin signaling. Proc Natl Acad Sci U S A 100: 2272-2277, 2003. Calderwood DA, Yan B, de Pereda JM, Alvarez BG, Fujioka Y, Liddington RC, Ginsberg MH. The phosphotyrosine binding-like domain of talin activates integrins. J Biol Chem 277: 21749-21758, 2002. Calderwood DA, Zent R, Grant R, Rees DJ, Hynes RO, Ginsberg MH. The Talin head domain binds to integrin beta subunit cytoplasmic tails and regulates integrin activation. J Biol Chem 274: 28071-28074, 1999. Camelliti P, Green CR, LeGrice I, Kohl P. Fibroblast network in rabbit sinoatrial node: Structural and functional identification of homogeneous and heterogeneous cell coupling. Circ Res 94: 828-835, 2004. Camenisch TD, Spicer AP, Brehm-Gibson T, Biesterfeldt J, Augustine ML, Calabro A, Jr., Kubalak S, Klewer SE, McDonald JA. Disruption of hyaluronan synthase-2 abrogates normal cardiac morphogenesis and hyaluronan-mediated transformation of epithelium to mesenchyme. J Clin Invest 106: 349-360, 2000. Campbell SE, Katwa LC. Angiotensin II stimulated expression of transforming growth factor-beta1 in cardiac fibroblasts and myofibroblasts. J Mol Cell Cardiol 29: 1947-1958, 1997. Campisi J. The biology of replicative senescence. Eur J Cancer 33: 703-709, 1997. Carey DJ. Syndecans: Multifunctional cell-surface co-receptors. Biochem J 327 (Pt 1): 1-16, 1997. Carisey A, Tsang R, Greiner AM, Nijenhuis N, Heath N, Nazgiewicz A, Kemkemer R, Derby B, Spatz J, Ballestrem C. Vinculin regulates the recruitment and release of core focal adhesion proteins in a forcedependent manner. Curr Biol 23: 271-281, 2013. Cary LA, Klinghoffer RA, Sachsenmaier C, Cooper JA. SRC catalytic but not scaffolding function is needed for integrin-regulated tyrosine phosphorylation, cell migration, and cell spreading. Mol Cell Biol 22: 2427-2440, 2002. Chan MW, Arora PD, Bozavikov P, McCulloch CA. FAK, PIP5KIgamma and gelsolin cooperatively mediate force-induced expression of alpha-smooth muscle actin. J Cell Sci 122: 2769-2781, 2009.

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

80. Chan MW, Chaudary F, Lee W, Copeland JW, McCulloch CA. Forceinduced myofibroblast differentiation through collagen receptors is dependent on mammalian diaphanous (mDia). J Biol Chem 285: 92739281, 2010. 81. Chang F, Lemmon CA, Park D, Romer LH. FAK potentiates Rac1 activation and localization to matrix adhesion sites: A role for betaPIX. Mol Biol Cell 18: 253-264, 2007. 82. Chang HY, Chi JT, Dudoit S, Bondre C, van de Rijn M, Botstein D, Brown PO. Diversity, topographic differentiation, and positional memory in human fibroblasts. Proc Natl Acad Sci U S A 99: 1287712882, 2002. 83. Chen CS. Mechanotransduction - a field pulling together? J Cell Sci 121: 3285-3292, 2008. 84. Chen HC, Appeddu PA, Parsons JT, Hildebrand JD, Schaller MD, Guan JL. Interaction of focal adhesion kinase with cytoskeletal protein talin. J Biol Chem 270: 16995-16999, 1995. 85. Chen K, Li D, Zhang X, Hermonat PL, Mehta JL. Anoxiareoxygenation stimulates collagen type-I and MMP-1 expression in cardiac fibroblasts: Modulation by the PPAR-gamma ligand pioglitazone. J Cardiovasc Pharm 44: 682-687, 2004. 86. Chen L, Wu Q, Guo F, Xia B, Zuo J. Expression of Dishevelled-1 in wound healing after acute myocardial infarction: Possible involvement in myofibroblast proliferation and migration. J Cell Mol Med 8: 257264, 2004. 87. Chen M, Masaki T, Sawamura T. LOX-1, the receptor for oxidized low-density lipoprotein identified from endothelial cells: Implications in endothelial dysfunction and atherosclerosis. Pharmacol Ther 95: 89-100, 2002. 88. Chen SJ, Yuan W, Lo S, Trojanowska M, Varga J. Interaction of smad3 with a proximal smad-binding element of the human alpha2(I) procollagen gene promoter required for transcriptional activation by TGF-beta. J Cell Physiol 183: 381-392, 2000. 89. Chen W, Saxena A, Li N, Sun J, Gupta A, Lee DW, Tian Q, Dobaczewski M, Frangogiannis NG. Endogenous IRAK-M attenuates postinfarction remodeling through effects on macrophages and fibroblasts. Arterioscl Throm Vas biol 32: 2598-2608, 2012. 90. Chen Y, Banda M, Speyer CL, Smith JS, Rabson AB, Gorski DH. Regulation of the expression and activity of the antiangiogenic homeobox gene GAX/MEOX2 by ZEB2 and microRNA-221. Mol Cell Biol 30: 3902-3913, 2010. 91. Chen Y, Shi-Wen X, van Beek J, Kennedy L, McLeod M, Renzoni EA, Bou-Gharios G, Wilcox-Adelman S, Goetinck PF, Eastwood M, Black CM, Abraham DJ, Leask A. Matrix contraction by dermal fibroblasts requires transforming growth factor-beta/activin-linked kinase 5, heparan sulfate-containing proteoglycans, and MEK/ERK: Insights into pathological scarring in chronic fibrotic disease. Am J Pathol 167: 1699-1711, 2005. 92. Chen YG, Liu F, Massague J. Mechanism of TGFbeta receptor inhibition by FKBP12. EMBO J 16: 3866-3876, 1997. 93. Cheng TH, Cheng PY, Shih NL, Chen IB, Wang DL, Chen JJ. Involvement of reactive oxygen species in angiotensin II-induced endothelin-1 gene expression in rat cardiac fibroblasts. J Am Coll Cardiol 42: 18451854, 2003. 94. Cheng Y, Liu X, Yang J, Lin Y, Xu DZ, Lu Q, Deitch EA, Huo Y, Delphin ES, Zhang C. MicroRNA-145, a novel smooth muscle cell phenotypic marker and modulator, controls vascular neointimal lesion formation. Circ Res 105: 158-166, 2009. 95. Chikumi H, Fukuhara S, Gutkind JS. Regulation of G protein-linked guanine nucleotide exchange factors for Rho, PDZ-RhoGEF, and LARG by tyrosine phosphorylation: Evidence of a role for focal adhesion kinase. J Biol Chem 277: 12463-12473, 2002. 96. Chiloeches A, Paterson HF, Marais R, Clerk A, Marshall CJ, Sugden PH. Regulation of Ras.GTP loading and Ras-Raf association in neonatal rat ventricular myocytes by G protein-coupled receptor agonists and phorbol ester. Activation of the extracellular signal-regulated kinase cascade by phorbol ester is mediated by Ras. J Biol Chem 274: 1976219770, 1999. 97. Chiquet M, Gelman L, Lutz R, Maier S. From mechanotransduction to extracellular matrix gene expression in fibroblasts. Biochim Biophys Acta 1793: 911-920, 2009. 98. Chohan PK, Singh RB, Dhalla NS, Netticadan T. L-arginine administration recovers sarcoplasmic reticulum function in ischemic reperfused hearts by preventing calpain activation. Cardiovasc Res 69: 152-163, 2006. 99. Choi CK, Vicente-Manzanares M, Zareno J, Whitmore LA, Mogilner A, Horwitz AR. Actin and alpha-actinin orchestrate the assembly and maturation of nascent adhesions in a myosin II motor-independent manner. Nat Cell Biol 10: 1039-1050, 2008. 100. Christia P, Bujak M, Gonzalez-Quesada C, Chen W, Dobaczewski M, Reddy A, Frangogiannis NG. Systematic characterization of myocardial inflammation, repair, and remodeling in a mouse model of reperfused myocardial infarction. J Histochem Cytochem 61: 555-570, 2013.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

101. Chu W, Li X, Li C, Wan L, Shi H, Song X, Liu X, Chen X, Zhang C, Shan H, Lu Y, Yang B. TGFBR3, a potential negative regulator of TGF-beta signaling, protects cardiac fibroblasts from hypoxia-induced apoptosis. J Cell Physiol 226: 2586-2594, 2011. 102. Chua CC, Hamdy RC, Chua BH. Angiotensin II induces TIMP-1 production in rat heart endothelial cells. Biochim Biophys Acta 1311: 175180, 1996. 103. Cicoira M, Rossi A, Bonapace S, Zanolla L, Golia G, Franceschini L, Caruso B, Marino PN, Zardini P. Independent and additional prognostic value of aminoterminal propeptide of type III procollagen circulating levels in patients with chronic heart failure. J Card Fail 10: 403-411, 2004. 104. Cleutjens JP, Kandala JC, Guarda E, Guntaka RV, Weber KT. Regulation of collagen degradation in the rat myocardium after infarction. J Mol Cell Cardiol 27: 1281-1292, 1995. 105. Cleutjens JP, Verluyten MJ, Smiths JF, Daemen MJ. Collagen remodeling after myocardial infarction in the rat heart. Am J Pathol 147: 325-338, 1995. 106. Colige A, Vandenberghe I, Thiry M, Lambert CA, Van Beeumen J, Li SW, Prockop DJ, Lapiere CM, Nusgens BV. Cloning and characterization of ADAMTS-14, a novel ADAMTS displaying high homology with ADAMTS-2 and ADAMTS-3. J Biol Chem 277: 5756-5766, 2002. 107. Colombelli J, Besser A, Kress H, Reynaud EG, Girard P, Caussinus E, Haselmann U, Small JV, Schwarz US, Stelzer EH. Mechanosensing in actin stress fibers revealed by a close correlation between force and protein localization. J Cell Sci 122: 1665-1679, 2009. 108. Conrad CH, Brooks WW, Hayes JA, Sen S, Robinson KG, Bing OH. Myocardial fibrosis and stiffness with hypertrophy and heart failure in the spontaneously hypertensive rat. Circulation 91: 161-170, 1995. 109. Cook TA, Nagasaki T, Gundersen GG. Rho guanosine triphosphatase mediates the selective stabilization of microtubules induced by lysophosphatidic acid. J Cell Biol 141: 175-185, 1998. 110. Corda S, Samuel JL, Rappaport L. Extracellular matrix and growth factors during heart growth. Heart Fail Rev 5: 119-130, 2000. 111. Cortez DM, Feldman MD, Mummidi S, Valente AJ, Steffensen B, Vincenti M, Barnes JL, Chandrasekar B. IL-17 stimulates MMP-1 expression in primary human cardiac fibroblasts via p38 MAPK- and ERK1/2-dependent C/EBP-beta, NF-ukappaB, and AP-1 activation. Am J Physiol Heart Circ Physiol 293: H3356-3365, 2007. 112. Crabos M, Roth M, Hahn AW, Erne P. Characterization of angiotensin II receptors in cultured adult rat cardiac fibroblasts. Coupling to signaling systems and gene expression. J Clin Invest 93: 2372-2378, 1994. 113. Cramer LP. Mechanism of cell rear retraction in migrating cells. Curr Opin Cell Biol 25: 591-599, 2013. 114. Cunnington RH, Northcott JM, Ghavami S, Filomeno KL, Jahan F, Kavosh MS, Davies JJ, Wigle JT, Dixon IM. The Ski-Zeb2-Meox2 pathway provides a novel mechanism for regulation of the cardiac myofibroblast phenotype. J Cell Sci 127: 40-49, 2014. 115. Cunnington RH, Wang B, Ghavami S, Bathe KL, Rattan SG, Dixon IM. Antifibrotic properties of c-Ski and its regulation of cardiac myofibroblast phenotype and contractility. Am J Physiol-Cell Ph 300: C176-186, 2011. 116. Czubryt MP. Common threads in cardiac fibrosis, infarct scar formation, and wound healing. Fibrogen Tissue Rep 5: 19, 2012. 117. D’Addario M, Arora PD, Ellen RP, McCulloch CA. Interaction of p38 and Sp1 in a mechanical force-induced, beta 1 integrin-mediated transcriptional circuit that regulates the actin-binding protein filamin-A. J Biol Chem 277: 47541-47550, 2002. 118. D’Angelo M, Billings PC, Pacifici M, Leboy PS, Kirsch T. Authentic matrix vesicles contain active metalloproteases (MMP). a role for matrix vesicle-associated MMP-13 in activation of transforming growth factor-beta. J Biol Chem 276: 11347-11353, 2001. 119. Dadson K, Chasiotis H, Wannaiampikul S, Tungtrongchitr R, Xu A, Sweeney G. Adiponectin mediated APPL1-AMPK signaling induces cell migration, MMP activation, and collagen remodeling in cardiac fibroblasts. J Cell Biochem 115: 785-793, 2014. 120. Dang I, Gorelik R, Sousa-Blin C, Derivery E, Guerin C, Linkner J, Nemethova M, Dumortier JG, Giger FA, Chipysheva TA, Ermilova VD, Vacher S, Campanacci V, Herrada I, Planson AG, Fetics S, Henriot V, David V, Oguievetskaia K, Lakisic G, Pierre F, Steffen A, Boyreau A, Peyrieras N, Rottner K, Zinn-Justin S, Cherfils J, Bieche I, Alexandrova AY, David NB, Small JV, Faix J, Blanchoin L, Gautreau A. Inhibitory signalling to the Arp2/3 complex steers cell migration. Nature 503: 281-284, 2013. 121. Darby I, Skalli O, Gabbiani G. Alpha-smooth muscle actin is transiently expressed by myofibroblasts during experimental wound healing. Lab Invest 63: 21-29, 1990. 122. Davis J, Burr AR, Davis GF, Birnbaumer L, Molkentin JD. A TRPC6dependent pathway for myofibroblast transdifferentiation and wound healing in vivo. Dev Cell 23: 705-715, 2012.

