161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene...

58
!"#$%% &' "()")(&(*+, -!-,+.)(+!-!+/-0+),+ //-0-1221

Transcript of 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene...

Page 1: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.
Page 2: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

Dissertation for the Degree of Doctor of Philosophy (Faculty of Medicine) in MolecularCell Biology presented at Uppsala University in 2002, from the Ludwig Institute forCancer Research

Abstract

Brodin, G. 2002 Smad7 in TGF-β Signalling. Acta Universitatis Upsaliensis.Comprehensive Summaries of Uppsala Dissertations from the Faculty of Medicine1135. 57pp. Uppsala. ISBN 91-554-5271-X.

Members of the transforming growth factor-β (TGF-β) superfamily of growth anddifferentiation factors regulate a vast array of biological functions in the adult, and areof great importance in governing cell fate determination and patterning in thedeveloping embryo. The TGF-β signal is propagated intracellularly by Smad proteinsresulting in transcriptional responses. Smad6 and Smad7 are inhibitory Smads known todownregulate the TGF-β signal and thereby possibly modulating the biologicalresponse. This thesis describes a functional analysis of the inhibitory Smad7 from an invitro and in vivo perspective.

The prostate gland is dependent on androgens for its growth and differentiation.Androgen withdrawal can cause regression and apoptosis in normal and malignantprostate. Previous studies suggest a role for TGF-β in the apoptotic mechanism. Weinvestigated the expression levels of Smad proteins in the rat ventral prostate as well asin an androgen sensitive prostate tumor model (Dunning R3327 PAP) byimmunohistochemistry. We observed an increased immunoreactivity for Smad3, Smad4and phosphorylated Smad2 in the rat ventral prostate epithelial cells after castration, aswell as in the prostate tumor cells. Expression of inhibitory Smad6 and Smad7 werealso increased in both normal and malignant prostate in response to castration.

Several studies have shown that Smad7 is upregulated in response to TGF-βstimuli, suggesting a role in a negative feedback loop attenuating the TGF-β response.We investigated the molecular mechanism behind that response by studying thetranscriptional regulation of the Smad7 gene. We identified a palindromic Smad bindingelement (SBE) in the promoter. Point mutations introduced into the SBE abolishedtranscriptional activation via TGF-β. We also observed that mutating or deletingbinding motifs for Sp1 and AP-1, led to an attenuation of the TGF-β mediatedtranscriptional induction as well as the basal promoter activity.

Gene ablation of Smad proteins has revealed specific physiological anddevelopmental roles. We analysed mice targeted on the Smad7 locus. The miceappeared viable and fertile with a slight reduction in litter size, suggesting a perinatalloss. Biochemical analysis of mouse embryonic fibroblasts (MEFs) showed no majordifference between wild type and mutant MEFs.

Greger Brodin, Ludwig Institute for Cancer Research, Biomedical Centre, Box 595, SE-751 24, Uppsala, Sweden

© Greger Brodin 2002

ISSN 0282-7476ISBN 91-554-5271-X

Printed in Sweden by Eklundshofs Grafiska, Uppsala 2002

Page 3: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

ToMy Parents

&Veronika

Page 4: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

4

This thesis is based on the following papers, referred to in the text by their Romannumerals:

I. Brodin, G., ten Dijke, P., Funa, K., Heldin, C., and Landström, M. (1999)Increased Smad expression and activation are associated with apoptosis innormal and malignant prostate after castration. Cancer Research 59, 2731-2738

II. Brodin, G., Åhgren, A., ten Dijke, P., Heldin, C., and Heuchel, R. (2000)Efficient TGF-β induction of the Smad7 gene requires co-operation betweenAP-1, Sp1, and Smad proteins on the mouse Smad7 promoter. J. Biol. Chem.275, 29023-29030

III. Brodin, G., Cheng, A., Åhgren, A., Kulkarni, S., Pawson, T., Heldin, C.H.,and Heuchel, R. Targeting of the mouse Smad7 gene. Manuscript.

Reprints were made with permission from the publishers

We shall not cease from explorationAnd the end of all our exploringWill be to arrive where we startedAnd know the place for the first time.

T.S. Eliot

Page 5: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

5

TABLE OF CONTENTS

PageAbbreviations 6Introduction 7Transforming growth factor-β (TGF-β) 8

TGF-β superfamily of ligands 8TGF-β 9Activin / Inhibin 9BMP 9Other members 10Control of TGF-β production and activation 11TGF-β superfamily binding proteins 11

TGF-β superfamily receptors 12Type I and II receptors 12Type III/Endoglin receptors 13

TGF-β superfamily intracellular signalling 13The Smad family of signal transducers 13The Smad pathway 15Non-Smad pathways 17Cross-talk between Smads and the MAP kinase pathway 18Specificity in Smad signalling 19Genetic targeting 20

Smad7 23Smad7 signalling 24Smad7 in physiological and pathological conditions 25

Vasculogenesis 25Lung morphogenesis 27Apoptosis 27Fibrosis 28Immunological disorders 28Carcinogenesis 29

Present investigations 30Smad expression in prostatic carcinoma (paper I) 30Transcriptional regulation of the Smad7 gene (paper II) 31Targeting of the Smad7 gene (paper III) 32

Future perspectives 34Acknowledgements 35References 37

Page 6: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

6

ABBREVIATIONS

ActR activin receptorALK activin receptor-like kinaseAMH/MIS anti-Müllerian hormone/Müllerian inhibiting

substanceATF2 activating transcription factor 2AVM arteriovenous malformationsBAMBI BMP and Activin membrane-bound inhibitorBMP bone morphogenetic proteinCBFA/AML core-binding factor A/acute myeloneus leukaemiaEMSA electrophoretic mobility shift assaysFAST forkhead activin signal transducerGDFs growth and differentiation factorsGDNF glial cell line-derived neurotrophic factorHAT histone acetyltransferaseHDAC histone deacetylasesJNK Jun N-terminal kinase LAP latency-associated proteinLEF1/TCF lymphoid enhancer-binding factor 1/T-cell specific

factorLLCs large latent complexesLTBP latent TGF-β-binding proteinMEF mouse embryonic fibroblastsMH1/2 Mad homology ½SAD Smad activation domainSARA Smad anchor for receptor activation SBE Smad binding elementSki Sloan-Kettering Institute proto-oncogeneSLC small latent complexSMURF Smad ubiquitylation regulatory factorSnoN Ski-related novel gene NTβR TGF-β receptorTAK1 TGF-β-activated kinaseTFE3 transcription factor binding to immunoglobulin

heavy constant mu enhancer 3TGF-β transforming growth factor-β TGIF TG3-intercating factorVSMC vascular smooth muscle cellsXIAP X-chromosome-linked inhibitor of apoptosis protein

Page 7: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

7

INTRODUCTION

We are all wired. The term “we” is used in a rather generous way referring to ourselvesand all the multicellular organisms sharing life with us. The fact that prevents us allfrom being a shapeless pile of individual cells is biochemical communication. This isachieved through an elaborate signalling network relaying a continuous flow ofinformation within cells and between the cells and their environment. Pathological ordysfunctional conditions may arise when the regulation of this information is perturbed,which ultimately can affect the development of an organism or the maintenance ofphysiological functions in the adult organism. Cells exchange information by means ofdirect cell to cell interactions, cell contact with extracellular matrix components or byresponding to soluble factors such as hormones, neurotransmitters, cytokines, growthfactors or small ions. The extracellular signal is converted into an intracellular one bythe use of a specific class of proteins, the so-called receptors. The initiated signalcascade will affect cellular functions such as survival, division, death, movement ordifferentiation.

Transforming growth factor-β (TGF-β) belongs to a superfamily of cytokines that exerta multitude of cellular functions on the organism affecting such diverse things as tissuehomeostasis, cell division and multicellular patterning and development. These diverseeffects are mediated through the combination and activation of specific type I and typeII serine/threonine receptor kinases that transduce the signal from the cellular surface tothe nucleus via downstream effector proteins, called Smads. These proteins can bedivided into three subclasses, receptor activated (R-Smads), common mediator Smad(Smad4) and the inhibitory Smads (Smad6 and Smad7). The R-Smads (Smad1, 2, 3, 5and 8) become phosphorylated in their C-termini and thus activated by the type Ireceptors. They associate with the common mediator Smad4 and subsequentlytranslocate to the nucleus where they trigger transcriptional responses. The inhibitorySmads exert their effect as attenuators and modulators of the TGF-β signal. Thequestion is how different Smads can influence such a vast array of cellular effects andhow pathological conditions such as fibrosis, rheumatoid arthritis and carcinogenesis,can arise in response to a deregulation of this TGF-β signal.

The scope of this thesis was to investigate the functional role of the inhibitory Smad7from an in vivo and an in vitro perspective. This was achieved by assessing the level ofSmad expression in a rat prostate carcinoma model, by characterising the transcriptionalregulation of the Smad7 promoter and finally, by disrupting the Smad7 gene in mice, inorder to identify phenotypic alterations that would suggest different functional roles forSmad7.

Page 8: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

8

TRANSFORMING GROWTH FACTOR-β (TGF-β)

The basis for the discovery of TGF-β was a finding by de Larco and Todaro, where theyobserved that a pool of polypeptide growth factors released from mouse fibroblaststransformed with murine sarcoma virus, was able to transform cells in a soft agar assay(De Larco and Todaro, 1978). Using a similar approach on normal rat kidney fibroblastsit was demonstrated that the growth factor cocktail actually consisted of two distinctpolypetide growth factors, coined transforming growth factor (TGF)- α and -β (Anzanoet al., 1983; Roberts et al., 1981). Whereas TGF-α displayed mitogenic activity, it laterbecame clear that TGF-β served as a potent growth inhibitor in most other cell types(Roberts and Sporn, 1990), rather suggesting a role as an important suppressor oftumourigenesis.

TGF-β superfamily of ligands

30 40 50 60 80 100% Identity

MIS/AMHTGF-βActivinNodal

BMP-3

Vgr-2

GDF-1

GDF-5

BMP-7

BMP-2/4

Dpp

Figure 1. The ligands of the TGF-β superfamily. Dendrogram indicating the relativelevel of amino acid sequence similarity between members.

Page 9: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

9

TGF-β serves as the prototype for the large and still growing TGF-β superfamily,consisting of more than 30 members which include bone morphogenetic proteins(BMPs), activins, inhibins, anti-müllerian hormone (AMH) and growth anddifferentiation factors (GDFs) (see Figure 1) (reviewed in Massague, 1998).