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

123. Davis J, Molkentin JD. Myofibroblasts: Trust your heart and let fate decide. J Mol Cell Cardiol 70: 9-18, 2014. 124. de Souza RR. Aging of myocardial collagen. Biogerontology 3: 325335, 2002. 125. Defilippi P, Di Stefano P, Cabodi S. p130Cas: A versatile scaffold in signaling networks. Trends Cell Biol 16: 257-263, 2006. 126. del Pozo MA, Alderson NB, Kiosses WB, Chiang HH, Anderson RG, Schwartz MA. Integrins regulate Rac targeting by internalization of membrane domains. Science 303: 839-842, 2004. 127. Delmas C, Manenti S, Boudjelal A, Peyssonnaux C, Eychene A, Darbon JM. The p42/p44 mitogen-activated protein kinase activation triggers p27Kip1 degradation independently of CDK2/cyclin E in NIH 3T3 cells. J Biol Chem 276: 34958-34965, 2001. 128. DeMali KA, Wennerberg K, Burridge K. Integrin signaling to the actin cytoskeleton. Curr Opin Cell Biol 15: 572-582, 2003. 129. Denhez F, Wilcox-Adelman SA, Baciu PC, Saoncella S, Lee S, French B, Neveu W, Goetinck PF. Syndesmos, a syndecan-4 cytoplasmic domain interactor, binds to the focal adhesion adaptor proteins paxillin and Hic-5. J Biol Chem 277: 12270-12274, 2002. 130. Deschamps AM, Spinale FG. Pathways of matrix metalloproteinase induction in heart failure: Bioactive molecules and transcriptional regulation. Cardiovasc Res 69: 666-676, 2006. 131. Desmouliere A, Gabbiani G. Modulation of fibroblastic cytoskeletal features during pathological situations: The role of extracellular matrix and cytokines. Cell Motil Cytoskeleton 29: 195-203, 1994. 132. Desmouliere A, Geinoz A, Gabbiani F, Gabbiani G. Transforming growth factor-beta 1 induces alpha-smooth muscle actin expression in granulation tissue myofibroblasts and in quiescent and growing cultured fibroblasts. J Cell Biol 122: 103-111, 1993. 133. Diehl JA, Cheng M, Roussel MF, Sherr CJ. Glycogen synthase kinase3beta regulates cyclin D1 proteolysis and subcellular localization. Genes Dev 12: 3499-3511, 1998. 134. Diez C, Nestler M, Friedrich U, Vieth M, Stolte M, Hu K, Hoppe J, Simm A. Down-regulation of Akt/PKB in senescent cardiac fibroblasts impairs PDGF-induced cell proliferation. Cardiovasc Res 49: 731-740, 2001. 135. Dinh DT, Frauman AG, Johnston CI, Fabiani ME. Angiotensin receptors: Distribution, signalling and function. Clin Sci (Lond) 100: 481492, 2001. 136. Dixon IM, Hao J, Reid NL, Roth JC. Effect of chronic AT(1) receptor blockade on cardiac Smad overexpression in hereditary cardiomyopathic hamsters. Cardiovasc Res 46: 286-297, 2000. 137. Dobaczewski M, Bujak M, Li N, Gonzalez-Quesada C, Mendoza LH, Wang XF, Frangogiannis NG. Smad3 signaling critically regulates fibroblast phenotype and function in healing myocardial infarction. Circ Res 107: 418-428, 2010. 138. Dong S, Ma W, Hao B, Hu F, Yan L, Yan X, Wang Y, Chen Z, Wang Z. microRNA-21 promotes cardiac fibrosis and development of heart failure with preserved left ventricular ejection fraction by up-regulating Bcl-2. Int J Clin Exp Pathol 7: 565-574, 2014. 139. Driesen RB, Nagaraju CK, Abi-Char J, Coenen T, Lijnen PJ, Fagard RH, Sipido KR, Petrov VV. Reversible and irreversible differentiation of cardiac fibroblasts. Cardiovasc Res 101: 411-422, 2014. 140. Du J, Xie J, Zhang Z, Tsujikawa H, Fusco D, Silverman D, Liang B, Yue L. TRPM7-mediated Ca2 +signals confer fibrogenesis in human atrial fibrillation. Circ Res 106: 992-1003, 2010. 141. Dubash AD, Wennerberg K, Garcia-Mata R, Menold MM, Arthur WT, Burridge K. A novel role for Lsc/p115 RhoGEF and LARG in regulating RhoA activity downstream of adhesion to fibronectin. J Cell Sci 120: 3989-3998, 2007. 142. Dubey RK, Gillespie DG, Mi Z, Jackson EK. Cardiac fibroblasts express the cAMP-adenosine pathway. Hypertension 36: 337-342, 2000. 143. Dubey RK, Gillespie DG, Mi Z, Jackson EK. Exogenous and endogenous adenosine inhibits fetal calf serum-induced growth of rat cardiac fibroblasts: Role of A2B receptors. Circulation 96: 2656-2666, 1997. 144. Dubey RK, Gillespie DG, Osaka K, Suzuki F, Jackson EK. Adenosine inhibits growth of rat aortic smooth muscle cells. Possible role of A2b receptor. Hypertension 27: 786-793, 1996. 145. Dugina V, Fontao L, Chaponnier C, Vasiliev J, Gabbiani G. Focal adhesion features during myofibroblastic differentiation are controlled by intracellular and extracellular factors. J Cell Sci 114: 3285-3296, 2001. 146. Dumbauld DW, Lee TT, Singh A, Scrimgeour J, Gersbach CA, Zamir EA, Fu J, Chen CS, Curtis JE, Craig SW, Garcia AJ. How vinculin regulates force transmission. Proc Natl Acad Sci U S A 110: 97889793, 2013. 147. Dunlevy JR, Couchman JR. Interleukin-8 induces motile behavior and loss of focal adhesions in primary fibroblasts. J Cell Sci 108(Pt 1): 311-321, 1995. 148. Eghbali M. Cardiac fibroblasts: Function, regulation of gene expression, and phenotypic modulation. Basic Res Cardiol 87(Suppl 2): 183-189, 1992.

751

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

149. Eghbali M, Weber KT. Collagen and the myocardium: Fibrillar structure, biosynthesis and degradation in relation to hypertrophy and its regression. Mol Cell Biochem 96: 1-14, 1990. 150. Eimer W, Niermann M, Eppe MA, Jockusch BM. Molecular shape of vinculin in aqueous solution. J Mol Biol 229: 146-152, 1993. 151. Ekholm SV, Reed SI. Regulation of G(1) cyclin-dependent kinases in the mammalian cell cycle. Curr Opin Cell Biol 12: 676-684, 2000. 152. Elledge SJ. Cell cycle checkpoints: Preventing an identity crisis. Science 274: 1664-1672, 1996. 153. Ellerbroek SM, Wennerberg K, Burridge K. Serine phosphorylation negatively regulates RhoA in vivo. J Biol Chem 278: 19023-19031, 2003. 154. Ellis IR, Jones SJ, Lindsay Y, Ohe G, Schor AM, Schor SL, Leslie NR. Migration Stimulating Factor (MSF) promotes fibroblast migration by inhibiting AKT. Cell Signal 22: 1655-1659, 2010. 155. Esparza-Lopez J, Montiel JL, Vilchis-Landeros MM, Okadome T, Miyazono K, Lopez-Casillas F. Ligand binding and functional properties of betaglycan, a co-receptor of the transforming growth factorbeta superfamily. Specialized binding regions for transforming growth factor-beta and inhibin A. J Biol Chem 276: 14588-14596, 2001. 156. Espira L, Lamoureux L, Jones SC, Gerard RD, Dixon IM, Czubryt MP. The basic helix-loop-helix transcription factor scleraxis regulates fibroblast collagen synthesis. J Mol Cell Cardiol 47: 188-195, 2009. 157. Evans RA, Tian YC, Steadman R, Phillips AO. TGF-beta1-mediated fibroblast-myofibroblast terminal differentiation-the role of Smad proteins. Exp Cell Res 282: 90-100, 2003. 158. Felsenfeld DP, Schwartzberg PL, Venegas A, Tse R, Sheetz MP. Selective regulation of integrin–cytoskeleton interactions by the tyrosine kinase Src. Nat Cell Biol 1: 200-206, 1999. 159. Feng J, Ito M, Ichikawa K, Isaka N, Nishikawa M, Hartshorne DJ, Nakano T. Inhibitory phosphorylation site for Rho-associated kinase on smooth muscle myosin phosphatase. J Biol Chem 274: 37385-37390, 1999. 160. Fisher SA, Absher M. Norepinephrine and ANG II stimulate secretion of TGF-beta by neonatal rat cardiac fibroblasts in vitro. Am J Physiol 268: C910-917, 1995. 161. Fitzgerald KA, Bowie AG, Skeffington BS, O’Neill LA. Ras, protein kinase C zeta, and I kappa B kinases 1 and 2 are downstream effectors of CD44 during the activation of NF-kappa B by hyaluronic acid fragments in T-24 carcinoma cells. J Immunol 164: 2053-2063, 2000. 162. Fonseca PM, Shin NY, Brabek J, Ryzhova L, Wu J, Hanks SK. Regulation and localization of CAS substrate domain tyrosine phosphorylation. Cell Signal 16: 621-629, 2004. 163. Frangogiannis NG. Matricellular proteins in cardiac adaptation and disease. Physiol Rev 92: 635-688, 2012. 164. Frangogiannis NG. The mechanistic basis of infarct healing. Antioxid Redox Signal 8: 1907-1939, 2006. 165. Frangogiannis NG. Regulation of the inflammatory response in cardiac repair. Circ Res 110: 159-173, 2012. 166. Frangogiannis NG, Michael LH, Entman ML. Myofibroblasts in reperfused myocardial infarcts express the embryonic form of smooth muscle myosin heavy chain (SMemb). Cardiovasc Res 48: 89-100, 2000. 167. Freed DH, Chilton L, Li Y, Dangerfield AL, Raizman JE, Rattan SG, Visen N, Hryshko LV, Dixon IM. Role of myosin light chain kinase in cardiotrophin-1-induced cardiac myofibroblast cell migration. Am J Physiol Heart Circ Physiol 301: H514-522, 2011. 168. Freed DH, Moon MC, Borowiec AM, Jones SC, Zahradka P, Dixon IM. Cardiotrophin-1: Expression in experimental myocardial infarction and potential role in post-MI wound healing. Mol Cell Biochem 254: 247-256, 2003. 169. Frey RS, Mulder KM. Involvement of extracellular signal-regulated kinase 2 and stress-activated protein kinase/Jun N-terminal kinase activation by transforming growth factor beta in the negative growth control of breast cancer cells. Cancer Res 57: 628-633, 1997. 170. Frey RS, Mulder KM. TGFbeta regulation of mitogen-activated protein kinases in human breast cancer cells. Cancer Lett 117: 41-50, 1997. 171. Fujishiro SH, Tanimura S, Mure S, Kashimoto Y, Watanabe K, Kohno M. ERK1/2 phosphorylate GEF-H1 to enhance its guanine nucleotide exchange activity toward RhoA. Biochem Bioph Res Co 368: 162-167, 2008. 172. Fukata M, Nakagawa M, Kaibuchi K. Roles of Rho-family GTPases in cell polarisation and directional migration. Curr Opin Cell Biol 15: 590-597, 2003. 173. Gabbiani G, Hirschel BJ, Ryan GB, Statkov PR, Majno G. Granulation tissue as a contractile organ. A study of structure and function. J Exp Med 135: 719-734, 1972. 174. Ganz A, Lambert M, Saez A, Silberzan P, Buguin A, Mege RM, Ladoux B. Traction forces exerted through N-cadherin contacts. Biol Cell 98: 721-730, 2006. 175. Gao X, He X, Luo B, Peng L, Lin J, Zuo Z. Angiotensin II increases collagen I expression via transforming growth factor-beta1 and extracellular signal-regulated kinase in cardiac fibroblasts. Eur J Pharmacol 606: 115-120, 2009.

752

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

176. Geiger B, Bershadsky A. Assembly and mechanosensory function of focal contacts. Curr Opin Cell Biol 13: 584-592, 2001. 177. Geiger B, Bershadsky A, Pankov R, Yamada KM. Transmembrane crosstalk between the extracellular matrix–cytoskeleton crosstalk. Nature Rev Mol Cell Biol 2: 793-805, 2001. 178. Geiger B, Spatz JP, Bershadsky AD. Environmental sensing through focal adhesions. Nature Rev Mol Cell Biol 10: 21-33, 2009. 179. Ginsberg MH, Wencel JD, White JG, Plow EF. Binding of fibronectin to alpha-granule-deficient platelets. J Cell Biol 97: 571-573, 1983. 180. Goffin JM, Pittet P, Csucs G, Lussi JW, Meister JJ, Hinz B. Focal adhesion size controls tension-dependent recruitment of alpha-smooth muscle actin to stress fibers. J Cell Biol 172: 259-268, 2006. 181. Gondokaryono SP, Ushio H, Niyonsaba F, Hara M, Takenaka H, Jayawardana ST, Ikeda S, Okumura K, Ogawa H. The extra domain A of fibronectin stimulates murine mast cells via toll-like receptor 4. J Leukoc Biol 82: 657-665, 2007. 182. Grigorian M, Andresen S, Tulchinsky E, Kriajevska M, Carlberg C, Kruse C, Cohn M, Ambartsumian N, Christensen A, Selivanova G, Lukanidin E. Tumor suppressor p53 protein is a new target for the metastasis-associated Mts1/S100A4 protein: Functional consequences of their interaction. J Biol Chem 276: 22699-22708, 2001. 183. Grinnell F. Fibroblasts, myofibroblasts, and wound contraction. J Cell Biol 124: 401-404, 1994. 184. Guilluy C, Garcia-Mata R, Burridge K. Rho protein crosstalk: Another social network? Trends Cell Biol 21: 718-726, 2011. 185. Guilluy C, Swaminathan V, Garcia-Mata R, O’Brien ET, Superfine R, Burridge K. The Rho GEFs LARG and GEF-H1 regulate the mechanical response to force on integrins. Nat Cell Biol 13: 722-727, 2011. 186. Gunning P, O’Neill G, Hardeman E. Tropomyosin-based regulation of the actin cytoskeleton in time and space. Physiol Rev 88: 1-35, 2008. 187. Gunning PW, Schevzov G, Kee AJ, Hardeman EC. Tropomyosin isoforms: Divining rods for actin cytoskeleton function. Trends Cell Biol 15: 333-341, 2005. 188. Guo WH, Wang YL. Retrograde fluxes of focal adhesion proteins in response to cell migration and mechanical signals. Mol Biol Cell 18: 4519-4527, 2007. 189. Guo XG, Uzui H, Mizuguchi T, Ueda T, Chen JZ, Lee JD. Imidaprilat inhibits matrix metalloproteinase-2 activity in human cardiac fibroblasts induced by interleukin-1beta via NO-dependent pathway. Int J Cardiol 126: 414-420, 2008. 190. Gupton SL, Anderson KL, Kole TP, Fischer RS, Ponti A, HitchcockDeGregori SE, Danuser G, Fowler VM, Wirtz D, Hanein D, WatermanStorer CM. Cell migration without a lamellipodium: Translation of actin dynamics into cell movement mediated by tropomyosin. J Cell Biol 168: 619-631, 2005. 191. Halliday NL, Tomasek JJ. Mechanical properties of the extracellular matrix influence fibronectin fibril assembly in vitro. Exp Cell Res 217: 109-117, 1995. 192. Han J, Lim CJ, Watanabe N, Soriani A, Ratnikov B, Calderwood DA, Puzon-McLaughlin W, Lafuente EM, Boussiotis VA, Shattil SJ, Ginsberg MH. Reconstructing and deconstructing agonist-induced activation of integrin alphaIIbbeta3. Curr Biol 16: 1796-1806, 2006. 193. Hanahan D, Weinberg RA. The hallmarks of cancer. Cell 100: 57-70, 2000. 194. Harada M, Luo X, Qi XY, Tadevosyan A, Maguy A, Ordog B, Ledoux J, Kato T, Naud P, Voigt N, Shi Y, Kamiya K, Murohara T, Kodama I, Tardif JC, Schotten U, Van Wagoner DR, Dobrev D, Nattel S. Transient receptor potential canonical-3 channel-dependent fibroblast regulation in atrial fibrillation. Circulation 126: 2051-2064, 2012. 195. Harris ES, Li F, Higgs HN. The mouse formin, FRLalpha, slows actin filament barbed end elongation, competes with capping protein, accelerates polymerization from monomers, and severs filaments. J Biol Chem 279: 20076-20087, 2004. 196. Hauck CR, Sieg DJ, Hsia DA, Loftus JC, Gaarde WA, Monia BP, Schlaepfer DD. Inhibition of focal adhesion kinase expression or activity disrupts epidermal growth factor-stimulated signaling promoting the migration of invasive human carcinoma cells. Cancer Res 61: 70797090, 2001. 197. Haudek SB, Cheng J, Du J, Wang Y, Hermosillo-Rodriguez J, Trial J, Taffet GE, Entman ML. Monocytic fibroblast precursors mediate fibrosis in angiotensin-II-induced cardiac hypertrophy. J Mol Cell Cardiol 49: 499-507, 2010. 198. Hayakawa K, Tatsumi H, Sokabe M. Actin filaments function as a tension sensor by tension-dependent binding of cofilin to the filament. J Cell Biol 195: 721-727, 2011. 199. He Y, Esser P, Schacht V, Bruckner-Tuderman L, Has C. Role of kindlin-2 in fibroblast functions: Implications for wound healing. J Invest Dermatol 131: 245-256, 2011. 200. Heath JP, Holifield BF. On the mechanisms of cortical actin flow and its role in cytoskeletal organisation of fibroblasts. Sym Soc Exp Biol 47: 35-56, 1993.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