TGF-β TGF-β exists in three highly similar isoforms termed TGF-β1 (Derynck et al., 1985),TGF-β2 (de Martin et al., 1987; Madisen et al., 1988), and TGF-β3 (Derynck et al.,1988; ten Dijke et al., 1988), all encoded by distinct genes. Besides forminghomodimers, it has been reported that heterodimers can form between TGF-β1 andTGF-β2, and between TGF-β2 and TGF-β3 (Cheifetz et al., 1987; Ogawa et al., 1992).The functions of TGF-β ligands include cell cycle arrest in epithelial andhaematopoietic cells and control of cell proliferation and differentiation in mesenchymalcells. They are also strong inducers of extracellular matrix production and are involvedin wound healing and immunosuppression, (Massague, 1990; Roberts and Sporn, 1990;Roberts and Sporn, 1993). The three TGF-β isoforms affect TGF-β signalling in a rathersimilar and redundant way in vitro, but display different in vivo expression patterns andfunctions (reviewed in (Roberts and Sporn, 1992). Gene analysis also revealed that eachTGF-β isoform is controlled by a unique and differently regulated promoter (reviewedin Roberts et al., 1991).

Activin/InhibinMembers of the activin subfamily can induce pituitary follicle-stimulating hormone(FSH) production. Other functions include erythroid cell differentiation and induction ofmesoderm as shown in Xenopous. The different members of the activins can formhomo- or hetero-dimers between different β-subunits: βA, βB, βC and βE (Gaddy-Kurten et al., 1995; Harland, 1994; Vale et al., 1990).

Inhibin consists of a distantly related α-subunit, which can form hetero-dimers with thedifferent activin β-subunits. Inhibin can act as a counterplayer of activin, and inhibitFSH production as well as other effects of activin (Gaddy-Kurten et al., 1995; Vale etal., 1990).

BMPBone morphogenetic proteins (BMPs) constitute the largest group within the TGF-βsuperfamily of growth and differentiation factors, with over 20 members. These can befurther subdivided into different subfamilies based on structural homology andphysiological effects.

Spencer et al. observed in 1982 that flies mutant for the Drosophila genedecaplentaplegic (dpp) displayed a variety of pattern deficiencies and structureduplications. They suggested that that the dpp gene complex influenced positional

Page 10: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

10

information within developing epidermal tissue (Spencer et al., 1982). In a search forfactors able to induce ectopical bone and cartilage formation, Wozney et al. identifiedtwo mammalian homologues of dpp, BMP2 and BMP4 (Wozney et al., 1988). BMP4and dpp was later shown to be able to functionally substitute for each other in flies andmammals, suggesting a high degree of conservation between structure and function(Padgett et al., 1993; Sampath et al., 1993).

The BMP2 subfamily consists of BMP2, BMP4 and DPP (a BMP2/4 homologue inDrosophila). This subfamily affects gastrulation, neurogenesis, and interdigitalapoptosis in mammalians. In Xenopus the BMPs influence patterning of the mesodermand in Drosophila they affect dorsalization and eye and wing development (Harland,1994; Hogan, 1996; Mehler et al., 1997). The second subfamily, BMP5, includesmembers like BMP5, 60A (a BMP5 homologue in Drosophila), BMP6/Vgr1,BMP7/OP1, and BMP8/OP2. Together with BMP2 and BMP4 this subfamily isinvolved in the development of almost every organ and plays many roles in neuronaldevelopment (Hogan, 1996; Mehler et al., 1997).

The BMP5 subfamily members BMP5 (Celeste et al., 1990), BMP7/OP1 (Celeste et al.,1990; Özkaynak et al., 1990), and BMP8/OP2 (Özkaynak et al., 1992), were allidentified as bone-inducing proteins. In a search for mammalian homologues for Vg1,previously discovered in Xenopus oocytes (Weeks and Melton, 1987), BMP6/Vgr1 wasidentified (Lyons et al., 1989). Wharton et al. isolated the Drosophila gene 60A whilesearching for homologues of TGF-β (Wharton et al., 1991).

Other membersGrowth and differentiation factors (GDFs) belonging to the GDF5 subfamily have beenreported to affect chondrogenesis in the developing limbs (Hogan, 1996; Kingsley,1994) while Vg1, belonging to the Vg1 subfamily, affects axial mesoderm induction infrog and fish (Kingsley, 1994).

Members of the BMP3 subfamily, such as BMP3/osteogenein and GDF10 have beenreported to influence osteogenic differentiation, endochondral bone formation andmonocyte chemotaxis (Cunningham et al., 1992).

Nodal, one of the intermediate members has been shown to be involved in axialmesoderm induction and left-right asymmetry (Beddington, 1996; Hogan, 1996). Otherintermediate members, i.e. Dorsalin and GDF8, are involved in regulation of celldifferentiation within the neural tube (Basler et al., 1993) and inhibition of skeletalmuscle growth (McPherron et al., 1997), respectively.

More distant members of the TGF-β superfamily include anti-Müllerian hormone(AMH) also called Müllerian inhibiting substance (MIS) and glial cell line-derivedneurotrophic factor (GDNF). AMH induces Müllerian duct regression, resulting in malegenital organs (Cate et al., 1986; Josso et al., 1993). GDNF is a factor that promotesdopaminergic neuron survival and differentiation (Lin et al., 1993), but also affects

Page 11: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

11

kidney development (Massague, 1996). Surprisingly, GDNF has also been reported tosignal via the Ret tyrosine kinase receptor (Massague and Weis-Garcia, 1996).

Control of TGF-β production and activationTGF-βs are synthesised as inactive dimeric precursors that later become proteolyticallyprocessed in the Golgi apparatus by furin, a member of the convertase family ofendproteases (Dubois et al., 1995). Furin cleaves the precursor at RXXR sites, located112-114 amino acids from the C-terminal end resulting in a C-terminal TGF-β part andan N-terminal remnant called latency-associated protein (LAP) (Cui et al., 1998;Munger et al., 1997b). The N-terminal and the mature part of TGF-β, form a stillinactive, non-covalent complex called the small latent complex (SLC) which is muchmore stable than the bioactive form of TGF-β. The process continues in the Golgi by theformation of disulphide bonds between the LAP and latent TGF-β-binding protein(LTBP), resulting in large latent complexes (LLCs). LTBPs serve to enhance stabilityand secretion of the SLC complex, ensure correct folding of TGF-β, and target the latentTGF-β complex either to the cell surface for activation, or to the extracellular matrix ofdistinct cells and tissues for storage (reviewed in Munger et al., 1997b; Taipale andKeski-Oja, 1997). An additional function of the LTBPs might be to influence TGF-β toactivate integrin signalling. It has been shown that LTBP-1, -2, and -4 have RGDsequences which are integrin binding sites (Munger et al., 1998b; Saharinen et al.,1998). Indeed, it has been shown that large latent TGF-β complexes can associatedirectly to integrin αVβ1 at the cell surface (Munger et al., 1998b). In addition, it hasbeen shown that the epithelial specific integrin αVβ6 can bind the RGD-motif in LAP,suggesting a cytoskeleton-mediated activation of TGF-β (Munger et al., 1998a).

Final activation of TGF-β is physically controlled by the binding of LAP to mannose-6-phosphate receptors, and by proteases such as plasmin and cathepsin that cleave LAP(Munger et al., 1997a; Taipale and Keski-Oja, 1997). Thrombospondin-1 appears to beanother important activator of TGF-β in vivo. It induces a conformational change inLAP, which results in TGF-β activation (Crawford et al., 1998). Interestingly, TGF-βinduces PAI-1 (plasminogen activator inhibitor-1), suggesting some level of self-regulation of the activation process. Another activation candidate is matrixmetalloproteinase-9 (MMP-9) which can activate latent TGF-β2 and TGF-β3 in vitro(Yu and Stamenkovic, 2000).

TGF-β superfamily binding proteinsBioactive TGF-β superfamily members can associate with several extracellular proteinsthat can modify their activity. Extracellular matrix proteoglycans: decorin and biglycan,have been shown to inhibit TGF-β activity (Yamaguchi et al., 1990). In addition, a 60-kDa protein has been reported to inhibit the interaction between the receptor and TGF-βligands (Piek et al., 1997). Furthermore, it has been demonstrated that α2-macroglobulin acts as a clearance factor for circulating TGF-βs and activins in serum(Mather, 1996).

Page 12: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

12

BMPs are regulated by soluble factors that play important roles during embryonicdevelopment. By competing with BMPs for ligand-receptor interactions, noggin(Zimmerman et al., 1996), chordin (Piccolo et al., 1996), short gastrulation (Sog) aDrosophila homologue of chordin (Francois and Bier, 1995), Dan and gremlin (Hsu etal., 1998), modulate differentiation of mesoderm and ectoderm. Follistatin can alsoinhibit receptor-ligand interactions by interacting directly with activin (Nakamura et al.,1990), as well as with BMPs (Yamashita et al., 1995a). In addition, cerberus has beenidentified as an antagonist not only for BMP and activin, but also for nodal signalling(Hsu et al., 1998).

TGF-β superfamily receptors

Receptors for polypeptide growth factors are transmembrane proteins that are able totransduce the extracellular message across the plasma membrane into an intracellularsignal. The TGF-βs and related factors signal through a group of transmembrane proteinserine/threonine kinases known as the TGF-β receptor family.

Type I and II receptorsThe TGF-β receptor family is divided into two subfamilies, type I and type II receptors,based on their structural and functional characteristics. The vertebrate type I receptorsubfamily forms three subgroups based on similarities in kinase domains and signallingactivities. In mammals, one group includes TβR-I, ActR-IB, and ALK7, anotherincludes BMPR-IA and -IB, and a third group includes ALK1 and ALK2 (reviewed in(Massague, 1998)). As a result of being cloned simultaneously by different groups, mosttype I receptors received different names. Initially the neutral ALK (activin receptor-like kinase) nomenclature was used whereas a more descriptive name was designatedafter the identification of a physiological ligand. As a consequence, the TGF-β type Ireceptor initially identified as ALK-5 (Franzen et al., 1993) is now called TβR-I(Yamashita et al., 1994). On a similar note, the activin receptor type I previously knownas ALK-4 (ten Dijke et al., 1993) is now called ActR-IB (Carcamo et al., 1994), andALK-3 and ALK-6 are referred to as BMPR-IA and BMPR-IB, respectively (Koenig etal., 1994; Yamashita et al., 1995b). Type I receptors with no known ligands includemammalian ALK-7 (Rydén et al., 1996; Tsuchida et al., 1996) and the related XenopusXTrR-I (Mahony and Gurdon, 1995). TGF-β can bind to ALK-1 (Attisano et al., 1993)but more weakly than to TβR-I (ten Dijke et al., 1994a). ALK-1 has been reportedmediate a TGF-β response (Attisano et al., 1993). ALK-2 is also known as ActR-I sinceit can bind to activin and mediate certain activin responses (Attisano et al., 1993;Yamashita et al., 1995a). ActR-I is a bit promiscuous since it also can binds BMP2 andBMP4 (Liu et al., 1995; ten Dijke et al., 1994b). In addition, the ActR-I mousehomologue can when overexpressed bind TGF-β (Ebner et al., 1993; ten Dijke et al.,1994a). It has also been proposed that ALK-2/ActR-I can function as an MIS/AMHtype I receptor (He et al., 1993).