201. Heldin CH, Miyazono K, ten Dijke P. TGF-beta signalling from cell membrane to nucleus through SMAD proteins. Nature 390: 465-471, 1997. 202. Helsten TL, Bunch TA, Kato H, Yamanouchi J, Choi SH, Jannuzi AL, Feral CC, Ginsberg MH, Brower DL, Shattil SJ. Differences in regulation of Drosophila and vertebrate integrin affinity by talin. Mol Biol Cell 19: 3589-3598, 2008. 203. Herman IM, Crisona NJ, Pollard TD. Relation between cell activity and the distribution of cytoplasmic actin and myosin. J Cell Biol 90: 84-91, 1981. 204. Heymans S, Schroen B, Vermeersch P, Milting H, Gao F, Kassner A, Gillijns H, Herijgers P, Flameng W, Carmeliet P, Van de Werf F, Pinto YM, Janssens S. Increased cardiac expression of tissue inhibitor of metalloproteinase-1 and tissue inhibitor of metalloproteinase-2 is related to cardiac fibrosis and dysfunction in the chronic pressureoverloaded human heart. Circulation 112: 1136-1144, 2005. 205. Hilfiker-Kleiner D, Hilfiker A, Fuchs M, Kaminski K, Schaefer A, Schieffer B, Hillmer A, Schmiedl A, Ding Z, Podewski E, Poli V, Schneider MD, Schulz R, Park JK, Wollert KC, Drexler H. Signal transducer and activator of transcription 3 is required for myocardial capillary growth, control of interstitial matrix deposition, and heart protection from ischemic injury. Circ Res 95: 187-195, 2004. 206. Hinz B. Formation and function of the myofibroblast during tissue repair. J Invest Dermatol 127: 526-537, 2007. 207. Hinz B. The myofibroblast: Paradigm for a mechanically active cell. J Biomech 43: 146-155, 2010. 208. Hinz B. Tissue stiffness, latent TGF-beta1 activation, and mechanical signal transduction: Implications for the pathogenesis and treatment of fibrosis. Curr Rheumatol Rep 11: 120-126, 2009. 209. 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. 210. Hirata H, Tatsumi H, Lim CT, Sokabe M. Force-dependent vinculin binding to talin in live cells: A crucial step in anchoring the actin cytoskeleton to focal adhesions. Am J Physiol-Cell Ph 306: C607-620, 2014. 211. Holmes JW, Borg TK, Covell JW. Structure and mechanics of healing myocardial infarcts. Annu Rev Biomed Eng 7: 223-253, 2005. 212. Holtz J. Pathophysiology of heart failure and the renin-angiotensinsystem. Basic Res Cardiol 88(Suppl 1): 183-201, 1993. 213. Horazdovsky W. Acquisition and selection of the educated. Quintessenz J 21: 817-820, 1991. 214. Hotulainen P, Lappalainen P. Stress fibers are generated by two distinct actin assembly mechanisms in motile cells. J Cell Biol 173: 383-394, 2006. 215. Hsia DA, Mitra SK, Hauck CR, Streblow DN, Nelson JA, Ilic D, Huang S, Li E, Nemerow GR, Leng J, Spencer KS, Cheresh DA, Schlaepfer DD. Differential regulation of cell motility and invasion by FAK. J Cell Biol 160: 753-767, 2003. 216. Hu K, Ji L, Applegate KT, Danuser G, Waterman-Storer CM. Differential transmission of actin motion within focal adhesions. Science 315: 111-115, 2007. 217. Huang X, Yang N, Fiore VF, Barker TH, Sun Y, Morris SW, Ding Q, Thannickal VJ, Zhou Y. Matrix stiffness-induced myofibroblast differentiation is mediated by intrinsic mechanotransduction. Am J Resp Cell Mol Biol 47: 340-348, 2012. 218. Humphries JD, Wang P, Streuli C, Geiger B, Humphries MJ, Ballestrem C. Vinculin controls focal adhesion formation by direct interactions with talin and actin. J Cell Biol 179: 1043-1057, 2007. 219. Humphries MJ. Integrin structure. Biochem Soc Tran 28: 311-339, 2000. 220. Husse B, Briest W, Homagk L, Isenberg G, Gekle M. Cyclical mechanical stretch modulates expression of collagen I and collagen III by PKC and tyrosine kinase in cardiac fibroblasts. Am J Physiol Regul Integr Comp Physiol 293: R1898-R1907, 2007. 221. Huveneers S, van den Bout I, Sonneveld P, Sancho A, Sonnenberg A, Danen EH. Integrin alpha v beta 3 controls activity and oncogenic potential of primed c-Src. Cancer Res 67: 2693-2700, 2007. 222. Hynes RO. Integrins: A family of cell surface receptors. Cell 48: 549554, 1987. 223. Ilic D, Furuta Y, Kanazawa S, Takeda N, Sobue K, Nakatsuji N, Nomura S, Fujimoto J, Okada M, Yamamoto T. Reduced cell motility and enhanced focal adhesion contact formation in cells from FAK-deficient mice. Nature 377: 539-544, 1995. 224. Ingber DE. Cellular basis of mechanotransduction. Biol Bullet 194: 323-325; discussion 325-327, 1998. 225. Inoue A, Yanagisawa M, Kimura S, Kasuya Y, Miyauchi T, Goto K, Masaki T. The human endothelin family: Three structurally and pharmacologically distinct isopeptides predicted by three separate genes. Proc Natl Acad Sci U S A 86: 2863-2867, 1989. 226. Inoue A, Yanagisawa M, Takuwa Y, Mitsui Y, Kobayashi M, Masaki T. The human preproendothelin-1 gene. Complete nucleotide sequence and regulation of expression. J Biol Chem 264: 14954-14959, 1989.

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

227. Ito M, Nakano T, Erdodi F, Hartshorne DJ. Myosin phosphatase: Structure, regulation and function. Mol Cell Biochem 259: 197-209, 2004. 228. Janmey PA, Wells RG, Assoian RK, McCulloch CA. From tissue mechanics to transcription factors. Differentiation 86: 112-120, 2013. 229. Jazbutyte V, Fiedler J, Kneitz S, Galuppo P, Just A, Holzmann A, Bauersachs J, Thum T. MicroRNA-22 increases senescence and activates cardiac fibroblasts in the aging heart. Age (Dordr) 35: 747-762, 2013. 230. Jiang X, Chen S, Asara JM, Balk SP. Phosphoinositide 3-kinase pathway activation in phosphate and tensin homolog (PTEN)-deficient prostate cancer cells is independent of receptor tyrosine kinases and mediated by the p110beta and p110delta catalytic subunits. J Biol Chem 285: 14980-14989, 2010. 231. Kahner BN, Kato H, Banno A, Ginsberg MH, Shattil SJ, Ye F. Kindlins, integrin activation and the regulation of talin recruitment to alphaIIbbeta3. PloS One 7: e34056, 2012. 232. Kakiashvili E, Speight P, Waheed F, Seth R, Lodyga M, Tanimura S, Kohno M, Rotstein OD, Kapus A, Szaszi K. GEF-H1 mediates tumor necrosis factor-alpha-induced Rho activation and myosin phosphorylation: Role in the regulation of tubular paracellular permeability. J Biol Chem 284: 11454-11466, 2009. 233. Kamikura DM, Khoury H, Maroun C, Naujokas MA, Park M. Enhanced transformation by a plasma membrane-associated met oncoprotein: Activation of a phosphoinositide 3′ -kinase-dependent autocrine loop involving hyaluronic acid and CD44. Mol Cell Biol 20: 3482-3496, 2000. 234. Kandalam V, Basu R, Abraham T, Wang X, Soloway PD, Jaworski DM, Oudit GY, Kassiri Z. TIMP2 deficiency accelerates adverse postmyocardial infarction remodeling because of enhanced MT1-MMP activity despite lack of MMP2 activation. Circ Res 106: 796-808, 2010. 235. Kanemura H, Iimuro Y, Takeuchi M, Ueki T, Hirano T, Horiguchi K, Asano Y, Fujimoto J. Hepatocyte growth factor gene transfer with naked plasmid DNA ameliorates dimethylnitrosamine-induced liver fibrosis in rats. Hepatol Res 38: 930-939, 2008. 236. Kapur NK, Qiao X, Paruchuri V, Mackey EE, Daly GH, Ughreja K, Morine KJ, Levine J, Aronovitz MJ, Hill NS, Jaffe IZ, Letarte M, Karas RH. Reducing endoglin activity limits calcineurin and TRPC-6 expression and improves survival in a mouse model of right ventricular pressure overload. J Am Heart Assoc 3: 2014. 237. Katoh K, Kano Y, Amano M, Kaibuchi K, Fujiwara K. Stress fiber organization regulated by MLCK and Rho-kinase in cultured human fibroblasts. Am J Physiol-Cell Ph 280: C1669-1679, 2001. 238. Katoh K, Kano Y, Amano M, Onishi H, Kaibuchi K, Fujiwara K. Rhokinase–mediated contraction of isolated stress fibers. J Cell Biol 153: 569-584, 2001. 239. Katoh K, Kano Y, Fujiwara K. Isolation and in vitro contraction of stress fibers. Method Enzymol 325: 369-380, 2000. 240. Katoh K, Kano Y, Masuda M, Onishi H, Fujiwara K. Isolation and contraction of the stress fiber. Mol Biol Cell 9: 1919-1938, 1998. 241. Katoh K, Kano Y, Ookawara S. Rho-kinase dependent organization of stress fibers and focal adhesions in cultured fibroblasts. Genes Cells 12: 623-638, 2007. 242. Kavsak P, Rasmussen RK, Causing CG, Bonni S, Zhu H, Thomsen GH, Wrana JL. Smad7 binds to Smurf2 to form an E3 ubiquitin ligase that targets the TGF beta receptor for degradation. Mol Cell 6: 1365-1375, 2000. 243. Kawabata S, Usukura J, Morone N, Ito M, Iwamatsu A, Kaibuchi K, Amano M. Interaction of Rho-kinase with myosin II at stress fibres. Genes Cells 9: 653-660, 2004. 244. Kawaguchi M, Takahashi M, Hata T, Kashima Y, Usui F, Morimoto H, Izawa A, Takahashi Y, Masumoto J, Koyama J, Hongo M, Noda T, Nakayama J, Sagara J, Taniguchi S, Ikeda U. Inflammasome activation of cardiac fibroblasts is essential for myocardial ischemia/reperfusion injury. Circulation 123: 594-604, 2011. 245. Kawano Y, Fukata Y, Oshiro N, Amano M, Nakamura T, Ito M, Matsumura F, Inagaki M, Kaibuchi K. Phosphorylation of myosin-binding subunit (MBS) of myosin phosphatase by Rho-kinase in vivo. J Cell Biol 147: 1023-1038, 1999. 246. Kelly D, Khan S, Cockerill G, Ng LL, Thompson M, Samani NJ, Squire IB. Circulating stromelysin-1 (MMP-3): A novel predictor of LV dysfunction, remodelling and all-cause mortality after acute myocardial infarction. Eur J Heart Failure 10: 133-139, 2008. 247. Kessler E, Takahara K, Biniaminov L, Brusel M, Greenspan DS. Bone morphogenetic protein-1: The type I procollagen C-proteinase. Science 271: 360-362, 1996. 248. Kida Y, Kobayashi M, Suzuki T, Takeshita A, Okamatsu Y, Hanazawa S, Yasui T, Hasegawa K. Interleukin-1 stimulates cytokines, prostaglandin E2 and matrix metalloproteinase-1 production via activation of MAPK/AP-1 and NF-kappaB in human gingival fibroblasts. Cytokine 29: 159-168, 2005. 249. Kiema T, Lad Y, Jiang P, Oxley CL, Baldassarre M, Wegener KL, Campbell ID, Ylanne J, Calderwood DA. The molecular basis of filamin

753

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

250. 251.

252. 253.

254.

255. 256.

257. 258.

259. 260. 261.

262.

263. 264. 265.

266. 267. 268.

269. 270. 271. 272.

754

binding to integrins and competition with talin. Mol Cell 21: 337-347, 2006. Kim C, Ye F, Ginsberg MH. Regulation of integrin activation. Annu Rev Cell Dev Biol 27: 321-345, 2011. Kim H, Nakamura F, Lee W, Hong C, Perez-Sala D, McCulloch CA. Regulation of cell adhesion to collagen via beta1 integrins is dependent on interactions of filamin A with vimentin and protein kinase C epsilon. Exp Cell Res 316: 1829-1844, 2010. Kim H, Sengupta A, Glogauer M, McCulloch CA. Filamin A regulates cell spreading and survival via beta1 integrins. Exp Cell Res 314: 834846, 2008. Kimura K, Ito M, Amano M, Chihara K, Fukata Y, Nakafuku M, Yamamori B, Feng J, Nakano T, Okawa K, Iwamatsu A, Kaibuchi K. Regulation of myosin phosphatase by Rho and Rho-associated kinase (Rho-kinase). Science 273: 245-248, 1996. Kinoshita T, Ishikawa Y, Arita M, Akishima-Fukasawa Y, Fujita K, Inomata N, Suzuki T, Namiki A, Mikami T, Ikeda T, Yamazaki J, Ishii T, Akasaka Y. Antifibrotic response of cardiac fibroblasts in hypertensive hearts through enhanced TIMP-1 expression by basic fibroblast growth factor. Cardiovasc Pathol 23: 92-100, 2014. Kishimoto T, Okumura E. In vivo regulation of the entry into M-phase: Initial activation and nuclear translocation of cyclin B/Cdc2. Prog Cell Cycle Res 3: 241-249, 1997. Klein EA, Yin L, Kothapalli D, Castagnino P, Byfield FJ, Xu T, Levental I, Hawthorne E, Janmey PA, Assoian RK. Cell-cycle control by physiological matrix elasticity and in vivo tissue stiffening. Curr Biol 19: 1511-1518, 2009. Klemke RL, Leng J, Molander R, Brooks PC, Vuori K, Cheresh DA. CAS/Crk coupling serves as a “molecular switch” for induction of cell migration. J Cell Biol 140: 961-972, 1998. Klingbeil CK, Hauck CR, Hsia DA, Jones KC, Reider SR, Schlaepfer DD. Targeting Pyk2 to beta 1-integrin-containing focal contacts rescues fibronectin-stimulated signaling and haptotactic motility defects of focal adhesion kinase-null cells. J Cell Biol 152: 97-110, 2001. Ko KS, Arora PD, McCulloch CA. Cadherins mediate intercellular mechanical signaling in fibroblasts by activation of stretch-sensitive calcium-permeable channels. J Biol Chem 276: 35967-35977, 2001. Kobielak A, Fuchs E. Alpha-catenin: At the junction of intercellular adhesion and actin dynamics. Nature Rev Mol Cell Biol 5: 614-625, 2004. Koestler SA, Steffen A, Nemethova M, Winterhoff M, Luo N, Holleboom JM, Krupp J, Jacob S, Vinzenz M, Schur F, Schluter K, Gunning PW, Winkler C, Schmeiser C, Faix J, Stradal TE, Small JV, Rottner K. Arp2/3 complex is essential for actin network treadmilling as well as for targeting of capping protein and cofilin. Mol Biol Cell 24: 2861-2875, 2013. Koitabashi N, Danner T, Zaiman AL, Pinto YM, Rowell J, Mankowski J, Zhang D, Nakamura T, Takimoto E, Kass DA. Pivotal role of cardiomyocyte TGF-beta signaling in the murine pathological response to sustained pressure overload. J Clin Invest 121: 2301-2312, 2011. Kong-Beltran M, Stamos J, Wickramasinghe D. The Sema domain of Met is necessary for receptor dimerization and activation. Cancer cell 6: 75-84, 2004. Kooistra MR, Dube N, Bos JL. Rap1: A key regulator in cell-cell junction formation. J Cell Sci 120: 17-22, 2007. Koudssi F, Lopez JE, Villegas S, Long CS. Cardiac fibroblasts arrest at the G1/S restriction point in response to interleukin (IL)-1beta. Evidence for IL-1beta-induced hypophosphorylation of the retinoblastoma protein. J Biol Chem 273: 25796-25803, 1998. Kozma R, Ahmed S, Best A, Lim L. The Ras-related protein Cdc42Hs and bradykinin promote formation of peripheral actin microspikes and filopodia in Swiss 3T3 fibroblasts. Mol Cell Biol 15: 1942-1952, 1995. Krause M, Dent EW, Bear JE, Loureiro JJ, Gertler FB. Ena/VASP proteins: Regulators of the actin cytoskeleton and cell migration. Annu Rev Cell Dev Biol 19: 541-564, 2003. Kumar S, Maxwell IZ, Heisterkamp A, Polte TR, Lele TP, Salanga M, Mazur E, Ingber DE. Viscoelastic retraction of single living stress fibers and its impact on cell shape, cytoskeletal organization, and extracellular matrix mechanics. Biophys J 90: 3762-3773, 2006. Kuwahara K, Barrientos T, Pipes GC, Li S, Olson EN. Muscle-specific signaling mechanism that links actin dynamics to serum response factor. Mol Cell Biol 25: 3173-3181, 2005. Kuwahara K, Wang Y, McAnally J, Richardson JA, Bassel-Duby R, Hill JA, Olson EN. TRPC6 fulfills a calcineurin signaling circuit during pathologic cardiac remodeling. J Clin Invest 116: 3114-3126, 2006. Laeremans H, Rensen SS, Ottenheijm HC, Smits JF, Blankesteijn WM. Wnt/frizzled signalling modulates the migration and differentiation of immortalized cardiac fibroblasts. Cardiovasc Res 87: 514-523, 2010. Lafuente EM, van Puijenbroek AA, Krause M, Carman CV, Freeman GJ, Berezovskaya A, Constantine E, Springer TA, Gertler FB, Boussiotis VA. RIAM, an Ena/VASP and Profilin ligand, interacts with Rap1-GTP and mediates Rap1-induced adhesion. Dev Cell 7: 585-595, 2004.