Page 13: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

13

The type II receptor subfamily includes in vertebrates TβR-II, BMPR-II, and AMHR,which bind selectively to TGF-β (Lin et al., 1992), BMPs (Liu et al., 1995; Nohno et al.,1995; Rosenzweig et al., 1995) and MIS, respectively (Baarends et al., 1994; diClemente et al., 1994). ActR-II and -IIB can together with the BMP type I receptorsbind BMP2, BMP4 and BMP7 as well as GDF5 (Hoodless et al., 1996; Nishitoh et al.,1996; Yamashita et al., 1995a). In addition, ActR-II and -IIB can also associate withactivins, either alone or in concert with activin type I receptors (Attisano et al., 1992;Mathews and Vale, 1991; Mathews et al., 1992).

The type I receptor contains a unique region, known as the GS-domain (Wrana et al.,1994). It has been shown that ligand-induced phosphorylation of TβR-I by the type IIreceptor takes place on serines and threonines in a TTSGSGSG sequence in the GS-domain. This is important for receptor activation and signalling (Souchelnytskyi et al.,1996; Wieser et al., 1995; Wrana et al., 1994).

Type III/Endoglin receptorsLigand crosslinking experiments have identified an additional class of accessoryreceptors, the type III receptors. These consists of two related proteins, betaglycan andendoglin (Gougos and Letarte, 1990; Wang et al., 1991). Since these receptors lackintrinsic signalling activity they are believed to regulate the TGF-β access to thesignalling receptors. Betaglycan has been reported to bind to all three TGF-β ligandswith high affinity (Cheifetz and Massagué, 1991; Segarini et al., 1989), and to facilitateligand binding to the type II receptor (Lopez-Casillas et al., 1993; Wang et al., 1991). Incontrast, endoglin has been shown to bind to TGF-β1 and -β3, but not to TGF-β2(Cheifetz et al., 1992).

TGF-β superfamily intracellular signalling

The signal initiated by the TGF-β superfamily ligands is transduced by type I and typeII serine/threonine kinase receptors into the intracellular space. The ligands bind type IIreceptor, forming a heterodimeric complex which can recruit and activate the type Ireceptor by phoshorylating serine and threonine residues located primarily in the GS-domain (Souchelnytskyi et al., 1996; Wrana et al., 1994). Ligands such as TGF-β1,TGF-β3, and activins have been shown to associate to the receptors through sequentialbinding (Massague, 1998). In contrast, TGF-β2 (Rodriguez et al., 1995), as well asBMP2 and BMP7, have affinity for both type I and II receptors, and associate with thereceptor complex through co-operative binding (Massague, 1998). The activatedreceptor can recruit downstream signalling molecules, known as Smads (figure 2).

The Smad family of signal transducersIn a genetic screen looking for enhancers of a weak decapentaplegic (dpp) maternalphenotype in Drosophila, a new gene mad (mothers against dpp) was isolated (Rafteryet al., 1995; Sekelsky et al., 1995). This was followed by the discovery of three Mad-

Page 14: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

14

homologues: sma-2, sma-3 and sma-4 in C. Elegans (Savage et al., 1996). Mutations ofthese sma genes resembled the small body sized phenotype observed in the Daf4mutants (type II serine/threonine receptor) of C. Elegans. The vertebrate homologues ofthe mad- and sma-genes are called smads. The proteins derived from these genes can bedivided into three different subclasses (figure 2), i) receptor activated Smads (R-Smads), ii) common mediator Smads (Co-Smads), and iii) inhibitory Smads (I-Smads)depending on their diverse roles in signalling. The N-terminal and C-terminal regions ofthe R-Smads and the Co-Smads display a great deal of homology and have beendesignated the Mad homology 1 (MH1) and Mad homology 2 (MH2) domainsrespectively. A proline-rich linker domain exists between the MH1 and MH2 domain.The R-Smads, but not the Co- or I-Smads, contain a C-terminal SSXS motif whichbecomes phosphorylated upon receptor interaction (reviewed in Heldin et al., 1997).

ActR-IIAActR-IIB

BMPR-IIActR-IIAActR-IIB

TGF-β Activin BMP

TβRII

TβRI/ALK-5 ALK-1

?

Smad1Smad5Smad8

ALK-4/ActR-IB

Smad2Smad3

ALK-2/ActR-IALK-3/BMPR-IAALK-6/BPMR-IB

Smad2Smad3

Smad1Smad5Smad8

Smad4 Smad4 Smad4 Smad4

Smad6Smad7

Smad6Smad7

Smad6Smad7

Smad6Smad7

TGF-β superfamily

Type-II R

Type-I R

R-Smads

Co-Smads

InhibitorySmads

Biologicalresponses

Inhibition of mitogenicityInduction of extracellular matrix

Induction of dorsal mesoderm, erythroid differentiation,and FSH-release

Induction of ventral mesoderm,cartilage and bone,and apoptosis

Figure 2. TGF-β superfamily members, their signalling molecules and some biologicalresponses. Adapted from Heldin et al., 1997.

The R-Smads Smad2 and Smad3 mediate signals from TGF-β and activin ligandsthrough the TβR-I/Alk-5 and ActR-IB receptors, respectively (see Figure 2) (Eppert etal., 1996; Macias-Silva et al., 1996; Zhang et al., 1996). BMP signalling is mediatedthrough R-Smads 1, 5 and 8 which become phosphorylated and activated by the ActR-I,BMPR-IA or BMPR-IB receptors (Thomsen, 1996). The determinant of specificity

Page 15: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

15

between the R-Smads and their interaction with either TGF-β/activin or BMP receptors,is the L3 loop region within the MH2 domain (Lo et al., 1998). However, some reportssuggest that Smad1, 5, and 8 might be promiscous towards the TGF-β receptors as well(Lux et al., 1999; Macias-Silva et al., 1998; Oh et al., 2000). The Co-Smad, Smad4(also known as DPC4, deleted in pancreatic carcinomas), appears to play a critical rolein both BMP- and TGF-β/activin-mediated pathways.

The affinity of Smad4 for R-Smads can be increased through phosphorylation of the C-terminal part of R-Smads leading to the formation of a complex (Souchelnytskyi et al.,1997). Moreover, Smad4 has a unique Smad activation domain (SAD) in the linkerregion, which governs transcriptional activation via the co-activator p300 (deCaestecker et al., 2000). One Co-Smad has been identified in mammals, but there mightbe others. Two Co-Smads have been found in Xenopus laevis (Howell et al., 1999;Masuyama et al., 1999).

The Smad pathwayUpon receptor activation Smad molecules can be recruited to the receptor complex(figure 3). It has been shown that microtubules play an important role in the guiding ofSmads to the plasma membrane (Dong et al., 2000), where a FYVE-domain protein,Smad anchor for receptor activation (SARA), presents the R-Smads to the receptor(Tsukazaki et al., 1998). The FYVE-domain is a protein structure that in other proteinshas been associated with anchoring to endosomes (Tsukazaki et al., 1998). This hasraised the possibility that the receptor needs to be internalised before it can bind theSARA-sequestered Smad proteins. Besides restraining the movement of Smads it seemsas if SARA also masks a region of Smad2 that otherwise mediates nuclear import (Xu etal., 2000). While receptor-mediated phosphorylation and activation of R-Smadsincreases the affinity for Smad4 it decreases the affinity for SARA at the same time (Xuet al., 2000).

The R-Smads are presented to the receptor complex and become activated byphosphorylation on the C-terminal SSXS motif (Heldin et al., 1997). The L45 loop, adistinct region in the type I receptor containing two β strands (number 4 and 5) flankingthe kinase domain, has been reported to determine the association of Smad molecules tothe receptor (Feng and Derynck, 1997). To underscore the importance of the L45 loopin determination of specificity, it has been shown that swapping of the L45 loop ofTβR-I and BMPR-IB also alters Smad1 and Smad2 recognition, subsequently leadingto switched transcriptional response (Chen et al., 1998; Persson et al., 1998). Afterreceptor activation, the R-Smads oligomerise with the Co-Smad, forming a heteromericcomplex that is translocated to the nucleus where it regulates transcription of targetgenes (Heldin et al., 1997). Smad3 and Smad4 have been shown to interact directly withspecific DNA sequences via their MH1 domain (Dennler et al., 1998; Vindevoghel etal., 1998; Yingling et al., 1997). In order to fully activate transcription of the targetpromoters, the Smad complexes must recruit additional factors, like the transcriptionfactor components AP-1 (Liberati et al., 1999), DNA-binding adaptors like FAST-1(Chen et al., 1996), or co-activators such as CBP/p300 (Feng et al., 1998; Janknecht et

Page 16: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

16

al., 1998; Nishihara et al., 1998; Shen et al., 1998; Topper et al., 1998). Thistransactivating role of the Smad proteins has been ascribed to the MH2 domain(reviewed by Massagué and Wotton, 2000).

R-SMAD

Co-SMAD

Nucleus

Ras

Erk

SNIP1

COREPRESSORS

TGIFSki/SnoN TGF-β

CO-ACTIVATORSP300CBP

SIGNALLINGRECEPTORS

ACCESSORY RECEPTORS

III

III

LIGANDS

DNA-BINDING COFACTORS

BetaglycanEndoglinCrypto

LEF1/TCFCBFA1CBFA3

JUN

β-cateninSMAD1SMAD2

JNK

BMPTGF-βTNF-αIFN-γEGF

SMAD1SMAD2NFκBSTAT

Co-SMAD

R-SMAD

CO

FAC

TO

R

BAMBISMURF1

SMAD7SMAD6

WntBMPTGF-βTNF-α

TGF-β superfamily

Figure 3. Schematic overview of the transforming growth factor-β (TGF-β) signallingpathway. Adapted from Massagué et al., 2000.

To date, two I-Smads have been identified in mammals, Smad6 and Smad7 (Imamura etal., 1997; Nakao et al., 1997; Topper et al., 1997). They have been characterised asinhibitors of TGF-β/activin and BMP signalling and have been proposed to function innegative feed-back loops, since their expression is induced by TGF-β/activin and BMP-members (Christian and Nakayama, 1999). It has been shown that I-Smads can interactstably with the type I receptor and block further activation of R-Smads (Imamura et al.,1997; Nakao et al., 1997; Souchelnytskyi et al., 1998). For Smad6 an additionalmechanism has been suggested, where Smad6 competes with Smad4 for binding toSmad1, thereby preventing the formation of a functional heteromeric Smad1/Smad4complex (Hata et al., 1998). Smad7 is considered as a general inhibitor of TGF-βsuperfamily-induced responses, whereas Smad6 is thought to preferentially block BMP-mediated signalling (Itoh et al., 1998), although this is controversial (Imamura et al.,1997).