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

273. Lamb NJ, Fernandez A, Conti MA, Adelstein R, Glass DB, Welch WJ, Feramisco JR. Regulation of actin microfilament integrity in living nonmuscle cells by the cAMP-dependent protein kinase and the myosin light chain kinase. J Cell Biol 106: 1955-1971, 1988. 274. Lang P, Gesbert F, Delespine-Carmagnat M, Stancou R, Pouchelet M, Bertoglio J. Protein kinase A phosphorylation of RhoA mediates the morphological and functional effects of cyclic AMP in cytotoxic lymphocytes. EMBO J 15: 510-519, 1996. 275. Lappalainen P, Drubin DG. Cofilin promotes rapid actin filament turnover in vivo. Nature 388: 78-82, 1997. 276. Lawson C, Lim ST, Uryu S, Chen XL, Calderwood DA, Schlaepfer DD. FAK promotes recruitment of talin to nascent adhesions to control cell motility. J Cell Biol 196: 223-232, 2012. 277. Lazard D, Sastre X, Frid MG, Glukhova MA, Thiery JP, Koteliansky VE. Expression of smooth muscle-specific proteins in myoepithelium and stromal myofibroblasts of normal and malignant human breast tissue. Proc Natl Acad Sci U S A 90: 999-1003, 1993. 278. Leask A. Potential therapeutic targets for cardiac fibrosis: TGFbeta, angiotensin, endothelin, CCN2, and PDGF, partners in fibroblast activation. Circ Res 106: 1675-1680, 2010. 279. Leask A. Targeting the TGFbeta, endothelin-1 and CCN2 axis to combat fibrosis in scleroderma. Cell Signal 20: 1409-1414, 2008. 280. Leask A. TGFbeta, cardiac fibroblasts, and the fibrotic response. Cardiovasc Res 74: 207-212, 2007. 281. Leask A, Abraham DJ. TGF-beta signaling and the fibrotic response. FASEB J 18: 816-827, 2004. 282. Leask A, Holmes A, Black CM, Abraham DJ. Connective tissue growth factor gene regulation. Requirements for its induction by transforming growth factor-beta 2 in fibroblasts. J Biol Chem 278: 13008-13015, 2003. 283. LeBleu VS, Taduri G, O’Connell J, Teng Y, Cooke VG, Woda C, Sugimoto H, Kalluri R. Origin and function of myofibroblasts in kidney fibrosis. Nature Med 19: 1047-1053, 2013. 284. Lee AA, Dillmann WH, McCulloch AD, Villarreal FJ. Angiotensin II stimulates the autocrine production of transforming growth factor-beta 1 in adult rat cardiac fibroblasts. J Mol Cell Cardiol 27: 2347-2357, 1995. 285. Lee HS, Lim CJ, Puzon-McLaughlin W, Shattil SJ, Ginsberg MH. RIAM activates integrins by linking talin to ras GTPase membranetargeting sequences. J Biol Chem 284: 5119-5127, 2009. 286. Li P, Wang D, Lucas J, Oparil S, Xing D, Cao X, Novak L, Renfrow MB, Chen YF. Atrial natriuretic peptide inhibits transforming growth factor beta-induced Smad signaling and myofibroblast transformation in mouse cardiac fibroblasts. Circ Res 102: 185-192, 2008. 287. Li YY, Feldman AM, Sun Y, McTiernan CF. Differential expression of tissue inhibitors of metalloproteinases in the failing human heart. Circulation 98: 1728-1734, 1998. 288. Li YY, Feng Y, McTiernan CF, Pei W, Moravec CS, Wang P, Rosenblum W, Kormos RL, Feldman AM. Downregulation of matrix metalloproteinases and reduction in collagen damage in the failing human heart after support with left ventricular assist devices. Circulation 104: 1147-1152, 2001. 289. Li Z, Wang C, Jiao X, Lu Y, Fu M, Quong AA, Dye C, Yang J, Dai M, Ju X, Zhang X, Li A, Burbelo P, Stanley ER, Pestell RG. Cyclin D1 regulates cellular migration through the inhibition of thrombospondin 1 and ROCK signaling. Mol Cell Biol 26: 4240-4256, 2006. 290. Liao YF, Gotwals PJ, Koteliansky VE, Sheppard D, Van De Water L. The EIIIA segment of fibronectin is a ligand for integrins alpha 9beta 1 and alpha 4beta 1 providing a novel mechanism for regulating cell adhesion by alternative splicing. J Biol Chem 277: 14467-14474, 2002. 291. Lie-Venema H, van den Akker NM, Bax NA, Winter EM, Maas S, Kekarainen T, Hoeben RC, deRuiter MC, Poelmann RE, Gittenbergerde Groot AC. Origin, fate, and function of epicardium-derived cells (EPDCs) in normal and abnormal cardiac development. ScientificWorldJournal 7: 1777-1798, 2007. 292. Lim Y, Han I, Jeon J, Park H, Bahk YY, Oh ES. Phosphorylation of focal adhesion kinase at tyrosine 861 is crucial for Ras transformation of fibroblasts. J Biol Chem 279: 29060-29065, 2004. 293. Lim Y, Lim ST, Tomar A, Gardel M, Bernard-Trifilo JA, Chen XL, Uryu SA, Canete-Soler R, Zhai J, Lin H, Schlaepfer WW, Nalbant P, Bokoch G, Ilic D, Waterman-Storer C, Schlaepfer DD. PyK2 and FAK connections to p190Rho guanine nucleotide exchange factor regulate RhoA activity, focal adhesion formation, and cell motility. J Cell Biol 180: 187-203, 2008. 294. Lindsey ML. MMP induction and inhibition in myocardial infarction. Heart Fail Rev 9: 7-19, 2004. 295. Liu S, Xu SW, Kennedy L, Pala D, Chen Y, Eastwood M, Carter DE, Black CM, Abraham DJ, Leask A. FAK is required for TGFbetainduced JNK phosphorylation in fibroblasts: Implications for acquisition of a matrix-remodeling phenotype. Mol Biol Cell 18: 2169-2178, 2007. 296. Liu X, Sun SQ, Hassid A, Ostrom RS. cAMP inhibits transforming growth factor-beta-stimulated collagen synthesis via inhibition of

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Comprehensive Physiology

297. 298. 299. 300. 301. 302.

303.

304. 305.

306.

307. 308.

309. 310. 311.

312. 313. 314. 315. 316.

317. 318.

319.

320. 321.

extracellular signal-regulated kinase 1/2 and Smad signaling in cardiac fibroblasts. Mol Pharmacol 70: 1992-2003, 2006. Lo RS, Wotton D, Massague J. Epidermal growth factor signaling via Ras controls the Smad transcriptional co-repressor TGIF. EMBO J 20: 128-136, 2001. Lodish H, Berk A, Zipursky SL, Matsudaira P, Baltimore D, Darnell J. Section 20.4, receptor tyrosine kinases. In: Molecular Cell Biology. New York, NY: W. H. Freeman, 2000. Lopez-Casillas F, Wrana JL, Massague J. Betaglycan presents ligand to the TGF beta signaling receptor. Cell 73: 1435-1444, 1993. Lundberg AS, Weinberg RA. Functional inactivation of the retinoblastoma protein requires sequential modification by at least two distinct cyclin-cdk complexes. Mol Cell Biol 18: 753-761, 1998. Lutz R, Sakai T, Chiquet M. Pericellular fibronectin is required for RhoA-dependent responses to cyclic strain in fibroblasts. J Cell Sci 123: 1511-1521, 2010. Ma Y, Chiao YA, Zhang J, Manicone AM, Jin YF, Lindsey ML. Matrix metalloproteinase-28 deletion amplifies inflammatory and extracellular matrix responses to cardiac aging. Microsc Microanal 18: 81-90, 2012. Ma Y, Zhang X, Bao H, Mi S, Cai W, Yan H, Wang Q, Wang Z, Yan J, Fan GC, Lindsey ML, Hu Z. Toll-like receptor (TLR) 2 and TLR4 differentially regulate doxorubicin induced cardiomyopathy in mice. PloS One 7: e40763, 2012. Mack CP, Somlyo AV, Hautmann M, Somlyo AP, Owens GK. Smooth muscle differentiation marker gene expression is regulated by RhoAmediated actin polymerization. J Biol Chem 276: 341-347, 2001. Maeda S, Dean DD, Gomez R, Schwartz Z, Boyan BD. The first stage of transforming growth factor beta1 activation is release of the large latent complex from the extracellular matrix of growth plate chondrocytes by matrix vesicle stromelysin-1 (MMP-3). Calcif Tissue Int 70: 54-65, 2002. Maekawa M, Ishizaki T, Boku S, Watanabe N, Fujita A, Iwamatsu A, Obinata T, Ohashi K, Mizuno K, Narumiya S. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science 285: 895-898, 1999. Maier S, Lutz R, Gelman L, Sarasa-Renedo A, Schenk S, Grashoff C, Chiquet M. Tenascin-C induction by cyclic strain requires integrinlinked kinase. Biochim Biophys Acta 1783: 1150-1162, 2008. Makino N, Sugano M, Satoh S, Oyama J, Maeda T. Peroxisome proliferator-activated receptor-gamma ligands attenuate brain natriuretic peptide production and affect remodeling in cardiac fibroblasts in reoxygenation after hypoxia. Cell Biochem Biophys 44: 65-71, 2006. Manabe I, Shindo T, Nagai R. Gene expression in fibroblasts and fibrosis: Involvement in cardiac hypertrophy. Circ Res 91: 1103-1113, 2002. Manabe R, Ohe N, Maeda T, Fukuda T, Sekiguchi K. Modulation of cell-adhesive activity of fibronectin by the alternatively spliced EDA segment. J Cell Biol 139: 295-307, 1997. Manickam N, Patel M, Griendling KK, Gorin Y, Barnes JL. RhoA/Rho kinase mediates TGF-beta1-induced kidney myofibroblast activation through Poldip2/Nox4-derived reactive oxygen species. Am J Physiol Renal Physiol 307: F159-171, 2014. Mann DL, Spinale FG. Activation of matrix metalloproteinases in the failing human heart: Breaking the tie that binds. Circulation 98: 16991702, 1998. Manser E, Loo TH, Koh CG, Zhao ZS, Chen XQ, Tan L, Tan I, Leung T, Lim L. PAK kinases are directly coupled to the PIX family of nucleotide exchange factors. Mol Cell 1: 183-192, 1998. Manso AM, Elsherif L, Kang SM, Ross RS. Integrins, membranetype matrix metalloproteinases and ADAMs: Potential implications for cardiac remodeling. Cardiovasc Res 69: 574-584, 2006. Manso AM, Kang SM, Ross RS. Integrins, focal adhesions, and cardiac fibroblasts. J Invest Med 57: 856-860, 2009. Martin MM, Buckenberger JA, Jiang J, Malana GE, Knoell DL, Feldman DS, Elton TS. TGF-beta1 stimulates human AT1 receptor expression in lung fibroblasts by cross talk between the Smad, p38 MAPK, JNK, and PI3K signaling pathways. Am J Physiol Lung Cell Mol Physiol 293: L790-799, 2007. Martinon F, Mayor A, Tschopp J. The inflammasomes: Guardians of the body. Annu Rev Immunol 27: 229-265, 2009. Martos R, Baugh J, Ledwidge M, O’Loughlin C, Conlon C, Patle A, Donnelly SC, McDonald K. Diastolic heart failure: Evidence of increased myocardial collagen turnover linked to diastolic dysfunction. Circulation 115: 888-895, 2007. Masumoto J, Taniguchi S, Ayukawa K, Sarvotham H, Kishino T, Niikawa N, Hidaka E, Katsuyama T, Higuchi T, Sagara J. ASC, a novel 22-kDa protein, aggregates during apoptosis of human promyelocytic leukemia HL-60 cells. J Biol Chem 274: 33835-33838, 1999. Mayorga M, Bahi N, Ballester M, Comella JX, Sanchis D. Bcl-2 is a key factor for cardiac fibroblast resistance to programmed cell death. J Biol Chem 279: 34882-34889, 2004. McClain PE. Characterization of cardiac muscle collagen. Molecular heterogeneity. J Biol Chem 249: 2303-2311, 1974.