Page 17: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

17

Self-regulation of the TGF-β pathway through negative feedback has been reported tooccur through other mechanisms than antagonistic I-Smads. BMP has been shown toexert a negative feedback control via the protein BAMBI (BMP and Activin membrane-bound inhibitor) which is a truncated and kinase-deficient type I receptor that interfereswith activation by binding to the BMP receptors (Onichtchouk et al., 1999).

Smads are disposed from the cytoplasm and the nucleus by ubiquitination andproteasomal degradation. The cytoplasmic Smad1 and Smad2 become targeted byproteolysis via the ubiquitin-ligase proteins Smurf1 and Smurf2, respectively (Lin et al.,2000; Zhu et al., 1999). The nuclear Smad2 on the other hand, is degraded via theUbcH5 family of ubiquitin-conjugating proteins (Lo and Massagué, 1999).

Non-Smad pathwaysTo this date, only Smads are recognised as TGF-β receptor substrates and signaltransducers. However, numerous reports indicate that TGF-β and BMP are also able totransduce signals via the mitogen-activated protein kinase (MAPK) pathway (figure 4)(reviewed by Massagué and Wotton, 2000) TGF-β has been reported to activate the JunN-terminal kinase (JNK) via MKK4 (MAPK kinase 4) in a rapid fashion in afibrosarcoma cell line (Hocevar et al., 1999). In addition, p38 seems to be activated byTGF-β in lung and kidney epithelial cell lines via MKK3 (Hanafusa et al., 1999) leadingto activation of the nuclear target ATF2 (activating transcription factor 2) (Sano et al.,1999). The biochemical link between the TGF-β receptor and the MKKs is to this datestill unknown. One candidate could be TAK1 (TGF-β activated kinase), a member ofthe MAPK kinase kinase (MAPKKK) family, implicated in activation of p38 viaMKK3/MKK6 (Moriguchi et al., 1996) and activation of JNKs via MKK4 (Shirakabe etal., 1997). JNK signalling by TAK1 is also reported during Drosophila development(Takatsu et al., 2000). Other factors such as TAB1 and XIAP might be located furtherupstream in the signalling pathway affecting TAK1.

TAB1 is one of the proposed activators of TAK1 (Shibuya et al., 1996). Also, sinceinjection of TAB1- and TAK1-mRNA into the dorsal marginal zone of Xenopusinduced ventral mesoderm, TAB1 and TAK1 are suggested to play a functional role inBMP signalling during early Xenopus development (Shibuya et al., 1998).

XIAP, a positive regulator of BMP signalling, has been proposed as a link between theBMP receptors and TAB1-TAK1. XIAP can associate not only with TAB1, but alsowith BMP-receptors in mammalian cells. Furthermore injection of XIAP mRNA intodorsal blastomeres enhanced the ventralisation of Xenopus embryos in a TAB1-TAK1dependent manner (Yamaguchi et al., 1999).

Page 18: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

18

GRB2

IIIPlasma membrane

TGF-β

TGF-β receptor

Growth factor

Growth factorreceptor

P P

SOS

Nuclear membrane

Ubiquitinationand

degradation

R-SMAD

Co-SMAD

R-SMAD

CO

FAC

TO

R

Fos Jun AFT2

TGIF

DNA

P

MKK4 MKK3

JNK p38

RAS MEK

P

XIAP

TAK1

Cytokinereceptor

Cytokine

ERK

Figure 4. Cross-talk between Smads and the MAP kinase pathway. Adapted fromMassagué et al., 2000.

Cross-talk between Smads and the MAP kinase pathwayThe Erk-activated Ras pathway has been reported to modify the TGF-β signallingpathways at different levels (figure 4). It has been shown that a hyperactive Raspathway can downregulate TGF-β receptors in H-Ras transformed rat intestinalepithelial cells (Zhao and Buick, 1995). Moreover, it has been reported that oncogenicRas can inhibit BMP and TGF-β signalling negatively by decreasing the accumulationof R-Smads in the nucleus. It was observed that EGF and HGF could phosphorylateSmad1, Smad2 and Smad3 on MAPK and Erk consensus sites located in the linkerregion, which led to a retention of R-Smads in the cytoplasm and reducedtranscriptional activity (Kretzschmar et al., 1997a; Kretzschmar et al., 1999;Kretzschmar et al., 1997b). When mutations were introduced into the Erkphosphorylation sites of Smad3, it resulted in a Ras-resistant form that could rescue thegrowth inhibitory response of TGF-β in Ras-transformed cells. Moreover, EGF induceda less extensive phosphorylation and cytoplasmic retention of Smad2 and Smad3, ascompared to oncogenic H-Ras, which might explain the silencing of anti-mitogenicTGF-β functions by hyperactive Ras in cancer cells (Kretzschmar et al., 1999).

The Erk/MAP kinase pathway has also been implicated in regulation of Smad mediatedtranscription at other levels than Smad retention. It has been reported that the

Page 19: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

19

transcriptional co-repressor TGIF becomes upregulated in response to Erk signalling(Lo et al., 2001).

Specificity in Smad signallingHow can specificity in Smad signal transduction be achieved? It has been shown thatthe different branches of TGF-β/activin and BMP signalling can access distinct sets oftarget genes, via their respective down-stream modulators such as Smad1 and Smad 2,(Chen et al., 1998). An example of the opposing and somewhat complementary rolesbetween Smad1- and Smad2-dependent pathways, is demonstrated in Xenopous laevisembryogenesis. Smad2 can, in response to Nodal related factors (Xnr-1, Xnr-2 andVg1), activate genes responsible for induction of dorsal mesoderm. In contrast Smad1,in response to BMP4, activates a different set of genes that affect induction of ventralmesoderm and suppression of neural fates. Ectopic expression of Smad1 or Smad2induces of ventral and dorsal structures, respectively (Baker and Harland, 1996; Graff etal., 1996). In a similar complementary fashion, BMP and Nodal dictate the specificationof left-right symmetry in the developing vertebrate embryo (Rodriguez Esteban et al.,1999; Saijoh et al., 2000; Yokouchi et al., 1999).

As transcription factors, R- and Co-Smad complexes are able to recognise and interactwith distinct DNA sequences, e.g. CAGAC, GTCTAGAC, and other GC richsequences. This interaction is rather weak which suggests that the Smads need theassistance of other proteins for correct binding (Shi et al., 1998). Since the pentamericCAGAC DNA-binding motif statistically occurs rather frequently in the genome, a highaffinity of the Smads to their recognition sequence would result in a rather general andunspecific association with DNA. Other factors are therefore required for increasedspecificity and target selection such as co-factors (figure 3). The co-factors arestructurally diverse proteins that share the ability to associate with Smad molecules anda neighbouring DNA sequence (reviewed (Massagué and Wotton, 2000)). Cell typespecific responses could be explained by the fact that certain co-factor combinations areexpressed in distinct cells or tissues (Chen et al., 1997; Hata et al., 2000).

One group of co-factors can be described as adaptors for the DNA-Smad interaction. Toachieve DNA binding Smad1 has been shown to bind to OAZ (olf-associated zincfinger) (Hata et al., 2000), while Smad2 binds FAST (forkhead activin signaltransducer) (Chen et al., 1996) and Mixer (Germain et al., 2000). The adaptor proteinsOAZ, FAST, and Mixer lack intrinsic transcriptional activity. The recognition betweenthe Smad molecule and the adaptor protein requires a bulging alpha-helix 2 region ofthe MH2 domain in the Smad proteins and a Smad-interaction domain, which isconserved in FAST and Mixer but not in OAZ (Chen et al., 1998; Germain et al., 2000;Hata et al., 2000).

Transcription factors are another group of proteins that can form functional complexeswith Smad molecules (figure 3). Examples are JunB (Zhang et al., 1998), TFE3(transcription factor binding to immunoglobulin heavy constant mu enhancer 3) (Hua etal., 1999), core-binding factor A/acute myeloneus leukaemia (CBFA/AML) proteins

Page 20: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

20

(Hanai et al., 1999; Pardali et al., 2000; Tsuji et al., 1998) and lymphoid enhancer-binding factor 1/T-cell specific factor (LEF1/TCF) (Labbe et al., 2000; Nishita et al.,2000). LEF1/TCF is the mediator of WNT/β-catenin signalling and is able to co-operatewith Smads in the activation of Xtwn (Xenopous twin) in response to Nodal relatedsignals (Nishita et al., 2000).

Smads are able to recruit transcriptional repressors as well as activators (figure 3).Smads can associate with repressors such as TG3-interacting factor (TGIF) (Wotton etal., 1999), Sloan-Kettering Institute proto-oncogene (Ski) (Luo et al., 1999) and Ski-related novel gene N (SnoN) (Sun et al., 1999). The repressors can bind histonedeacetylases (HDAC) which are generally implicated in chromatin condensation andtranscriptional silencing. The HDAC binding would counteract the effect of histoneacetyltransferase (HAT) activity associated with the co-activators (figure 3) CBP andp300, and transcriptional activation (Massagué and Wotton, 2000). The relative levelsof repressor versus co-activator could determine the final outcome of transcription. Forexample, it has been reported that the repressor TGIF can associate with Smad2 andSmad3 in competition with the co-activator p300 (Wotton et al., 1999).

Genetic targetingThe functional importance of TGF-β superfamily members in embryonic patterning andtissue homeostasis has been assessed by gene ablation experiments in mice (Table 1-3).They revealed a plethora of effects that range from defective vasculogenesis,inflammation, autoimmunity, and skeletal abnormalities to early malformationsaffecting egg cylinder formation, gastrulation and mesoderm formation. Interestingly,Smad3 deficient mice developed metastatic colorectal cancer at 4-6 months of age(Table 3) (Zhu et al., 1998), highlighting the potential role of TGF-β as a potent tumoursuppressor. This is a complex issue, since other groups have reported phenotypicalvariations for the same targeted genes, perhaps revealing the importance of geneticstrain background in determining the phenotypic outcome of specific null mutants.

Page 21: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

21

Table 1. Major defects in TGF-β superfamily ligand-deficient mice. Adapted fromGoumans and Mummery, 2000.

Targeted gene Phenotype ReferencesTGF-β1

TGF-β2

TGF-β3

- Defective yolk sac vasculogenesis andhaematopoiesis. Embryonic lethal (E9.5-11.5)

- Inflammation and autoimmunity

- Cardiac, lung, craniofacial, limb, spinalcolumn, eye, inner ear, urogenitaldefects. Perinatal lethality. - Cleft palate, delayed lung maturation.Perinatal lethality.

(Dickson et al., 1995)

(Shull et al., 1992)(Kulkarni et al., 1993)(Letterio and Roberts,1996)(Sanford et al., 1997)

(Proetzel et al., 1995)(Kaartinen et al., 1995)

BMP-2

BMP-4

BMP-7

- Embryonic lethal (E7.5-10.5)- Failure of proamniotic canal to close,heart malformation.- Embryonic lethal (E7.5-9.5)- 1. Arrest at egg cylinder stage, lack ofmesoderm.- 2. Develop until early somite stage w.disorganised/truncated posteriorstructures and reduced extra-embryonicmesoderm.- Defects in primordial germ cells andallantois formation.- Skull, eye and kidney defects.- Perinatal lethality

(Zhang and Bradley,1996)

(Winnier et al., 1995)

(Lawson et al., 1999)

(Dudley et al., 1995)(Luo et al., 1995)

Activin βA

Activin βB

Activin βA/βB

- Cleft pallet, lack of whiskers andincisors.- Failure of eyelid fusion. Females showimpaired reproductive ability.- No additional defects.

(Matzuk, 1995) (Schrewe et al., 1994)(Vassalli et al., 1994)(Matzuk, 1995)

Nodal - Embryonic lethal (E7.5). Failure ingastrulation and primitive streakformation

(Conlon et al., 1994)

Page 22: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

22

Table 2. Major defects in TGF-β superfamily receptor-deficient mice. Adapted fromGoumans and Mummery, 2000.

Targeted gene Phenotype ReferencesALK-1

ALK-2

ALK-3

ALK-4

ALK-5

ALK-6

- Embryonic lethal (E10.5-11.5).- Defects in angiogenesis and vascularsmooth muscle cell differentiation.- Embryonic lethal (E7.5-9.5)- Failure in primitive streak elongation,delayed mesoderm formation andmalformed visceral endoderm. - Embryonic lethal (E7.5-9.5)- Fail to form mesoderm, reducedproliferation of the epiblast.- Embryonic lethal (E7.5-9.5)- Defect in epiblast and extraembryonicectoderm organisation and gastrulation.- Embryonic lethal around E 10.5. - Severe defects in vascular developmentof placenta and yolk sac.- Intact haematopoietic potential ofprecursors.- Defects in endothelial cell proliferation,migration and fibronectin production.- Defects in digit formation and fore- andhindlimb development

(Oh et al., 2000)

(Gu et al., 1999)

(Mishina et al., 1995)

(Gu et al., 1998)

(Larsson et al., 2001)

(Yi et al., 2000)(Baur et al., 2000)

TβRII

ActR-IIA

ActR-IIB

ActR-IIA/ActR-IIb

- Embryonic lethal (E10)- Defective yolk sac vasculogenesis.- Skeletal and facial abnormalities inpercentage of mice.- Perinatal lethality- Cardiac defects associated with defectsin left-right asymmetry- Homeotic transformation of theskeleton - Embryonic lethal- Defect in primitive streak formationand gastrulation.

(Oshima et al., 1996)

(Matzuk, 1995)

(Oh and Li, 1997)

(Song et al., 1999)

Endoglin -Embryonic lethal (E10.5-11.5). -Defective yolk sac vasculogenesis,embryonic angiogenesis and vascularsmooth muscle cell development. - Cardiac malformations.

(Arthur et al., 1999)(Li et al., 1999)

(Bourdeau et al., 1999)

Page 23: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

23

Table 3. Major defects in Smad-deficient mice. Adapted from Goumans and Mummery,2000.

Targeted gene Phenotype ReferencesSmad1

Smad2

Smad3

Smad4

Smad5

Smad6

-Embryonic lethal (E9.5) - Failure in establishing chorion-allantoiccirculation. - Embryonic lethal (E7.5-8.5)- Failure in egg cylinder elongation,gastrulation and mesoderm formation.- Metastatic colorectal cancer (4-6months of age)- Impaired immunity and chronic infection.

- Accelerated wound healing.- Embryonic lethal (E7.5-8.5)- Growth retardation, no mesodermformation, abnormal visceral endoderm.- Embryonic lethal (E9.5-10.5)- Defect in angiogenesis, left/rightasymmetry, craniofascial abnormalitiesand induced mesenchymal apoptosis.- Cardiovascular abnormalities.- Defect in endocardial cushiontransformation

(Lechleider et al.,2001)

(Weinstein et al., 1998)(Nomura and Li, 1998)(Waldrip et al., 1998)(Zhu et al., 1998)

(Datto et al., 1999)(Yang et al., 1999b)

(Ashcroft et al., 1999)(Yang et al., 1998)(Sirard et al., 1998)

(Chang et al.,1999),(Chang et al.,2000)(Yang et al., 1999a)(Galvin et al., 2000)

SMAD7

Two research groups, using two different experimental approaches, discovered Smad7.In a screening of Smad homologues in mouse EST and human cDNA libraries, Nakao etal. identified a protein of 426 amino acids and termed it Smad7 (Nakao et al., 1997).Topper et al. using a differential display approach identified Smad6 and Smad7 as novelgenes upregulated in response to laminar vascular flow-stress in human endothelial cells(Topper et al., 1997).

Page 24: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

24

Smad7 signalling

Smad7 has been proposed to function as an antagonist of TGF-β signalling. Nakao et al.observed that Smad7 could block responses initiated by TGF-β1 in Mv1Lu- and HaCaTcells. They also demonstrated that injection of Smad7 mRNA in a Xenopous embryoblocked the activin/TGF-β signalling. Moreover, ectopic expression of Smad7 resultedin a failure to form head and tail structures and showed a negative effect on theformation of mesodermal derivatives such as muscle and notochord (Nakao et al.,1997). This mimicked the effect of a dominant negative form of the activin receptor(Hemmati-Brivanlou and Melton, 1992), or dominant negative Smad4 (Lagna et al.,1996). The expression of Brachyury, a mesodermal marker gene, was alsodownregulated by Smad7 injection in the blastomere of two-cell Xenopus embryos(Nakao et al., 1997). Taken together, these data underscored the possible function ofSmad7 as a possible attenuator of TGF-β/activin- mediated signalling, both in vivo andin vitro.

Based on the observations that TGF-β mediated phosphorylation of Smad2 and Smad3is blocked in Smad7-transfected cells and that overexpressed Smad7 associates with theTGF-β type I receptor in COS cells, a molecular mechanism was proposed stipulatingthat the inhibitory effect resides in the ability of Smad7 to compete with R-Smads forthe type I receptor (Hayashi et al., 1997; Nakao et al., 1997).

An additional Smad7 function was proposed by Kavsak et al. where Smad7 can targetthe TGF-β receptor for proteasomal and lysosomal degradation, through the interactionwith an ubiquitin ligase, Smurf2. In addition, they also observed that the associationbetween Smurf2 and Smad7 induced a translocation of Smurf2 from nucleus to theactivated TGF-β receptor. This Smurf2/Smad7-complex formation and TGF-β receptorturnover could be further enhanced by interferon-γ (IFN-γ) (Kavsak et al., 2000).

On a similar note, Ebisawa et al. observed that Smurf1, an E3 ubiquitin ligase for BMP-specific Smads, also could bind Smad7 and thereby induce TGF-β receptor degradation(Ebisawa et al., 2001). Moreover, it has been demonstrated that STRAP, a WD40 repeatprotein could associate with both TβR-I and TβR-II (Datta et al., 1998). This stabilisedthe interaction between Smad7 and the activated receptor, thereby assisting in Smad7inhibition of TGF-β signalling (Datta et al., 1998; Datta and Moses, 2000).

Based on observations that Smad7 mRNA expression could be induced by TGF-βstimulation, Smad7 has been proposed to take part in a negative feedback loopdownregulating the TGF-β signal (Nakao et al., 1997). Additional pathways have beenshown to upregulate Smad7 transcription. Bitzer et al. reported that the mRNAexpression of Smad7 was upregulated via the NF-κB/RelA pathway by the pro-inflammatory molecules tumour necrosis factor-α and interleukin 1β (Bitzer et al.,2000). In addition, it was demonstrated that IFN-γ also could upregulate Smad7transcription via Jak1/Stat1 (Ulloa et al., 1999).

Page 25: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

25

I-Smads have conserved C-terminal Mad homology 2 (MH2) domains, whereas theamino acid sequences of their N-terminal regions (N-domain) are highly divergent, notonly between the Smad families but also between the I-Smads of different species.(Christian and Nakayama, 1999; Nakayama et al., 2000). Both the N- and C-terminalMH2-domains are important to obtain full activity of Smad6 and Smad7 in Xenopus(Nakayama et al., 2001). The N-domain and MH2-domain can bind to each other, whichfacilitates the interaction between Smad7 and TGF-β receptors and thereby enhancesthe inhibitory effect of Smad7/MH2-complex (Hanyu et al., 2001).

Smad7 has been suggested to exert other functional roles distinct from the antagonisticeffect in receptor-mediated Smad activation. It has been reported that overexpressedSmad7 is predominantly located to the nucleus in the absence of ligand and becomestranslocated to the cytoplasm in response to TGF-β stimulation (Itoh et al., 1998).Moreover, Pulaski et al., observed that phosphorylation of Smad7 at serine residue 249could affect Smad7-dependent transcriptional activation on an SV40 minimal promoter(using the GAL4 DNA binding domain fused to Smad7) (Pulaski et al., 2001). Inaddition, it has been proposed that I-Smads can act as transcriptional co-repressors byrecruiting histone deacetylases (HDACs) (Bai and Cao, 2002; Bai et al., 2000).