Volume 5, April 2015

February 12, 2015 10:56 8in×10.75in

Intracellular Signaling of Cardiac Fibroblasts

322. Meerson FZ. A mechanism of hypertrophy and wear of the myocardium. Am J Cardiol 15: 755-760, 1965. 323. Meloche S, Pouyssegur J. The ERK1/2 mitogen-activated protein kinase pathway as a master regulator of the G1- to S-phase transition. Oncogene 26: 3227-3239, 2007. 324. Meran S, Thomas D, Stephens P, Martin J, Bowen T, Phillips A, Steadman R. Involvement of hyaluronan in regulation of fibroblast phenotype. J Biol Chem 282: 25687-25697, 2007. 325. Miki H, Miura K, Takenawa T. N-WASP, a novel actin-depolymerizing protein, regulates the cortical cytoskeletal rearrangement in a PIP2dependent manner downstream of tyrosine kinases. EMBO J 15: 53265335, 1996. 326. Miki H, Suetsugu S, Takenawa T. WAVE, a novel WASP-family protein involved in actin reorganization induced by Rac. EMBO J 17: 69326941, 1998. 327. Miller NL, Lawson C, Chen XL, Lim ST, Schlaepfer DD. Rgnef (p190RhoGEF) knockout inhibits RhoA activity, focal adhesion establishment, and cell motility downstream of integrins. PloS One 7: e37830, 2012. 328. Minamino T, Miyauchi H, Yoshida T, Tateno K, Kunieda T, Komuro I. Vascular cell senescence and vascular aging. J Mol Cell Cardiol 36: 175-183, 2004. 329. Miralles F, Posern G, Zaromytidou AI, Treisman R. Actin dynamics control SRF activity by regulation of its coactivator MAL. Cell 113: 329-342, 2003. 330. Mitchell MD, Laird RE, Brown RD, Long CS. IL-1beta stimulates rat cardiac fibroblast migration via MAP kinase pathways. Am J Physiol Heart Circ Physiol 292: H1139-1147, 2007. 331. Mitra SK, Hanson DA, Schlaepfer DD. Focal adhesion kinase: In command and control of cell motility. Nature Rev Mol Cell Biol 6: 56-68, 2005. 332. Miyazono K, Heldin CH. Role for carbohydrate structures in TGF-beta 1 latency. Nature 338: 158-160, 1989. 333. Mochitate K, Pawelek P, Grinnell F. Stress relaxation of contracted collagen gels: Disruption of actin filament bundles, release of cell surface fibronectin, and down-regulation of DNA and protein synthesis. Exp Cell Res 193: 198-207, 1991. 334. Mogilner A, Oster G. Polymer motors: Pushing out the front and pulling up the back. Curr Biol 13: R721-733, 2003. 335. Mollmann H, Nef HM, Kostin S, von Kalle C, Pilz I, Weber M, Schaper J, Hamm CW, Elsasser A. Bone marrow-derived cells contribute to infarct remodelling. Cardiovasc Res 71: 661-671, 2006. 336. Montanez E, Ussar S, Schifferer M, Bosl M, Zent R, Moser M, Fassler R. Kindlin-2 controls bidirectional signaling of integrins. Genes Dev 22: 1325-1330, 2008. 337. Mookerjee I, Unemori EN, Du XJ, Tregear GW, Samuel CS. Relaxin modulates fibroblast function, collagen production, and matrix metalloproteinase-2 expression by cardiac fibroblasts. Ann NY Acad Sci 1041: 190-193, 2005. 338. Moore L, Fan D, Basu R, Kandalam V, Kassiri Z. Tissue inhibitor of metalloproteinases (TIMPs) in heart failure. Heart Fail Rev 17: 693706, 2012. 339. Morgan DO. Cyclin-dependent kinases: Engines, clocks, and microprocessors. Annu Rev Cell Dev Biol 13: 261-291, 1997. 340. Morgan DO. Principles of CDK regulation. Nature 374: 131-134, 1995. 341. Morgan DO, Fisher RP, Espinoza FH, Farrell A, Nourse J, Chamberlin H, Jin P. Control of eukaryotic cell cycle progression by phosphorylation of cyclin-dependent kinases. Cancer J Sci Am 4(Suppl 1): S77-83, 1998. 342. Morton DP. Heart rate responses and fluid balance of competitive crosscountry hang gliding pilots. Int J Sports Physiol Perform 5: 55-63, 2010. 343. Mulder KM, Morris SL. Activation of p21ras by transforming growth factor beta in epithelial cells. J Biol Chem 267: 5029-5031, 1992. 344. Muranyi A, Derkach D, Erdodi F, Kiss A, Ito M, Hartshorne DJ. Phosphorylation of Thr695 and Thr850 on the myosin phosphatase target subunit: Inhibitory effects and occurrence in A7r5 cells. FEBS Lett 579: 6611-6615, 2005. 345. Murayama R, Kobayashi M, Takeshita A, Yasui T, Yamamoto M. MAPKs, activator protein-1 and nuclear factor-kappaB mediate production of interleukin-1beta-stimulated cytokines, prostaglandin E(2) and MMP-1 in human periodontal ligament cells. J Periodont Res 46: 568-575, 2011. 346. Murphy-Ullrich JE, Poczatek M. Activation of latent TGF-beta by thrombospondin-1: Mechanisms and physiology. Cytokine Growth Factor Rev 11: 59-69, 2000. 347. Nagahama KY, Togo S, Holz O, Magnussen H, Liu X, Seyama K, Takahashi K, Rennard SI. Oncostatin M modulates fibroblast function via signal transducers and activators of transcription proteins-3. Am J Resp Cell Mol Biol 49: 582-591, 2013. 348. Nakamoto T, Sakai R, Ozawa K, Yazaki Y, Hirai H. Direct binding of C-terminal region of p130Cas to SH2 and SH3 domains of Src kinase. J Biol Chem 271: 8959-8965, 1996.

755

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

349. Nakao A, Afrakhte M, Moren A, Nakayama T, Christian JL, Heuchel R, Itoh S, Kawabata M, Heldin NE, Heldin CH, ten Dijke P. Identification of Smad7, a TGFbeta-inducible antagonist of TGF-beta signalling. Nature 389: 631-635, 1997. 350. Nakao A, Imamura T, Souchelnytskyi S, Kawabata M, Ishisaki A, Oeda E, Tamaki K, Hanai J, Heldin CH, Miyazono K, ten Dijke P. TGFbeta receptor-mediated signalling through Smad2, Smad3 and Smad4. EMBO J 16: 5353-5362, 1997. 351. Nayal A, Webb DJ, Horwitz AF. Talin: An emerging focal point of adhesion dynamics. Curr Opin Cell Biol 16: 94-98, 2004. 352. Nevins JR, DeGregori J, Jakoi L, Leone G. Functional analysis of E2F transcription factor. Method Enzymol 283: 205-219, 1997. 353. Nevins JR, Leone G, DeGregori J, Jakoi L. Role of the Rb/E2F pathway in cell growth control. J Cell Physiol 173: 233-236, 1997. 354. Nigg EA. Mitotic kinases as regulators of cell division and its checkpoints. Nature Rev Mol Cell Biol 2: 21-32, 2001. 355. Nimnual AS, Taylor LJ, Bar-Sagi D. Redox-dependent downregulation of Rho by Rac. Nat Cell Biol 5: 236-241, 2003. 356. Nishida M, Onohara N, Sato Y, Suda R, Ogushi M, Tanabe S, Inoue R, Mori Y, Kurose H. Galpha12/13-mediated up-regulation of TRPC6 negatively regulates endothelin-1-induced cardiac myofibroblast formation and collagen synthesis through nuclear factor of activated T cells activation. J Biol Chem 282: 23117-23128, 2007. 357. Nobes CD, Hall A. Rho GTPases control polarity, protrusion, and adhesion during cell movement. J Cell Biol 144: 1235-1244, 1999. 358. Nobes CD, Hall A. Rho, rac, and cdc42 GTPases regulate the assembly of multimolecular focal complexes associated with actin stress fibers, lamellipodia, and filopodia. Cell 81: 53-62, 1995. 359. Numata T, Shimizu T, Okada Y. TRPM7 is a stretch- and swellingactivated cation channel involved in volume regulation in human epithelial cells. Am J Physiol-Cell Ph 292: C460-467, 2007. 360. O’Neill GM, Fashena SJ, Golemis EA. Integrin signalling: A new Cas(t) of characters enters the stage. Trends Cell Biol 10: 111-119, 2000. 361. Oakes PW, Beckham Y, Stricker J, Gardel ML. Tension is required but not sufficient for focal adhesion maturation without a stress fiber template. J Cell Biol 196: 363-374, 2012. 362. Ogunjimi AA, Briant DJ, Pece-Barbara N, Le Roy C, Di Guglielmo GM, Kavsak P, Rasmussen RK, Seet BT, Sicheri F, Wrana JL. Regulation of Smurf2 ubiquitin ligase activity by anchoring the E2 to the HECT domain. Mol Cell 19: 297-308, 2005. 363. Oliferenko S, Kaverina I, Small JV, Huber LA. Hyaluronic acid (HA) binding to CD44 activates Rac1 and induces lamellipodia outgrowth. J Cell Biol 148: 1159-1164, 2000. 364. Olson ER, Naugle JE, Zhang X, Bomser JA, Meszaros JG. Inhibition of cardiac fibroblast proliferation and myofibroblast differentiation by resveratrol. Am J Physiol Heart Circ Physiol 288: H1131-1138, 2005. 365. Olson ER, Shamhart PE, Naugle JE, Meszaros JG. Angiotensin IIinduced extracellular signal-regulated kinase 1/2 activation is mediated by protein kinase Cdelta and intracellular calcium in adult rat cardiac fibroblasts. Hypertension 51: 704-711, 2008. 366. Olson MF, Ashworth A, Hall A. An essential role for Rho, Rac, and Cdc42 GTPases in cell cycle progression through G1. Science 269: 1270-1272, 1995. 367. Onichtchouk D, Chen YG, Dosch R, Gawantka V, Delius H, Massague J, Niehrs C. Silencing of TGF-beta signalling by the pseudoreceptor BAMBI. Nature 401: 480-485, 1999. 368. Orlowski RZ, Small GW, Shi YY. Evidence that inhibition of p44/42 mitogen-activated protein kinase signaling is a factor in proteasome inhibitor-mediated apoptosis. J Biol Chem 277: 27864-27871, 2002. 369. Otey CA, Carpen O. Alpha-actinin revisited: A fresh look at an old player. Cell Motil Cytoskeleton 58: 104-111, 2004. 370. Oxley CL, Anthis NJ, Lowe ED, Vakonakis I, Campbell ID, Wegener KL. An integrin phosphorylation switch: The effect of beta3 integrin tail phosphorylation on Dok1 and talin binding. J Biol Chem 283: 5420-5426, 2008. 371. Padrick SB, Rosen MK. Physical mechanisms of signal integration by WASP family proteins. Annual review of biochemistry 79: 707-735, 2010. 372. Palazzo AF, Cook TA, Alberts AS, Gundersen GG. mDia mediates Rho-regulated formation and orientation of stable microtubules. Nat Cell Biol 3: 723-729, 2001. 373. Palazzo AF, Eng CH, Schlaepfer DD, Marcantonio EE, Gundersen GG. Localized stabilization of microtubules by integrin- and FAK-facilitated Rho signaling. Science 303: 836-839, 2004. 374. Palmer JN, Hartogensis WE, Patten M, Fortuin FD, Long CS. Interleukin-1 beta induces cardiac myocyte growth but inhibits cardiac fibroblast proliferation in culture. J Clin Invest 95: 2555-2564, 1995. 375. Pan CH, Wen CH, Lin CS. Interplay of angiotensin II and angiotensin(17) in the regulation of matrix metalloproteinases of human cardiocytes. Exp Physiol 93: 599-612, 2008. 376. Pantaloni D, Carlier MF. How profilin promotes actin filament assembly in the presence of thymosin beta 4. Cell 75: 1007-1014, 1993.

756

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

377. Parker KK, Ingber DE. Extracellular matrix, mechanotransduction and structural hierarchies in heart tissue engineering. Philos Trans R Soc Lond B, Biol Sci 362: 1267-1279, 2007. 378. Parsons JT. Focal adhesion kinase: The first ten years. J Cell Sci 116: 1409-1416, 2003. 379. 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. 380. Patel A, Sharif-Naeini R, Folgering JR, Bichet D, Duprat F, Honore E. Canonical TRP channels and mechanotransduction: From physiology to disease states. Pflugers Arch 460: 571-581, 2010. 381. Pedersen SF, Nilius B. Transient receptor potential channels in mechanosensing and cell volume regulation. Method Enzymol 428: 183-207, 2007. 382. Pelech SL, Charest DL, Mordret GP, Siow YL, Palaty C, Campbell D, Charlton L, Samiei M, Sanghera JS. Networking with mitogen-activated protein kinases. Mol Cell Biochem 127-128: 157-169, 1993. 383. Pelham RJ, Jr., Wang Y. Cell locomotion and focal adhesions are regulated by substrate flexibility. Proc Natl Acad Sci U S A 94: 13661-13665, 1997. 384. Pelham RJ, Jr., Wang Y. High resolution detection of mechanical forces exerted by locomoting fibroblasts on the substrate. Mol Biol Cell 10: 935-945, 1999. 385. Peterson LJ, Rajfur Z, Maddox AS, Freel CD, Chen Y, Edlund M, Otey C, Burridge K. Simultaneous stretching and contraction of stress fibers in vivo. Mol Biol Cell 15: 3497-3508, 2004. 386. Petit V, Thiery JP. Focal adhesions: Structure and dynamics. Biol Cell 92: 477-494, 2000 387. Petrich BG. Talin-dependent integrin signalling in vivo. Thrombosis and haemostasis 101: 1020-1024, 2009. 388. Petrov VV, Fagard RH, Lijnen PJ. Stimulation of collagen production by transforming growth factor-beta1 during differentiation of cardiac fibroblasts to myofibroblasts. Hypertension 39: 258-263, 2002. 389. Petrov VV, Fagard RH, Lijnen PJ. Transforming growth factor-beta(1) induces angiotensin-converting enzyme synthesis in rat cardiac fibroblasts during their differentiation to myofibroblasts. J Renin Angiotensin Aldosterone Syst 1: 342-352, 2000. 390. Phan SH. Biology of fibroblasts and myofibroblasts. Proc Am Thorac Soc 5: 334-337, 2008. 391. Pillai MS, Sapna S, Shivakumar K. p38 MAPK regulates G1-S transition in hypoxic cardiac fibroblasts. Int J Biochem Cell Biol 43: 919-927, 2011. 392. Pinto VI, Senini VW, Wang Y, Kazembe MP, McCulloch CA. Filamin A protects cells against force-induced apoptosis by stabilizing talinand vinculin-containing cell adhesions. FASEB J 28: 453-463, 2014. 393. Pipes GC, Creemers EE, Olson EN. The myocardin family of transcriptional coactivators: Versatile regulators of cell growth, migration, and myogenesis. Genes Dev 20: 1545-1556, 2006. 394. Ponti A, Machacek M, Gupton SL, Waterman-Storer CM, Danuser G. Two distinct actin networks drive the protrusion of migrating cells. Science 305: 1782-1786, 2004. 395. Porter KE, Turner NA. Cardiac fibroblasts: At the heart of myocardial remodeling. Pharmacol Ther 123: 255-278, 2009. 396. Porter KE, Turner NA, O’Regan DJ, Ball SG. Tumor necrosis factor alpha induces human atrial myofibroblast proliferation, invasion and MMP-9 secretion: Inhibition by simvastatin. Cardiovasc Res 64: 507515, 2004. 397. Powell SR, Wang P, Katzeff H, Shringarpure R, Teoh C, Khaliulin I, Das DK, Davies KJ, Schwalb H. Oxidized and ubiquitinated proteins may predict recovery of postischemic cardiac function: Essential role of the proteasome. Antioxid Redox Signal 7: 538-546, 2005. 398. Prager-Khoutorsky M, Lichtenstein A, Krishnan R, Rajendran K, Mayo A, Kam Z, Geiger B, Bershadsky AD. Fibroblast polarization is a matrix-rigidity-dependent process controlled by focal adhesion mechanosensing. Nat Cell Biol 13: 1457-1465, 2011. 399. Pramod S, Shivakumar K. Mechanisms in cardiac fibroblast growth: An obligate role for Skp2 and FOXO3a in ERK1/2 MAPK-dependent regulation of p27kip1. Am J Physiol Heart Circ Physiol 306: H844-855, 2014. 400. Qiu P, Feng XH, Li L. Interaction of Smad3 and SRF-associated complex mediates TGF-beta1 signals to regulate SM22 transcription during myofibroblast differentiation. J Mol Cell Cardiol 35: 1407-1420, 2003. 401. Raffetto JD, Khalil RA. Matrix metalloproteinases and their inhibitors in vascular remodeling and vascular disease. Biochem Pharmacol 75: 346-359, 2008. 402. Rajkumar VS, Howell K, Csiszar K, Denton CP, Black CM, Abraham DJ. Shared expression of phenotypic markers in systemic sclerosis indicates a convergence of pericytes and fibroblasts to a myofibroblast lineage in fibrosis. Arthritis Res Ther 7: R1113-1123, 2005. 403. Reddig PJ, Juliano RL. Clinging to life: Cell to matrix adhesion and cell survival. Cancer Metast Rev 24: 425-439, 2005.

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

404. Redondo J, Bishop JE, Wilkins MR. Effect of atrial natriuretic peptide and cyclic GMP phosphodiesterase inhibition on collagen synthesis by adult cardiac fibroblasts. Br J Pharmacol 124: 1455-1462, 1998. 405. Ren XD, Kiosses WB, Schwartz MA. Regulation of the small GTPbinding protein Rho by cell adhesion and the cytoskeleton. EMBO J 18: 578-585, 1999. 406. Ren XD, Wang R, Li Q, Kahek LA, Kaibuchi K, Clark RA. Disruption of Rho signal transduction upon cell detachment. J Cell Sci 117: 35113518, 2004. 407. Riddick N, Ohtani K, Surks HK. Targeting by myosin phosphataseRhoA interacting protein mediates RhoA/ROCK regulation of myosin phosphatase. J Cell Biochem 103: 1158-1170, 2008. 408. Ridley AJ, Hall A. Signal transduction pathways regulating Rhomediated stress fibre formation: Requirement for a tyrosine kinase. EMBO J 13: 2600-2610, 1994. 409. Ridley AJ, Hall A. The small GTP-binding protein rho regulates the assembly of focal adhesions and actin stress fibers in response to growth factors. Cell 70: 389-399, 1992. 410. Ridley AJ, Paterson HF, Johnston CL, Diekmann D, Hall A. The small GTP-binding protein rac regulates growth factor-induced membrane ruffling. Cell 70: 401-410, 1992. 411. Ridley AJ, Schwartz MA, Burridge K, Firtel RA, Ginsberg MH, Borisy G, Parsons JT, Horwitz AR. Cell migration: Integrating signals from front to back. Science 302: 1704-1709, 2003. 412. Riveline D, Zamir E, Balaban NQ, Schwarz US, Ishizaki T, Narumiya S, Kam Z, Geiger B, Bershadsky AD. Focal contacts as mechanosensors: Externally applied local mechanical force induces growth of focal contacts by an mDia1-dependent and ROCK-independent mechanism. J Cell Biol 153: 1175-1186, 2001. 413. Roca-Cusachs P, del Rio A, Puklin-Faucher E, Gauthier NC, Biais N, Sheetz MP. Integrin-dependent force transmission to the extracellular matrix by alpha-actinin triggers adhesion maturation. Proc Natl Acad Sci U S A 110: E1361-1370, 2013. 414. Rodriguez-Vita J, Sanchez-Lopez E, Esteban V, Ruperez M, Egido J, Ruiz-Ortega M. Angiotensin II activates the Smad pathway in vascular smooth muscle cells by a transforming growth factor-beta-independent mechanism. Circulation 111: 2509-2517, 2005. 415. Romer LH, Birukov KG, Garcia JG. Focal adhesions: Paradigm for a signaling nexus. Circ Res 98: 606-616, 2006. 416. Ronnov-Jessen L, Petersen OW. Induction of alpha-smooth muscle actin by transforming growth factor-beta 1 in quiescent human breast gland fibroblasts. Implications for myofibroblast generation in breast neoplasia. Lab Invest 68: 696-707, 1993. 417. Rosenkranz AC, Woods RL, Dusting GJ, Ritchie RH. Antihypertrophic actions of the natriuretic peptides in adult rat cardiomyocytes: Importance of cyclic GMP. Cardiovasc Res 57: 515-522, 2003. 418. Rottner K, Hall A, Small JV. Interplay between Rac and Rho in the control of substrate contact dynamics. Curr Biol 9: 640-648, 1999. 419. Sadoshima J, Izumo S. Signal transduction pathways of angiotensin II–induced c-fos gene expression in cardiac myocytes in vitro. Roles of phospholipid-derived second messengers. Circ Res 73: 424-438, 1993. 420. Saharinen J, Keski-Oja J. Specific sequence motif of 8-Cys repeats of TGF-beta binding proteins, LTBPs, creates a hydrophobic interaction surface for binding of small latent TGF-beta. Mol Biol Cell 11: 26912704, 2000. 421. Saltel F, Mortier E, Hytonen VP, Jacquier MC, Zimmermann P, Vogel V, Liu W, Wehrle-Haller B. New PI(4,5)P2- and membrane proximal integrin-binding motifs in the talin head control beta3-integrin clustering. J Cell Biol 187: 715-731, 2009. 422. Sanders MA, Basson MD. p130cas but not paxillin is essential for Caco-2 intestinal epithelial cell spreading and migration on collagen IV. J Biol Chem 280: 23516-23522, 2005. 423. Santamaria-Kisiel L, Rintala-Dempsey AC, Shaw GS. Calciumdependent and -independent interactions of the S100 protein family. Biochem J 396: 201-214, 2006. 424. Santander C, Brandan E. Betaglycan induces TGF-beta signaling in a ligand-independent manner, through activation of the p38 pathway. Cell Signal 18: 1482-1491, 2006. 425. Santiago JJ, Dangerfield AL, Rattan SG, Bathe KL, Cunnington RH, Raizman JE, Bedosky KM, Freed DH, Kardami E, Dixon IM. Cardiac fibroblast to myofibroblast differentiation in vivo and in vitro: Expression of focal adhesion components in neonatal and adult rat ventricular myofibroblasts. Dev Dyn 239: 1573-1584, 2010. 426. Sappino AP, Schurch W, Gabbiani G. Differentiation repertoire of fibroblastic cells: Expression of cytoskeletal proteins as marker of phenotypic modulations. Lab Invest 63: 144-161, 1990. 427. Sarasa-Renedo A, Tunc-Civelek V, Chiquet M. Role of RhoA/ROCKdependent actin contractility in the induction of tenascin-C by cyclic tensile strain. Exp Cell Res 312: 1361-1370, 2006. 428. Sato Y, Kataoka K, Matsumori A, Sasayama S, Yamada T, Ito H, Takatsu Y. Measuring serum aminoterminal type III procollagen peptide, 7S domain of type IV collagen, and cardiac troponin T in patients

Volume 5, April 2015

Intracellular Signaling of Cardiac Fibroblasts

429.

430.

431. 432.

433.

434.

435. 436.

437. 438.

439. 440.

441. 442.

443.

444.

445. 446. 447. 448.

449. 450. 451. 452. 453.

with idiopathic dilated cardiomyopathy and secondary cardiomyopathy. Heart 78: 505-508, 1997. Sauzeau V, Le Jeune H, Cario-Toumaniantz C, Smolenski A, Lohmann SM, Bertoglio J, Chardin P, Pacaud P, Loirand G. Cyclic GMPdependent protein kinase signaling pathway inhibits RhoA-induced Ca2+ sensitization of contraction in vascular smooth muscle. J Biol Chem 275: 21722-21729, 2000. Savvatis K, Pappritz K, Becher PM, Lindner D, Zietsch C, Volk HD, Westermann D, Schultheiss HP, Tschope C. Interleukin-23 deficiency leads to impaired wound healing and adverse prognosis after myocardial infarction. Circ Heart Fail 7: 161-171, 2014. Sawada Y, Tamada M, Dubin-Thaler BJ, Cherniavskaya O, Sakai R, Tanaka S, Sheetz MP. Force sensing by mechanical extension of the Src family kinase substrate p130Cas. Cell 127: 1015-1026, 2006. Sawai H, Okada Y, Funahashi H, Matsuo Y, Takahashi H, Takeyama H, Manabe T. Activation of focal adhesion kinase enhances the adhesion and invasion of pancreatic cancer cells via extracellular signal-regulated kinase-1/2 signaling pathway activation. Mol Cancer 4: 37, 2005. Saxena A, Chen W, Su Y, Rai V, Uche OU, Li N, Frangogiannis NG. IL-1 induces proinflammatory leukocyte infiltration and regulates fibroblast phenotype in the infarcted myocardium. J Immunol 191: 4838-4848, 2013. Schalla S, Bekkers SC, Dennert R, van Suylen RJ, Waltenberger J, Leiner T, Wildberger J, Crijns HJ, Heymans S. Replacement and reactive myocardial fibrosis in idiopathic dilated cardiomyopathy: Comparison of magnetic resonance imaging with right ventricular biopsy. Eur J Heart Fail 12: 227-231, 2010. Schaller MD. Biochemical signals and biological responses elicited by the focal adhesion kinase. Biochim Biophys Acta 1540: 1-21, 2001. Schaller MD, Hildebrand JD, Shannon JD, Fox JW, Vines RR, Parsons JT. Autophosphorylation of the focal adhesion kinase, pp125FAK, directs SH2-dependent binding of pp60src. Mol Cell Biol 14: 16801688, 1994. Schaller MD, Parsons JT. pp125FAK-dependent tyrosine phosphorylation of paxillin creates a high-affinity binding site for Crk. Mol Cell Biol 15: 2635-2645, 1995. Schlaepfer DD, Broome MA, Hunter T. Fibronectin-stimulated signaling from a focal adhesion kinase-c-Src complex: Involvement of the Grb2, p130cas, and Nck adaptor proteins. Mol Cell Biol 17: 1702-1713, 1997. Schlaepfer DD, Hauck CR, Sieg DJ. Signaling through focal adhesion kinase. Prog Biophys Mol Bio 71: 435-478, 1999. Schlaepfer DD, Jones KC, Hunter T. Multiple Grb2-mediated integrinstimulated signaling pathways to ERK2/mitogen-activated protein kinase: Summation of both c-Src- and focal adhesion kinase-initiated tyrosine phosphorylation events. Mol Cell Biol 18: 2571-2585, 1998. Schoppet M, Chavakis T, Al-Fakhri N, Kanse SM, Preissner KT. Molecular interactions and functional interference between vitronectin and transforming growth factor-beta. Lab Invest 82: 37-46, 2002. Schor SL, Ellis I, Dolman C, Banyard J, Humphries MJ, Mosher DF, Grey AM, Mould AP, Sottile J, Schor AM. Substratum-dependent stimulation of fibroblast migration by the gelatin-binding domain of fibronectin. J Cell Sci 109(Pt 10): 2581-2590, 1996. Schram K, Ganguly R, No EK, Fang X, Thong FS, Sweeney G. Regulation of MT1-MMP and MMP-2 by leptin in cardiac fibroblasts involves Rho/ROCK-dependent actin cytoskeletal reorganization and leads to enhanced cell migration. Endocrinology 152: 2037-2047, 2011. Schram K, Wong MM, Palanivel R, No EK, Dixon IM, Sweeney G. Increased expression and cell surface localization of MT1-MMP plays a role in stimulation of MMP-2 activity by leptin in neonatal rat cardiac myofibroblasts. J Mol Cell Cardiol 44: 874-881, 2008. Schwartz MA, DeSimone DW. Cell adhesion receptors in mechanotransduction. Curr Opin Cell Biol 20: 551-556, 2008. Sechi AS, Wehland J. ENA/VASP proteins: Multifunctional regulators of actin cytoskeleton dynamics. Front Biosci 9: 1294-1310, 2004. Serhan CN. Novel lipid mediators and resolution mechanisms in acute inflammation: To resolve or not? Am J Pathol 177: 1576-1591, 2010. Serini G, Bochaton-Piallat ML, Ropraz P, Geinoz A, Borsi L, Zardi L, Gabbiani G. The fibronectin domain ED-A is crucial for myofibroblastic phenotype induction by transforming growth factor-beta1. J Cell Biol 142: 873-881, 1998. Shamhart PE, Meszaros JG. Non-fibrillar collagens: Key mediators of post-infarction cardiac remodeling? J Mol Cell Cardiol 48: 530-537, 2010. Shao H, Wang JH, Pollak MR, Wells A. alpha-actinin-4 is essential for maintaining the spreading, motility and contractility of fibroblasts. PloS One 5: e13921, 2010. Sharma A, Mayer BJ. Phosphorylation of p130Cas initiates Rac activation and membrane ruffling. BMC Cell Biol 9: 50, 2008. Shattil SJ. Integrins and Src: Dynamic duo of adhesion signaling. Trends Cell Biol 15: 399-403, 2005. Shattil SJ, Kim C, Ginsberg MH. The final steps of integrin activation: The end game. Nature Rev Mol Cell Biol 11: 288-300, 2010.

757

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

454. Shen XM, Wu YP, Feng YB, Luo ML, Du XL, Zhang Y, Cai Y, Xu X, Han YL, Zhang X, Zhan QM, Wang MR. Interaction of MT1-MMP and laminin-5gamma2 chain correlates with metastasis and invasiveness in human esophageal squamous cell carcinoma. Clin Exp Metastas 24: 541-550, 2007. 455. Sherr CJ. Cancer cell cycles. Science 274: 1672-1677, 1996. 456. Sherr CJ, Roberts JM. CDK inhibitors: Positive and negative regulators of G1-phase progression. Genes Dev 13: 1501-1512, 1999. 457. Shi-wen X, Kennedy L, Renzoni EA, Bou-Gharios G, du Bois RM, Black CM, Denton CP, Abraham DJ, Leask A. Endothelin is a downstream mediator of profibrotic responses to transforming growth factor beta in human lung fibroblasts. Arthritis Rheum 56: 4189-4194, 2007. 458. Shi-wen X, Parapuram SK, Pala D, Chen Y, Carter DE, Eastwood M, Denton CP, Abraham DJ, Leask A. Requirement of transforming growth factor beta-activated kinase 1 for transforming growth factor beta-induced alpha-smooth muscle actin expression and extracellular matrix contraction in fibroblasts. Arthritis Rheum 60: 234-241, 2009. 459. Shifrin Y, Pinto VI, Hassanali A, Arora PD, McCulloch CA. Forceinduced apoptosis mediated by the Rac/Pak/p38 signalling pathway is regulated by filamin A. Biochem J 445: 57-67, 2012. 460. Siddesha JM, Valente AJ, Sakamuri SS, Yoshida T, Gardner JD, Somanna N, Takahashi C, Noda M, Chandrasekar B. Angiotensin II stimulates cardiac fibroblast migration via the differential regulation of matrixins and RECK. J Mol Cell Cardiol 65: 9-18, 2013. 461. Sieg DJ, Hauck CR, Ilic D, Klingbeil CK, Schaefer E, Damsky CH, Schlaepfer DD. FAK integrates growth-factor and integrin signals to promote cell migration. Nat Cell Biol 2: 249-256, 2000. 462. Sieg DJ, Hauck CR, Schlaepfer DD. Required role of focal adhesion kinase (FAK) for integrin-stimulated cell migration. J Cell Sci 112(Pt 16): 2677-2691, 1999. 463. Simm A, Nestler M, Hoppe V. Mitogenic effect of PDGF-AA on cardiac fibroblasts. Basic Res Cardiol 93(Suppl 3): 40-43, 1998. 464. Simm A, Nestler M, Hoppe V. PDGF-AA, a potent mitogen for cardiac fibroblasts from adult rats. J Mol Cell Cardiol 29: 357-368, 1997. 465. Simonson WT, Franco SJ, Huttenlocher A. Talin1 regulates TCRmediated LFA-1 function. J Immunol 177: 7707-7714, 2006. 466. Simpson DG, Majeski M, Borg TK, Terracio L. Regulation of cardiac myocyte protein turnover and myofibrillar structure in vitro by specific directions of stretch. Circ Res 85: e59-69, 1999. 467. Singh RB, Chohan PK, Dhalla NS, Netticadan T. The sarcoplasmic reticulum proteins are targets for calpain action in the ischemicreperfused heart. J Mol Cell Cardiol 37: 101-110, 2004. 468. Siwik DA, Chang DL, Colucci WS. Interleukin-1beta and tumor necrosis factor-alpha decrease collagen synthesis and increase matrix metalloproteinase activity in cardiac fibroblasts in vitro. Circ Res 86: 12591265, 2000. 469. Siwik DA, Pagano PJ, Colucci WS. Oxidative stress regulates collagen synthesis and matrix metalloproteinase activity in cardiac fibroblasts. Am J Physiol-Cell Ph 280: C53-60, 2001. 470. Small EM, Thatcher JE, Sutherland LB, Kinoshita H, Gerard RD, Richardson JA, Dimaio JM, Sadek H, Kuwahara K, Olson EN. Myocardin-related transcription factor-a controls myofibroblast activation and fibrosis in response to myocardial infarction. Circ Res 107: 294-304, 2010. 471. Small JV, Anderson K, Rottner K. Actin and the coordination of protrusion, attachment and retraction in cell crawling. Bioscience Rep 16: 351-368, 1996. 472. Songyang Z, Shoelson SE, Chaudhuri M, Gish G, Pawson T, Haser WG, King F, Roberts T, Ratnofsky S, Lechleider RJ, et al. SH2 domains recognize specific phosphopeptide sequences. Cell 72: 767-778, 1993. 473. Soranno T, Bell E. Cytostructural dynamics of spreading and translocating cells. J Cell Biol 95: 127-136, 1982. 474. Sottile J, Hocking DC. Fibronectin polymerization regulates the composition and stability of extracellular matrix fibrils and cell-matrix adhesions. Mol Biol Cell 13: 3546-3559, 2002. 475. Spinale FG. Myocardial matrix remodeling and the matrix metalloproteinases: Influence on cardiac form and function. Physiol Rev 87: 1285-1342, 2007. 476. Spinale FG, Coker ML, Heung LJ, Bond BR, Gunasinghe HR, Etoh T, Goldberg AT, Zellner JL, Crumbley AJ. A matrix metalloproteinase induction/activation system exists in the human left ventricular myocardium and is upregulated in heart failure. Circulation 102: 1944-1949, 2000. 477. Stacy LB, Yu Q, Horak K, Larson DF. Effect of angiotensin II on primary cardiac fibroblast matrix metalloproteinase activities. Perfusion 22: 51-55, 2007. 478. Stamos J, Lazarus RA, Yao X, Kirchhofer D, Wiesmann C. Crystal structure of the HGF beta-chain in complex with the Sema domain of the Met receptor. EMBO J 23: 2325-2335, 2004. 479. Staudinger LA, Spano SJ, Lee W, Coelho N, Rajshankar D, Bendeck MP, Moriarty T, McCulloch CA. Interactions between the discoidin domain receptor 1 and beta1 integrin regulate attachment to collagen. Biol Open 2: 1148-1159, 2013.