Smad7 in physiological and pathological conditions

VasculogenesisThe developing vascular system of the early embryo originates from a small populationof mesodermal endothelial precursor cells. These cells join to form a capillary plexusthat is gradually remodelled by the development of distinct arterial-venous parts and therecruitment of mural cells such as capillary-associated pericytes and vascular smoothmuscle cells (VSMC). These structures constitute the basis for future angiogenesisusing existing vessels to form new ones (Daniel and Abrahamson, 2000; Hungerfordand Little, 1999; Risau, 1997; Yancopoulos et al., 2000). A multitude of TGF-βsuperfamily ligands, receptors and signalling molecules have been implicated in thedevelopment of the vascular system (Gatherer et al., 1990; Pelton et al., 1990; Pelton etal., 1989; Roberts and Sporn, 1992; Schmid et al., 1991). Several in vitro studiessuggest that TGF-β can regulate the growth of endothelial cells as well as influencetheir migration and fusion into capillary tubes. In addition, TGF-β can also affect vessellumen size (Gajdusek et al., 1993; Madri et al., 1988; Merwin et al., 1990; Pepper,1997; Pepper et al., 1990; Pepper et al., 1993; Roberts and Sporn, 1989). However, thebehaviour of the endothelial cell in response to TGF-β stimulation, varies greatlybetween experimental set ups, making functional assessments difficult (Daniel andAbrahamson, 2000; Klagsbrun and D'Amore, 1991). A different approach used toinvestigate the function of individual genes in vasculogenesis in vivo is gene targeting.Such studies have revealed that activin receptor-like kinase 1 (ALK1) a TGF-β/activinbinding receptor, endoglin (Eng) a TGF-β superfamily co-receptor, and Smad5 areexamples of proteins necessary for normal angiogenesis. Null mutants lacking either ofthese three gene functions display fragile, haemorrhagic and dilated vessels. In addition,

Page 26: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

26

extracellular matrix remodelling enzymes have abnormal expression levels and therecruitment of vascular smooth muscle cells (VSMC) to the vascular endothelium isimpaired (Chang et al., 1999; Li et al., 1999; Oh et al., 2000; Pepper, 1997; Urness etal., 2000; Yang et al., 1999b). Moreover, ALK1 null mutant mice display anarteriovenous malformation phenotype, characterised by the development of shuntsbetween the arterial and venous circulatory systems (Urness et al., 2000). Interestingly,these phenotypes resemble hereditary haemorrhagic telangiectasia, a humanpathological condition associated to haploinsufficiency of ALK1 or Eng (Shovlin andLetarte, 1999).

What is the role of TGF-β signalling in the developing vascular system? Recent modelssuggest that the primary function is to promote VSMC recruitment and differentiationwhile inhibiting endothelial proliferation through VSMC-endothelial interactions.However, non-TGF-β superfamily (e.g. angiopoeitins, PDGFs) signalling mutants alsoexhibit vessel dilations, fragility and haemorrhage associated with disturbed VSMC-endothelial interactions (Puri et al., 1995; Sato et al., 1995; Suri et al., 1996; Vikkula etal., 1996). It is therefore not clear to what extent loss of VSMCs, vessel dilation andreduced vessel integrity is a direct consequence of impaired TGF-β signalling or if it isa secondary effect caused by other disturbed TGF-β dependent functions (Folkman andD'Amore, 1996; Li et al., 1999; Oh et al., 2000).

Smad6 and Smad7 were originally identified as genes induced in vascular endotheliumin response to steady laminar shear stress, a physiologic biomechanical stimulus. Thissystem mimicking the effect of blood flowing past the endothelial cells, suggested animportant role in the maintenance of the vascular endothelium in the adult organism.But could these Smads also play a role in the developing endothelium? Zwijsen et al.investigated the expression of Smad7 mRNA in early mouse development by RT-PCR.They found that Smad7 was indeed upregulated in the developing vascular system ofthe mouse embryo, especially in endothelial cells of larger blood vessels, possibly as aresult of a larger blood flow. In addition, they observed a very early Smad 7 expressionduring preimplantation and gastrulation. Furthermore, overexpression of Smad7 in themouse zygote inhibited development at the 2-cell stage (Zwijsen et al., 2000).

In an approach that resembled the inhibition of TGF-β signals modulating angiogenesisand vasculogenesis, Vargesson et al. virally misexpressed Smad7 in the developingchick limb and head. They found that the larger vessels became dilated and frequentlydeveloped arteriovenous shunts similar to arteriovenous malformations (AVM).Expression of constitutively active BMP receptor could counteract the effect of Smad7overexpression, suggesting that a BMP-like pathway in contrast to TGF-β signallingwas the target of Smad7 inhibition. Moreover, Smad7 overexpressing vessels were nothaemorrhagic and had a normal structure. In addition, the dilation of vessels wasindependent of VSMCs and the recruitment of VSMCs was not affected. Takentogether, these findings suggest that the TGF-β pathway regulates vessel calibre and isnecessary for correct vessel connectivity in this experimental setting. They also indicatethat vessel dilation not necessarily must lead to vessel rupture and haemorrhages.Furthermore, the findings suggest that a VSMC coat is not a prerequisite for vessel

Page 27: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

27

maintenance and that TGF-β signalling is not involved in VSMC recruitment anddifferentiation (Vargesson and Laufer, 2001). Another study suggests that Smad7overexpression can attenuate the growth inhibitory effects of TGF-β on cultured ratsmooth muscle cells (Kato et al., 2001).

Lung morphogenesisTGF-β is involved in the negative regulation of lung growth and development duringearly lung organogenesis (Warburton and Lee, 1999). It had been shown that receptorregulated Smad2 and Smad3 were required for the TGF-β mediated inhibition ofembryonic lung branching morphogenesis and epithelial cell differentiation (Zhao et al.,1998). The role of Smad7 in lung morphogenesis was investigated by introducingSmad7 antisense oligonucleotides in cellular embryonic lung explants. There it wasdemonstrated that TGF-β mediated inhibition of lung branching was significantlyincreased in cells with abrogated Smad7 gene expression. In addition, it was observedby immunohistochemistry that Smad7 together with Smad2 and Smad3 co-localised indistal bronchial epithelial cells (Zhao et al., 2000a). A different approach to assess therole of Smad7 in lung morphogenesis, used adenoviral Smad7 overexpression. Theobservation was that Smad7, but not Smad6, could abolish TGF-β mediated branchinginhibition. In addition, Smad7 also inhibited TGF-β induced down regulation ofsurfactant protein C, a bronchial epithelial differentiation marker (Zhao et al., 2000b).

ApoptosisThe idea that Smad7 works as a modulator and attenuator of TGF-β mediated signallinghas gained acceptance over time. The fact that Smad7 can inhibit TGF-β signals alsosuggested that Smad7 might inhibit apoptosis initiated by TGF-β. Indeed, studies inmouse B- and T-cells, supported this theory and suggested that Smad7 could act as aninhibitor of activin-induced growth arrest and apoptosis (Ishisaki et al., 1998).Moreover, studies in WEHI 231 B-lymphocytes further corroborated this notion,showing that Smad7 could protect these cells from TGF-β induced apoptosis (Patil etal., 2000). Other reports suggested that the picture is much more complicated. Somestudies indicated that Smad7 by itself could work as an inducer of apoptosis. Transgenicmice overexpressing TGF-β1 under the control of the albumin promoter displayedincreased apoptosis associated with a depletion of podocytes in progressiveglomerulosclerosis (Schiffer et al., 2001). TGF-β1 and Smad7 both seemed able toinduce apoptosis, but through different pathways. The TGF-β mediated effect requiredthe activation of p38 MAP kinase and caspase-3, which was not required for Smad7-mediated apoptosis. In contrast to TGF-β, Smad7 was able to inhibit the nucleartranslocation and transcriptional activity of the cell survival-promoting factor, NF-kB(Schiffer et al., 2001). Additional evidence in prostatic carcinoma cells (PC-3U)suggested that Smad7 antisense RNA could inhibit TGF-β induced apoptosis(Landström et al., 2000). Moreover, overexpressing Smad7 alone using an induciblepromoter also resulted in apoptosis, implicating Smad7 itself as an effector of apoptosis(Landström et al., 2000).

Page 28: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

28

What signalling pathway could serve as the transducer of this Smad7 mediatedapoptosis? One candidate is the c-Jun N-terminal kinase (JNK) cascade. Mazars et al.demonstrated that expression of Smad7 caused a strong and sustained activation ofJNK. The use of a dominant-interfering mutant of mitogen-activated protein kinasekinase 4 (MKK4) completely abolished the Smad7-induced activation of JNK.Furthermore, expression of the mutant MKK4 also blocked the ability of Smad7 topromote cell death (Mazars et al., 2001).

FibrosisTGF-β is a potent regulator of extracellular matrix (ECM) by inducing proteins such ascollagen, fibronectin but also PAI-1 and TIMP. It is believed that a sustainedoverproduction of TGF-β caused by repeated chemical and or biological injury canresult in pathological amounts of ECM being collected in the tissue leading to afunctional deterioration (Border and Noble, 1994). TGF-β has been implicated in manyfibrotic disorders such as idiopathic pulmonary fibrosis (Broekelmann et al., 1991),autoimmune lung diseases (Deguchi, 1992), and bleomycin-induced lung fibrosis inanimal models (Khalil et al., 1989; Khalil et al., 1993; Westergren-Thorsson et al.,1993; Zhang et al., 1995). TGF-β was found to be necessary for the bleomycin-inducedtissue fibrosis in rodents (Giri et al., 1993). Based on these findings, Nakao et al.introduced Smad7 and Smad6 by adenoviral transient gene transfer into bleomycintreated mice. They observed that Smad7, but not Smad6, was able to reduce expressionof type I precollagen mRNA and abolish the morphological fibrotic responses intransgenic mice. In addition, the phospho-Smad2 immunoreactivity was reduced inbleomycin treated Smad7 transgenic mice, suggesting inhibition of the TGF-β/Smadpathway (Nakao et al., 1999).

Immunological disordersTGF-β is a powerful modulator of immune responses. It affects the differentiation,proliferation and state of activation of all immune cells. The dysregulation of TGF-βsignalling has been implicated in immune abnormalities related to autoimmunity,opportunistic infections and fibrotic complications (Letterio and Roberts, 1998). Studiesin murine models clearly underscore the connection between disrupted TGF-βsignalling and inflammatory disease. Deletion of the TGF-β1 gene in mice results insystemic inflammation and early death (Shull et al., 1992). In addition, overexpressionof a dominant negative TGF-β type II receptor gives rise to CD4+ T-cell hyperactivityand autoimmunity (Gorelik and Flavell, 2000). On a similar note, targeted disruption ofthe Smad3 gene resulted in inflammation of mucosal surfaces (Yang et al., 1999c).Given the strong immunosuppressive role of TGF-β1, what would the effect be if aTGF-β signalling inhibitor like Smad7 was overexpressed in immune cells? Nakao et al.studied the effect of Smad7 overexpression in peripheral T-cells in transgenic mice. Themutant T-cells were no longer growth inhibited by TGF-β but no overt phenotype wasobserved in the unchallenged transgenic mice. However, when subjected to antigen-

Page 29: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

29

induced airway inflammation the transgenic mice displayed an increased airwayreactivity and inflammation (Nakao et al., 2000). Despite the fact that this was anexample of a highly artificial system it can still give us some idea how Smad7 couldinfluence TGF-β signalling in vivo. Interestingly, a human disorder exists where theexpression levels of Smad7 are increased. Monteleone et al. (2001) demonstrated thatthe Smad7 expression was highly upregulated in the colon mucosa and in T-cellspurified from the colonic mucosa in patients with inflammatory bowel disease (IBD)and Crohns disease. The diseased tissue displayed reduced immunoreactivity forphospho-Smad3, indicating a reduced level of TGF-β signalling and responsiveness,despite an increased level of TGF-β ligand. When subjecting cells derived from IBDpatient tissue to antisense Smad7 oligonucleotides, they showed a reduction in Smad7protein expression. More importantly, silencing of Smad7 not only restored the abilityto respond to exogenous TGF-β, but lead also to reduced mRNA levels of the pro-inflammatory cytokines IFN-γ and TNF-α, the major counterplayers of TGF-β in theregulation of the inflammation status. This resulted in a TGF-β mediated inhibition ofIFN-γ and TNF-α production and suppression of inflammation (Monteleone et al.,2001).

CarcinogenesisHuman Smad7 has been mapped to the chromosomal region 18q21, between the Mad-Related-2 (MADR2) and Deleted in Pancreatic Cancer-4 (DPC4) locus. MADR2 andDPC4 encode the Smad2 and Smad4 proteins, respectively (Boulay et al., 2001; Röijeret al., 1998). Deletions in the 18q21 region are the most common genetic variationsobserved in colorectal carcinoma. An analysis of the gene copy number from colorectaltumour biopsies revealed that Smad2 or Smad4 were more often deleted than Smad7.On the contrary, Smad7 appeared to be more often amplified than the other genes(Boulay et al., 2001). The most common genetic alteration of Smad genes in the tumoursamples were simultaneous deletions of Smad2 and Smad4 while Smad7 was ofteneither retained by normal diploidy or amplified (Boulay et al., 2001).

Mutational analyses performed in other tumour types such as hepatocellular carcinoma,ovarian cancers and pancreatic cancer made it seem unlikely that deletion or duplicationof the Smad7 gene would influence cancer progression (Jonson et al., 1999; Wang et al.,2000; Kawate et al., 2001). Kleeff et al. (1999) however, claims that Smad7 canenhance tumourigenicity in pancreatic cancer. They observed increased Smad7 mRNAlevels in human pancreatic cancer. When they transfected COLO-357 human pancreaticcancer cells with Smad7, the cells lost their growth inhibitory response to TGF-β1. Inaddition these transfected cells showed increased anchorage-independent growth andaccelerated growth in nude mice (Kleeff et al., 1999).

Page 30: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

30

PRESENT INVESTIGATIONS

Smad expression in normal and malignant prostate (paper I)

The prostate gland is dependent on androgens for its growth and differentiation. It hasbeen observed that androgen ablation (i.e. castration) leads to regression and apoptosis(Kerr et al., 1972) in the rat ventral prostate (Kyprianou and Isaacs, 1988) as well as inPC-82 human prostate cancer (Kyprianou and Isaacs, 1988). Additional studies haveimplicated TGF-β as an inducer of apoptosis in the normal prostate as well as inmalignant prostatic carcinoma epithelial cells (Hsing et al., 1996; Kyprianou and Isaacs,1989; Landström et al., 1994; Rajah et al., 1997). Moreover, castration and, inparticular, castration in combination with estrogen treatment in a rat prostaticadenocarcinoma model caused an increased protein expression of TGF-β1, as well as ofTGF-β receptors (Landström et al., 1996). Taken together, these observations suggesteda role for TGF-β in the induction of programmed cell death in the normal and androgen-sensitive malignant prostate. However, the molecular mechanism behind such a TGF-βmediated apoptosis remained poorly understood.

We wanted to investigate a possible functional role for Smad proteins, the transducersof the TGF-β signal, in the normal and malignant prostate after castration. To addressthis issue, we examined the immunoreactivity of receptor-activated Smads (1, 2, 3 and5), the common mediator Smad4 and inhibitory Smads (6 and 7), in the rat ventralprostate and in an androgen and estrogen sensitive prostate tumour model (DunningR3327 PAP) (Isaacs, 1987). In parallel, we performed TUNEL-staining to assess thelevel and localisation of apoptotic cells within the two prostate models, which allowedus to correlate the Smad expression profile with the presence of apoptotic cells.

We observed increased expression levels of the Smads involved in TGF-β signaltransduction (i.e. Smad2, Smad3 and Smad4) in the epithelial cells of the normalprostate after castration, as well as in the Dunning tumour cells. Furthermore, in prostateepithelial cells after castration, we demonstrated an increased activation of Smad2 asdetected by antisera specific for phosphorylated Smad2 (Chen et al., 1998; Persson etal., 1998; Piek et al., 1999). These elevated levels of phosphorylated Smad2 were alsoobserved in the Dunning tumour after castration, but were less prominent. Interestingly,the expression of Smad2 was very low in the Dunning tumour cells prior to castration,in contrast to the rat ventral prostate, where the Smad2 levels were elevated beforecastration and remained high after androgen ablation. Smad3 immunostaining was alsoelevated in response to castration, both in normal and to a lesser degree in malignantprostate. Interestingly, in the rat ventral prostate Smad4 expression first increased aftercastration, only to become sharply reduced at later time points. The expression levels ofinhibitory Smad6 and Smad7 were low in the Dunning tumour when compared withnormal prostate, but increased significantly after treatment.

Page 31: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

31

There was a significant increase of the number of apoptotic cells in the normal andmalignant prostate in response to castration. Estrogen treatment of rats transplanted withthe Dunning tumours had an additive effect on castration-induced apoptosis.Interestingly, we were able to correlate areas of apoptotic cells with increased Smadprotein expression.

In summary, we observed elevated protein expression levels for Smad2, Smad3, andSmad4, increased activation of Smad2, as well as elevated protein expression levels ofinhibitory Smads in normal prostatic epithelial cells and, to some extent also in themalignant prostatic epithelial cells, after castration. These observations lend credence tothe notion that there is a correlation between the TGF-β/Smad pathway and apoptosis invivo. Deregulated expression or inactivation of components in this pathway mayinterfere with TGF-β induced apoptosis and create a favourable milieu for prostatictumour development.

Transcriptional regulation of the Smad7 gene (paper II)

Smad7 has been proposed to be an inhibitor of TGF-β signalling (Hayashi et al., 1997;Nakao et al., 1997). Previous studies have shown that Smad7 is upregulated in responseto TGF-β stimuli (Afrakhte et al., 1998; Nakao et al., 1997), suggesting a possible rolein a negative feedback loop modulating the TGF-β signal. We wanted to investigate themolecular mechanism underlying Smad7 activation, by studying the transcriptionalregulation of the mouse Smad7 gene promoter by TGF-β.

To obtain the mouse Smad7 promoter, we isolated several overlapping λ-phage clonesspanning the Smad7 genomic region. Using the promoterless pGL3 luciferase reportersystem we found that a 3 kb (kilo base pair) Xho1 fragment conferred TGF-β inducibleluciferase activity in HepG2 cells. Sequential deletions of the 3 kb promoter fragmentrevealed the presence of a TGF-β-responsive region containing a palindromicGTCTAGAC motif. Point mutations introduced into the palindromic sequenceabolished TGF-β mediated transcriptional activation. This palindromic element wasinitially identified by Zawel and co-workers (Zawel et al., 1998) in a artificial PCR-based screen in search for optimal Smad binding motifs. Indeed we could show byelectrophoretic mobility shift assays (EMSA) that Smad2, Smad3, and Smad4 were partof a complex associating with the palindromic Smad binding element (SBE).

In addition we observed the presence of consensus binding sequences for AP-1 and Sp1,which we showed subsequently to bind c-Jun and c-Fos, and Sp1, respectively, in vitro,in electrophoretic mobility shift assays (EMSA). Deletion or mutation of either AP-1 orSp1 binding motifs led to a dramatic decrease in promoter activity while retaining someTGF-β inducibility. In contrast, the major consequence of deletion or point mutation ofthe SBE site, was the complete loss of TGF-β inducibility. These findings suggested theimportance of co-operation between Smads and general transcription factors, such asAP-1 and Sp1.

Page 32: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

32

What could explain the ability of Sp1 and AP-1 to support the TGF-β response?Previous studies had shown that Smad3 and Smad4 could physically interact with c-Jun,JunB, JunD, and c-Fos and mediate transcription on the 12-O-tetradecanoyl-13-acetate-responsive gene promoter element (Liberati et al., 1999; Zhang et al., 1998). Otherstudies performed on a number of different TGF-β responsive promoters revealed thatSp1 sites could serve as major TGF-β responsive promoter elements (Datto et al., 1995;Greenwel et al., 1997; Li et al., 1995; Li et al., 1998; Moustakas and Kardassis, 1998).Moreover, it had been shown that c-Jun could superactivate Sp1 on the p21Waf/Cip1promoter (Kardassis et al., 1999). Our findings, however, suggested that thegeneral transcription factors for AP-1 and Sp1 had to collaborate together with theSmads in order to orchestrate the basal and TGF-β inducible transcriptional activity ofthe Smad7 promoter.

Targeting of the Smad7 gene (paper III)

The family of Smad proteins represents the major intracellular transducer of TGF-ßsignalling. Gene ablation of the different Smad proteins has revealed specificphysiological and developmental roles (see Table 1-3). In an effort to elucidate thefunctional relevance of Smad7 in vivo, we targeted the Smad7 locus.

We used a targeting vector where the coding region of exon-I including the translationalstart site, was replaced by a neomycin phosphotransferase (neo) selection cassette. Themorula aggregation method was used to generate mutant mice from targeted embryonicstem (ES) cells. The Smad7 homozygous mutant (-/-) mice were viable and fertile. Allmajor organs of the targeted mice such as brain, heart, kidneys, lung, liver, spleen andthymus as well as the larger and smaller intestines, were examined by autopsy andrevealed no consistent differences compared to wild type animals. Also the skeleton wasstudied with no consistent macroscopic morphological variations observed. However,some of the homozygous Smad7 (-/-) pups were significantly smaller than their wildtype (wt) and heterozygous (+/-) siblings. Genotyping of new-borns and mouse embryosof different ages indicated a certain amount of postnatal lethality of -/- mice within thefirst 1-4 weeks of life.

We observed by Northern blot and RT-PCR analyses that there was still TGF-ßinducible Smad7 mRNA expression in embryonic fibroblasts derived from wt or -/-E12.5 mouse embryos. As expected, the mutant mRNA also contained the neosequence, but lacked the coding sequence for the Smad7 exon-I. Interestingly, we foundthe mRNA for Smad6 to be highly upregulated in unstimulated mutant cells. However,we were unable to unambiguously identify Smad7 protein in mouse embryonicfibroblasts (MEFs) or mouse tissues, independent of the genotype.