758

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

480. Stawowy P, Margeta C, Kallisch H, Seidah NG, Chretien M, Fleck E, Graf K. Regulation of matrix metalloproteinase MT1-MMP/MMP-2 in cardiac fibroblasts by TGF-beta1 involves furin-convertase. Cardiovasc Res 63: 87-97, 2004. 481. Steffen A, Ladwein M, Dimchev GA, Hein A, Schwenkmezger L, Arens S, Ladwein KI, Margit Holleboom J, Schur F, Victor Small J, Schwarz J, Gerhard R, Faix J, Stradal TE, Brakebusch C, Rottner K. Rac function is crucial for cell migration but is not required for spreading and focal adhesion formation. J Cell Sci 126: 4572-4588, 2013. 482. Steffensen B, Wallon UM, Overall CM. Extracellular matrix binding properties of recombinant fibronectin type II-like modules of human 72-kDa gelatinase/type IV collagenase. High affinity binding to native type I collagen but not native type IV collagen. J Biol Chem 270: 11555-11566, 1995. 483. Stillman B. Cell cycle control of DNA replication. Science 274: 16591664, 1996. 484. Stocks SZ, Taylor SM, Shiels IA. Transforming growth factor-beta1 induces alpha-smooth muscle actin expression and fibronectin synthesis in cultured human retinal pigment epithelial cells. Clin Experiment Ophthalmol 29: 33-37, 2001. 485. Streblow DN, Vomaske J, Smith P, Melnychuk R, Hall L, Pancheva D, Smit M, Casarosa P, Schlaepfer DD, Nelson JA. Human cytomegalovirus chemokine receptor US28-induced smooth muscle cell migration is mediated by focal adhesion kinase and Src. J Biol Chem 278: 50456-50465, 2003. 486. Sun Y, Weber KT. Infarct scar: A dynamic tissue. Cardiovasc Res 46: 250-256, 2000. 487. Suraneni P, Rubinstein B, Unruh JR, Durnin M, Hanein D, Li R. The Arp2/3 complex is required for lamellipodia extension and directional fibroblast cell migration. J Cell Biol 197: 239-251, 2012. 488. Suzuki H, Yagi K, Kondo M, Kato M, Miyazono K, Miyazawa K. cSki inhibits the TGF-beta signaling pathway through stabilization of inactive Smad complexes on Smad-binding elements. Oncogene 23: 5068-5076, 2004. 489. Tadokoro S, Shattil SJ, Eto K, Tai V, Liddington RC, de Pereda JM, Ginsberg MH, Calderwood DA. Talin binding to integrin beta tails: A final common step in integrin activation. Science 302: 103-106, 2003. 490. Taipale J, Miyazono K, Heldin CH, Keski-Oja J. Latent transforming growth factor-beta 1 associates to fibroblast extracellular matrix via latent TGF-beta binding protein. J Cell Biol 124: 171-181, 1994. 491. Takahra T, Smart DE, Oakley F, Mann DA. Induction of myofibroblast MMP-9 transcription in three-dimensional collagen I gel cultures: Regulation by NF-kappaB, AP-1 and Sp1. Int J Biochem Cell Biol 36: 353-363, 2004. 492. Takeda N, Manabe I, Uchino Y, Eguchi K, Matsumoto S, Nishimura S, Shindo T, Sano M, Otsu K, Snider P, Conway SJ, Nagai R. Cardiac fibroblasts are essential for the adaptive response of the murine heart to pressure overload. J Clin Invest 120: 254-265, 2010. 493. Takemura G, Ohno M, Hayakawa Y, Misao J, Kanoh M, Ohno A, Uno Y, Minatoguchi S, Fujiwara T, Fujiwara H. Role of apoptosis in the disappearance of infiltrated and proliferated interstitial cells after myocardial infarction. Circ Res 82: 1130-1138, 1998. 494. Takenawa T, Suetsugu S. The WASP-WAVE protein network: Connecting the membrane to the cytoskeleton. Nature Rev Mol Cell Biol 8: 37-48, 2007. 495. Tamada M, Sheetz MP, Sawada Y. Activation of a signaling cascade by cytoskeleton stretch. Dev Cell 7: 709-718, 2004. 496. Tamaki Y, Iwanaga Y, Niizuma S, Kawashima T, Kato T, Inuzuka Y, Horie T, Morooka H, Takase T, Akahashi Y, Kobuke K, Ono K, Shioi T, Sheikh SP, Ambartsumian N, Lukanidin E, Koshimizu TA, Miyazaki S, Kimura T. Metastasis-associated protein, S100A4 mediates cardiac fibrosis potentially through the modulation of p53 in cardiac fibroblasts. J Mol Cell Cardiol 57: 72-81, 2013. 497. ten Dijke P, Arthur HM. Extracellular control of TGFbeta signalling in vascular development and disease. Nature Rev Mol Cell Biol 8: 857-869, 2007. 498. ten Klooster JP, Jaffer ZM, Chernoff J, Hordijk PL. Targeting and activation of Rac1 are mediated by the exchange factor beta-Pix. J Cell Biol 172: 759-769, 2006. 499. Tetsu O, McCormick F. Beta-catenin regulates expression of cyclin D1 in colon carcinoma cells. Nature 398: 422-426, 1999. 500. Theriot JA. Accelerating on a treadmill: ADF/cofilin promotes rapid actin filament turnover in the dynamic cytoskeleton. J Cell Biol 136: 1165-1168, 1997. 501. Thievessen I, Thompson PM, Berlemont S, Plevock KM, Plotnikov SV, Zemljic-Harpf A, Ross RS, Davidson MW, Danuser G, Campbell SL, Waterman CM. Vinculin-actin interaction couples actin retrograde flow to focal adhesions, but is dispensable for focal adhesion growth. J Cell Biol 202: 163-177, 2013. 502. Thodeti CK, Paruchuri S, Meszaros JG. A TRP to cardiac fibroblast differentiation. Channels (Austin) 7: 211-214, 2013. 503. Thomas CV, Coker ML, Zellner JL, Handy JR, Crumbley AJ, III, Spinale FG. Increased matrix metalloproteinase activity and selective

Volume 5, April 2015

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

504.

505.

506. 507. 508. 509.

510. 511. 512.

513.

514.

515.

516.

517. 518. 519.

520.

521. 522. 523. 524. 525.

526.

upregulation in LV myocardium from patients with end-stage dilated cardiomyopathy. Circulation 97: 1708-1715, 1998. Thum T, Galuppo P, Wolf C, Fiedler J, Kneitz S, van Laake LW, Doevendans PA, Mummery CL, Borlak J, Haverich A, Gross C, Engelhardt S, Ertl G, Bauersachs J. MicroRNAs in the human heart: A clue to fetal gene reprogramming in heart failure. Circulation 116: 258-267, 2007. Thum T, Gross C, Fiedler J, Fischer T, Kissler S, Bussen M, Galuppo P, Just S, Rottbauer W, Frantz S, Castoldi M, Soutschek J, Koteliansky V, Rosenwald A, Basson MA, Licht JD, Pena JT, Rouhanifard SH, Muckenthaler MU, Tuschl T, Martin GR, Bauersachs J, Engelhardt S. MicroRNA-21 contributes to myocardial disease by stimulating MAP kinase signalling in fibroblasts. Nature 456: 980-984, 2008. Tokiwa G, Dikic I, Lev S, Schlessinger J. Activation of Pyk2 by stress signals and coupling with JNK signaling pathway. Science 273: 792794, 1996. Tomar A, Lim ST, Lim Y, Schlaepfer DD. A FAK-p120RasGAPp190RhoGAP complex regulates polarity in migrating cells. J Cell Sci 122: 1852-1862, 2009. Tomasek JJ, Gabbiani G, Hinz B, Chaponnier C, Brown RA. Myofibroblasts and mechano-regulation of connective tissue remodelling. Nature Rev Mol Cell Biol 3: 349-363, 2002. Tomasek JJ, Haaksma CJ, Eddy RJ, Vaughan MB. Fibroblast contraction occurs on release of tension in attached collagen lattices: Dependency on an organized actin cytoskeleton and serum. Anat Rec 232: 359-368, 1992. Toole BP. Hyaluronan is not just a goo! J Clin Invest 106: 335-336, 2000. Torsoni AS, Marin TM, Velloso LA, Franchini KG. RhoA/ROCK signaling is critical to FAK activation by cyclic stretch in cardiac myocytes. Am J Physiol Heart Circ Physiol 289: H1488-1496, 2005. Totsukawa G, Wu Y, Sasaki Y, Hartshorne DJ, Yamakita Y, Yamashiro S, Matsumura F. Distinct roles of MLCK and ROCK in the regulation of membrane protrusions and focal adhesion dynamics during cell migration of fibroblasts. J Cell Biol 164: 427-439, 2004. Totsukawa G, Yamakita Y, Yamashiro S, Hartshorne DJ, Sasaki Y, Matsumura F. Distinct roles of ROCK (Rho-kinase) and MLCK in spatial regulation of MLC phosphorylation for assembly of stress fibers and focal adhesions in 3T3 fibroblasts. J Cell Biol 150: 797-806, 2000. Trombetta-eSilva J, Eadie EP, Zhang Y, Norris RA, Borg TK, Bradshaw AD. The effects of age and the expression of SPARC on extracellular matrix production by cardiac fibroblasts in 3-D cultures. PloS One 8: e79715, 2013. Tsuji T, Ishizaki T, Okamoto M, Higashida C, Kimura K, Furuyashiki T, Arakawa Y, Birge RB, Nakamoto T, Hirai H, Narumiya S. ROCK and mDia1 antagonize in Rho-dependent Rac activation in Swiss 3T3 fibroblasts. J Cell Biol 157: 819-830, 2002. Tsuruda T, Boerrigter G, Huntley BK, Noser JA, Cataliotti A, CostelloBoerrigter LC, Chen HH, Burnett JC, Jr. Brain natriuretic Peptide is produced in cardiac fibroblasts and induces matrix metalloproteinases. Circ Res 91: 1127-1134, 2002. Tummalapalli CM, Heath BJ, Tyagi SC. Tissue inhibitor of metalloproteinase-4 instigates apoptosis in transformed cardiac fibroblasts. J Cell Biochem 80: 512-521, 2001. Turner CE. Paxillin interactions. J Cell Sci 113(Pt 23): 4139-4140, 2000. Turner CE, Brown MC, Perrotta JA, Riedy MC, Nikolopoulos SN, McDonald AR, Bagrodia S, Thomas S, Leventhal PS. Paxillin LD4 motif binds PAK and PIX through a novel 95-kD ankyrin repeat, ARFGAP protein: A role in cytoskeletal remodeling. J Cell Biol 145: 851863, 1999. Turner CE, Miller JT. Primary sequence of paxillin contains putative SH2 and SH3 domain binding motifs and multiple LIM domains: Identification of a vinculin and pp125Fak-binding region. J Cell Sci 107(Pt 6): 1583-1591, 1994. Tyagi SC, Campbell SE, Reddy HK, Tjahja E, Voelker DJ. Matrix metalloproteinase activity expression in infarcted, noninfarcted and dilated cardiomyopathic human hearts. Mol Cell Biochem 155: 13-21, 1996. Tyagi SC, Kumar S, Glover G. Induction of tissue inhibitor and matrix metalloproteinase by serum in human heart-derived fibroblast and endomyocardial endothelial cells. J Cell Biochem 58: 360-371, 1995. Tyagi SC, Kumar SG, Banks J, Fortson W. Co-expression of tissue inhibitor and matrix metalloproteinase in myocardium. J Mol Cell Cardiol 27: 2177-2189, 1995. Ulmer TS, Calderwood DA, Ginsberg MH, Campbell ID. Domainspecific interactions of talin with the membrane-proximal region of the integrin beta3 subunit. Biochemistry 42: 8307-8312, 2003. Valente AJ, Sakamuri SS, Siddesha JM, Yoshida T, Gardner JD, Prabhu R, Siebenlist U, Chandrasekar B. TRAF3IP2 mediates interleukin-18induced cardiac fibroblast migration and differentiation. Cell Signal 25: 2176-2184, 2013. Valente AJ, Yoshida T, Gardner JD, Somanna N, Delafontaine P, Chandrasekar B. Interleukin-17A stimulates cardiac fibroblast proliferation

Volume 5, April 2015

Intracellular Signaling of Cardiac Fibroblasts

527.

528.

529.

530. 531.

532. 533. 534.

535.

536. 537. 538. 539. 540. 541.

542.

543. 544. 545. 546. 547. 548. 549. 550. 551.

and migration via negative regulation of the dual-specificity phosphatase MKP-1/DUSP-1. Cell Signal 24: 560-568, 2012. Verrecchia F, Chu ML, Mauviel A. Identification of novel TGFbeta /Smad gene targets in dermal fibroblasts using a combined cDNA microarray/promoter transactivation approach. J Biol Chem 276: 17058-17062, 2001. Verreck FA, de Boer T, Langenberg DM, Hoeve MA, Kramer M, Vaisberg E, Kastelein R, Kolk A, de Waal-Malefyt R, Ottenhoff TH. Human IL-23-producing type 1 macrophages promote but IL-10-producing type 2 macrophages subvert immunity to (myco)bacteria. Proc Natl Acad Sci U S A 101: 4560-4565, 2004. Villar AV, Garcia R, Llano M, Cobo M, Merino D, Lantero A, Tramullas M, Hurle JM, Hurle MA, Nistal JF. BAMBI (BMP and activin membrane-bound inhibitor) protects the murine heart from pressure-overload biomechanical stress by restraining TGF-beta signaling. Biochim Biophys Acta 1832: 323-335, 2013. Visse R, Nagase H. Matrix metalloproteinases and tissue inhibitors of metalloproteinases: Structure, function, and biochemistry. Circ Res 92: 827-839, 2003. Vivar R, Humeres C, Varela M, Ayala P, Guzman N, Olmedo I, Catalan M, Boza P, Munoz C, Diaz Araya G. Cardiac fibroblast death by ischemia/reperfusion is partially inhibited by IGF-1 through both PI3K/Akt and MEK-ERK pathways. Exp Mol Pathol 93: 1-7, 2012. Vousden KH, Lane DP. p53 in health and disease. Nature Rev Mol Cell Biol 8: 275-283, 2007. Vuori K, Hirai H, Aizawa S, Ruoslahti E. Introduction of p130cas signaling complex formation upon integrin-mediated cell adhesion: A role for Src family kinases. Mol Cell Biol 16: 2606-2613, 1996. Walker GA, Masters KS, Shah DN, Anseth KS, Leinwand LA. Valvular myofibroblast activation by transforming growth factor-beta: Implications for pathological extracellular matrix remodeling in heart valve disease. Circ Res 95: 253-260, 2004. Wang D, Chang PS, Wang Z, Sutherland L, Richardson JA, Small E, Krieg PA, Olson EN. Activation of cardiac gene expression by myocardin, a transcriptional cofactor for serum response factor. Cell 105: 851-862, 2001. Wang GL, Jiang BH, Rue EA, Semenza GL. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci U S A 92: 5510-5514, 1995. Wang J, Zohar R, McCulloch CA. Multiple roles of alpha-smooth muscle actin in mechanotransduction. Exp Cell Res 312: 205-214, 2006. Wang LP, Wang Y, Zhao LM, Li GR, Deng XL. Angiotensin II upregulates K(Ca)3.1 channels and stimulates cell proliferation in rat cardiac fibroblasts. Biochem Pharmacol 85: 1486-1494, 2013. Wang P, Ballestrem C, Streuli CH. The C terminus of talin links integrins to cell cycle progression. J Cell Biol 195: 499-513, 2011. Wang X, Khaidakov M, Ding Z, Dai Y, Mercanti F, Mehta JL. LOX-1 in the maintenance of cytoskeleton and proliferation in senescent cardiac fibroblasts. J Mol Cell Cardiol 60: 184-190, 2013. Wang X, Zhang L, Wei Z, Zhang X, Gao Q, Ma Y, Liu X, Jiang Y, Guo C. The inhibitory action of PDCD4 in lipopolysaccharide/Dgalactosamine-induced acute liver injury. Lab Invest 93: 291-302, 2013. Wang Y, Xu F, Chen J, Shen X, Deng Y, Xu L, Yin J, Chen H, Teng F, Liu X, Wu W, Jiang B, Guo DA. Matrix metalloproteinase-9 induces cardiac fibroblast migration, collagen and cytokine secretion: Inhibition by salvianolic acid B from Salvia miltiorrhiza. Phytomedicine 19: 1319, 2011. Wang YS, Li SH, Guo J, Mihic A, Wu J, Sun L, Davis K, Weisel RD, Li RK. Role of miR-145 in cardiac myofibroblast differentiation. J Mol Cell Cardiol 66: 94-105, 2014. Watanabe N, Kato T, Fujita A, Ishizaki T, Narumiya S. Cooperation between mDia1 and ROCK in Rho-induced actin reorganization. Nat Cell Biol 1: 136-143, 1999. Webber J, Jenkins RH, Meran S, Phillips A, Steadman R. Modulation of TGFbeta1-dependent myofibroblast differentiation by hyaluronan. Am J Pathol 175: 148-160, 2009. Webber J, Meran S, Steadman R, Phillips A. Hyaluronan orchestrates transforming growth factor-beta1-dependent maintenance of myofibroblast phenotype. J Biol Chem 284: 9083-9092, 2009. Weber KT. Monitoring tissue repair and fibrosis from a distance. Circulation 96: 2488-2492, 1997. Weber KT, Brilla CG. Pathological hypertrophy and cardiac interstitium. Fibrosis and renin-angiotensin-aldosterone system. Circulation 83: 1849-1865, 1991. Weber KT, Janicki JS, Pick R, Capasso J, Anversa P. Myocardial fibrosis and pathologic hypertrophy in the rat with renovascular hypertension. Am J Cardiol 65: 1G-7G, 1990. Weber KT, Pick R, Jalil JE, Janicki JS, Carroll EP. Patterns of myocardial fibrosis. J Mol Cell Cardiol 21(Suppl 5): 121-131, 1989. Wegener KL, Partridge AW, Han J, Pickford AR, Liddington RC, Ginsberg MH, Campbell ID. Structural basis of integrin activation by talin. Cell 128: 171-182, 2007.

759

JWBT335-c140044

JWBT335-CompPhys-3G-v1

Printer: Yet to Come

Intracellular Signaling of Cardiac Fibroblasts

552. Weidemann A, Johnson RS. Biology of HIF-1alpha. Cell Death Differ 15: 621-627, 2008. 553. Weinberg RA. The retinoblastoma protein and cell cycle control. Cell 81: 323-330, 1995. 554. Wells RG, Discher DE. Matrix elasticity, cytoskeletal tension, and TGFbeta: The insoluble and soluble meet. Sci Signal 1: pe13, 2008. 555. Willems IE, Havenith MG, De Mey JG, Daemen MJ. The alpha-smooth muscle actin-positive cells in healing human myocardial scars. Am J Pathol 145: 868-875, 1994. 556. Wilson DP, Susnjar M, Kiss E, Sutherland C, Walsh MP. Thromboxane A2-induced contraction of rat caudal arterial smooth muscle involves activation of Ca2+ entry and Ca2+ sensitization: Rho-associated kinasemediated phosphorylation of MYPT1 at Thr-855, but not Thr-697. Biochem J 389: 763-774, 2005. 557. Wilson MA, Shukitt-Hale B, Kalt W, Ingram DK, Joseph JA, Wolkow CA. Blueberry polyphenols increase lifespan and thermotolerance in Caenorhabditis elegans. Aging Cell 5: 59-68, 2006. 558. Wittmann T, Waterman-Storer CM. Cell motility: Can Rho GTPases and microtubules point the way? J Cell Sci 114: 3795-3803, 2001. 559. Wong VW, Rustad KC, Akaishi S, Sorkin M, Glotzbach JP, Januszyk M, Nelson ER, Levi K, Paterno J, Vial IN, Kuang AA, Longaker MT, Gurtner GC. Focal adhesion kinase links mechanical force to skin fibrosis via inflammatory signaling. Nat Med 18: 148-152, 2012. 560. Worth DC, Hodivala-Dilke K, Robinson SD, King SJ, Morton PE, Gertler FB, Humphries MJ, Parsons M. Alpha v beta3 integrin spatially regulates VASP and RIAM to control adhesion dynamics and migration. J Cell Biol 189: 369-383, 2010. 561. Wozniak MA, Cheng CQ, Shen CJ, Gao L, Olarerin-George AO, Won KJ, Hogenesch JB, Chen CS. Adhesion regulates MAP kinase/ternary complex factor exchange to control a proliferative transcriptional switch. Curr Biol 22: 2017-2026, 2012. 562. Wozniak MA, Desai R, Solski PA, Der CJ, Keely PJ. ROCK-generated contractility regulates breast epithelial cell differentiation in response to the physical properties of a three-dimensional collagen matrix. J Cell Biol 163: 583-595, 2003. 563. Wrana JL. The secret life of Smad4. Cell 136: 13-14, 2009. 564. Wylie DE, Damsky CH, Buck CA. Studies on the function of cell surface glycoproteins. I. Use of antisera to surface membranes in the identification of membrane components relevant to cell-substrate adhesion. J Cell Biol 80: 385-402, 1979. 565. Xu D, Kishi H, Kawamichi H, Kajiya K, Takada Y, Kobayashi S. Sphingosylphosphorylcholine induces stress fiber formation via activation of Fyn-RhoA-ROCK signaling pathway in fibroblasts. Cell Signal 24: 282-289, 2012. 566. Xuan Nguyen TL, Choi JW, Lee SB, Ye K, Woo SD, Lee KH, Ahn JY. Akt phosphorylation is essential for nuclear translocation and retention in NGF-stimulated PC12 cells. Biochem Bioph Res Co 349: 789-798, 2006. 567. Yam CH, Fung TK, Poon RY. Cyclin A in cell cycle control and cancer. Cell Mol Life Sci 59: 1317-1326, 2002. 568. Yamamoto T, Ebisuya M, Ashida F, Okamoto K, Yonehara S, Nishida E. Continuous ERK activation downregulates antiproliferative genes throughout G1 phase to allow cell-cycle progression. Curr Biol 16: 1171-1182, 2006. 569. Yang B, He K, Zheng F, Wan L, Yu X, Wang X, Zhao D, Bai Y, Chu W, Sun Y, Lu Y. Over-expression of hypoxia-inducible factor-1 alpha in vitro protects the cardiac fibroblasts from hypoxia-induced apoptosis. J Cardiovasc Med (Hagerstown) 15: 579-586, 2014. 570. Yang F, Chung AC, Huang XR, Lan HY. Angiotensin II induces connective tissue growth factor and collagen I expression via transforming growth factor-beta-dependent and -independent Smad pathways: The role of Smad3. Hypertension 54: 877-884, 2009. 571. Yang F, Zhu XL, Wang LP, Song XD, Wang RM, Li ZG, Luo L, Hu WM, Ma WD, Pei X, Zhang LJ, Li QJ. Role of AcSDKP on collagen synthesis and degradation in cultured rat cardiac fibroblast. Zhonghua Xin Xue Guan Bing Za Zhi 34: 843-846, 2006. 572. Ye F, Hu G, Taylor D, Ratnikov B, Bobkov AA, McLean MA, Sligar SG, Taylor KA, Ginsberg MH. Recreation of the terminal events in physiological integrin activation. J Cell Biol 188: 157-173, 2010.

760

February 12, 2015 10:56 8in×10.75in

Comprehensive Physiology

573. Ye N, Verma D, Meng F, Davidson MW, Suffoletto K, Hua SZ. Direct observation of alpha-actinin tension and recruitment at focal adhesions during contact growth. Exp Cell Res 327: 57-67, 2014. 574. Yi X, Li X, Zhou Y, Ren S, Wan W, Feng G, Jiang X. Hepatocyte growth factor regulates the TGF-beta1-induced proliferation, differentiation and secretory function of cardiac fibroblasts. Int J Mol Med 34: 381390, 2014. 575. Yonemura S, Wada Y, Watanabe T, Nagafuchi A, Shibata M. alphaCatenin as a tension transducer that induces adherens junction development. Nat Cell Biol 12: 533-542, 2010. 576. Yu L, Hebert MC, Zhang YE. TGF-beta receptor-activated p38 MAP kinase mediates Smad-independent TGF-beta responses. EMBO J 21: 3749-3759, 2002. 577. Yu Q, Stamenkovic I. Cell surface-localized matrix metalloproteinase9 proteolytically activates TGF-beta and promotes tumor invasion and angiogenesis. Genes Dev 14: 163-176, 2000. 578. Yue L, Xie J, Nattel S. Molecular determinants of cardiac fibroblast electrical function and therapeutic implications for atrial fibrillation. Cardiovasc Res 89: 744-753, 2011. 579. Yue Z, Zhang Y, Xie J, Jiang J, Yue L. Transient receptor potential (TRP) channels and cardiac fibrosis. Curr Top Med Chem 13: 270-282, 2013. 580. Zaidel-Bar R, Ballestrem C, Kam Z, Geiger B. Early molecular events in the assembly of matrix adhesions at the leading edge of migrating cells. J Cell Sci 116: 4605-4613, 2003. 581. Zamilpa R, Lopez EF, Chiao YA, Dai Q, Escobar GP, Hakala K, Weintraub ST, Lindsey ML. Proteomic analysis identifies in vivo candidate matrix metalloproteinase-9 substrates in the left ventricle postmyocardial infarction. Proteomics 10: 2214-2223, 2010. 582. Zeisberg EM, Tarnavski O, Zeisberg M, Dorfman AL, McMullen JR, Gustafsson E, Chandraker A, Yuan X, Pu WT, Roberts AB, Neilson EG, Sayegh MH, Izumo S, Kalluri R. Endothelial-to-mesenchymal transition contributes to cardiac fibrosis. Nat Med 13: 952-961, 2007. 583. Zhang G, Moorhead PJ, el Nahas AM. Myofibroblasts and the progression of experimental glomerulonephritis. Exp Nephrol 3: 308-318, 1995. 584. Zhang HY, Gharaee-Kermani M, Zhang K, Karmiol S, Phan SH. Lung fibroblast alpha-smooth muscle actin expression and contractile phenotype in bleomycin-induced pulmonary fibrosis. Am J Pathol 148: 527-537, 1996. 585. Zhang X, Chattopadhyay A, Ji QS, Owen JD, Ruest PJ, Carpenter G, Hanks SK. Focal adhesion kinase promotes phospholipase C-gamma1 activity. Proc Natl Acad Sci U S A 96: 9021-9026, 1999. 586. Zhang X, Jiang G, Cai Y, Monkley SJ, Critchley DR, Sheetz MP. Talin depletion reveals independence of initial cell spreading from integrin activation and traction. Nat Cell Biol 10: 1062-1068, 2008. 587. Zhang X, Moore SW, Iskratsch T, Sheetz MP. N-WASP-directed actin polymerization activates Cas phosphorylation and lamellipodium spreading. J Cell Sci 127: 1394-1405, 2014. 588. Zhao XH, Laschinger C, Arora P, Szaszi K, Kapus A, McCulloch CA. Force activates smooth muscle alpha-actin promoter activity through the Rho signaling pathway. J Cell Sci 120: 1801-1809, 2007. 589. Zhao ZS, Manser E, Loo TH, Lim L. Coupling of PAK-interacting exchange factor PIX to GIT1 promotes focal complex disassembly. Mol Cell Biol 20: 6354-6363, 2000. 590. Zhu F, Li Y, Zhang J, Piao C, Liu T, Li HH, Du J. Senescent cardiac fibroblast is critical for cardiac fibrosis after myocardial infarction. PloS One 8: e74535, 2013. 591. Ziegler WH, Liddington RC, Critchley DR. The structure and regulation of vinculin. Trends Cell Biol 16: 453-460, 2006. 592. Zimmermann J, Brunner C, Enculescu M, Goegler M, Ehrlicher A, Kas J, Falcke M. Actin filament elasticity and retrograde flow shape the force-velocity relation of motile cells. Biophys J 102: 287-295, 2012. 593. Zou Y, Komuro I, Yamazaki T, Kudoh S, Aikawa R, Zhu W, Shiojima I, Hiroi Y, Tobe K, Kadowaki T, Yazaki Y. Cell type-specific angiotensin II-evoked signal transduction pathways: Critical roles of Gbetagamma subunit, Src family, and Ras in cardiac fibroblasts. Circ Res 82: 337-345, 1998.

Volume 5, April 2015

Intracellular signaling of cardiac fibroblasts.

Long regarded as a mere accessory cell for the cardiomyocyte, the cardiac fibroblast is now recognized as a critical determinant of cardiac function i...
4MB Sizes 0 Downloads 13 Views