Since Smad7 had been proposed as a negative modulator of TGF-ß signalling weexamined a number of classical TGF-β responses using MEFs. Assays for plasminogenactivator inhibitor-1 (PAI-1), a well-studied TGF-ß target, showed no differencebetween wt and -/- MEFs, and neither did assays for mitogenicity. In order to test a

Page 33: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

33

”pure” Smad-dependent cellular response, we transfected a CAGA12-luciferase reporterconstruct into wt and -/- MEFs. Stimulation of these cells with TGF-ß resulted in similarlevels of reporter vector activation in wt and -/- MEFs. This finding was consistent withthe observation that we were unable to detect any significant differences in thephosphorylation pattern of Smad2 upon TGF-ß stimulation between the two cell types.

Taken together, we did not find any indication for a significantly changed TGF-ßresponse in fibroblasts from mutant mouse embryos, despite the effect of the Smad7gene targeting on the size and survival of Smad7 homozygous new-born mice. Apossible explanation could be that Smad6, the other known negative regulator of TGF-βsignalling, was able to compensate for the loss of Smad7. Since we were unable toidentify endogenous Smad7 protein, we can not exclude that a truncated form of Smad7was still produced by the mutant organism, resulting in a hypomorph. In addition, thereis also the possibility that Smad7 is not involved in the modulation of TGF-β signallingin cells of fibroblast origin.

Page 34: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

34

FUTURE PERSPECTIVES

A major emphasis for future work will lie in the improvement of tools for theidentification of endogenous Smad7 protein in order to verify if the targeting hasresulted in a true knockout or a hypomorph. In the case there is residual protein, it willlack the N-domain, due to the targeting strategy. Very little is known about the N-domain of Smad7 and most of known Smad7-functions regarding negative regulation ofTGF-β signalling reside in the MH2-domain. However, there are observations that bothdomains are necessary for proper function (Hanyu et al., 2001; Nakayama et al., 2001).

The fact that some homozygous animals do survive creates the option to investigatemore about the function of Smad7 by challenging the animals in specific ways.

The back crossing of the Smad7 mutation into different inbred mouse strains will addfurther important information to our picture of Smad7 and TGF-β signalling.

To investigate the biological function of Smad7, many research groups have ectopicallyoverexpressed Smad7 from very strong promoters, resulting in presumablyunphysiological protein levels. We therefore believe that Smad7 gene ablation will beinstrumental in dissecting the physiological role of Smad7 in vivo.

Page 35: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

35

ACKNOWLEDGEMENTS

The work presented in this thesis was carried out at the Ludwig Institute for CancerResearch in Uppsala. I have had the privilege to interact with so many wonderful peoplethat not only have made the scientific hurdles so much easier to negotiate, but have alsoprovided me with so many joyful memories to be savoured in the future…I would like to express my sincere gratitude to the following people:

Rainer Heuchel, my excellent supervisor…for possessing a scientific enthusiasm andoptimism (certainly temporal optimism…) of a global and highly contagious nature. Butmost of all, for having a great HEART…

Aive Åhgren, my wonderful college, co-worker and “MOM”… for being the centre ofgravity around which the more disorganised celestial bodies of the Gene TargetingGroup revolve…

Maréne Landström, my excellent second supervisor… for your endless support andencouragement…and for long discussions about things that REALLY matters…

Carl-Henrik Heldin, for allowing me to explore the realms of science… for hisoutstanding scientific passion and knowledge…not to mention his ability to digestpainstaking piles of “less than perfect” PhD student scribble…and still make sense of it.

Akira Ishisaki and Takanori Murayama, my crazy Japanese buddies from the Clinlab days… For broadening my point of view and being great friends…

Keiko Funa, my first supervisor… for giving me an excellent introduction to therealities of science.

Olle Nylander, my Studirektor från Biomedicinarlinjen… for your support andenthusiasm ….och vad hände egentligen med Kalle och den där fästingen ?

Michael Cross, for his unquestionable wit and reassuring words: “Greger…..you knowit makes sense…”

All my fellow PhD students through the years, Tobias, Anders, Johan, Simon, Olli-Pekka, Ninna, Sofia, Ann-Sofi, Urban, Peter, Maria, Roya, Catrin, Daniel,Kristian, Jonas, Inga, Kaisa, Camilla, Jan, Katerina and many more, for sharingand verbally expressing “The Ludwig Spleen”, and explaining why words like“motgång föder motgång”, “delad bitterhet är dubbel bitterhet” samt “inget gott sominte har något ont med sig..” actually helps you through the day…

All ye fantastic Lab Technicians, Aino, Anita, Jill, Lotti and Gudrun… for makingthe day at the bench quite enjoyable

Page 36: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

36

Ulf “Velocipedofilen” Hellman…for his magic greasy touch and everything else…

Ulla and Christer …for long discussions on different topics…mostly economy

Lasse, CPU-Uffe, Ingegärd, Eva H, Inger, and Gullbritt…for making life in thescience-lane run a lot smoother…

Lucie and Annegret…for soulfood and lengthy chats…

People who have broadened my scientific horizon and much, much more…such asLars, Johan E, Valeria, Ester, Peter tD, Aris, Marie-José, Serhiy, Pontus, Ivan, DrKoza, Norbert and Arne…MILLE GRACIE!!

Johan Sandin…my old “Biomed-sambo”…for great laughs and vocal partyperformances…

Eva “Pling” Grönroos, my former fellow commuter and the better half of the dynamicsong writer duo “Reclaim the bad lyrics..”, for some really “artistic” experiences..

Bengt “Becke” Ejdesjö, for his ability to violate natural laws in order to deliver or fixANYTHING … and for making people question the linear concept of time… You arethe best!

My buddies, Ola, Lars-Erik, Gustav and Kaj…who, after all these years, are stilltalking to me, the Eremitus Upsaliensis….

The Toronto crowd, Tony, Berri, Nicki and Venus …for making my stay “overthere”highly enjoyable and educational… Thanks a bunch!!!

Pernilla Grehag, for allowing me to share some really beautiful moments…

My parents, Inger and Lars, for your endless support and love…

Veronika, mitt allt, min snutta…för allt…Älskar Dig.

Page 37: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

37

REFERENCES

Afrakhte, M., Morén, A., Jossan, S., Itoh, S., Sampath, K., Westermark, B., Heldin, C.-H., Heldin, N.-E., and ten Dijke, P. (1998). Induction of inhibitory Smad6 andSmad7 mRNA by TGF-b family members. Biochem Biophys Res Commun 249,505-511.

Anzano, M. A., Roberts, A. B., Smith, J. M., Sporn, M. B., and De Larco, J. E. (1983).Sarcoma growth factor from conditioned medium of virally transformed cells iscomposed of both type alpha and type beta transforming growth factors. Proc NatlAcad Sci U S A 80, 6264-6268.

Arthur, H. M., Ure, J., Smith, A. J. H., Renforth, G., Wilson, D. I., Torsney, E.,Charlton, R., Parums, D. V., Jowett, T., Marchuk, D. A., et al. (1999). Endoglin, anancillary TGFb receptor, is required for extraembryonic angiogenesis and plays akey role in heart development. Dev Biol 271, 42-53.

Ashcroft, G. S., Yang, X., Glick, A. B., Weinstein, M., Letterio, J. L., Mizel, D. E.,Anzano, M., Greenwell-Wild, T., Wahl, S. M., Deng, C., and Roberts, A. B. (1999).Mice lacking Smad3 show accelerated wound healing and an impaired localinflammatory response. Nat Cell Biol 1, 260-266.

Attisano, L., Cárcamo, J., Ventura, F., Weis, F., Massagué, J., and Wrana, J. L. (1993).Identification of human activin and TGFbeta type I receptors that form heteromerickinase complexes with type II receptors. Cell 75, 671-680.

Attisano, L., Wrana, J. L., Cheifetz, S., and Massagué, J. (1992). Novel activinreceptors: Distinct genes and alternative mRNA splicing generate a repertoire ofserine/threonine kinase receptors. Cell 68, 97-108.

Baarends, W. M., Van, H. M., Post, M., Van, d. S. P., Hoogerbrugge, J. W., De Winter,J. P., Uilenbroek, J., Karels, B., Wilming, L. G., Meijers, J., et al. (1994). A novelmember of the transmembrane serine/threonine kinase receptor family is specificallyexpressed in the gonads and in mesenchymal cells adjacent to the müllerian duct.Development 120, 189-197.

Bai, S., and Cao, X. (2002). A nuclear antagonistic mechanism of inhibitory Smads intransforming growth factor-beta signaling. J Biol Chem 277, 4176-4182.

Bai, S., Shi, X., Yang, X., and Cao, X. (2000). Smad6 as a transcriptional corepressor. JBiol Chem 275, 8267-8270.

Baker, J. C., and Harland, R. M. (1996). A novel mesoderm inducer, Madr2, functionsin the activin signal transduction pathway. Genes Dev 10, 1880-1889.

Basler, K., Edlund, T., Jessell, T. M., and Yamada, T. (1993). Control of cell pattern inthe neural tube: regulation of cell differentiation by dorsalin-1, a novel TGF betafamily member. Cell 73, 687-702.

Baur, S. T., Mai, J. J., and Dymecki, S. M. (2000). Combinatorial signaling throughBMP receptor IB and GDF5: shaping of the distal mouse limb and the geneticts ofdistal limb diversity. Development 127, 605-619.

Beddington, R. (1996). Left, right, left... turn. Nature 381, 116-117.Bitzer, M., von Gersdorff, G., Liang, D., Dominguez-Rosales, A., Beg, A. A., Rojkind,

M., and Bottinger, E. P. (2000). A mechanism of suppression of TGF-b/SMAD

Page 38: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

38

signaling by NF-k B/RelA. Genes Dev 14, 187-197.

Border, W. A., and Noble, N. A. (1994). Transforming growth factor b in tissue fibrosis.N Engl J Med 331, 1286-1292.

Boulay, J. L., Mild, G., Reuter, J., Lagrange, M., Terracciano, L., Lowy, A., Laffer, U.,Orth, B., Metzger, U., Stamm, B., et al. (2001). Combined copy status of 18q21genes in colorectal cancer shows frequent retention of SMAD7. GenesChromosomes Cancer 31, 240-247.

Bourdeau, A., Dumont, D. J., and Letarte, M. (1999). A murine model of hereditaryhemorrhagic telangiectasia. J Clin Invest 104, 1343-1351.

Broekelmann, T. J., Limper, A. H., Colby, T. V., and McDonald, J. A. (1991).Transforming growth factor-b1 is present at sites of extracellular matrix geneexpression in human pulmonary fibrosis. Proc Natl Acad Sci USA 88, 6642-?

Carcamo, J., Weis, F. M., Ventura, F., Wieser, R., Wrana, J. L., Attisano, L., andMassague, J. (1994). Type I receptors specify growth-inhibitory and transcriptionalresponses to transforming growth factor beta and activin. Mol Cell Biol 14, 3810-3821.

Cate, R. L., Mattaliano, R. J., Hession, C., Tizard, R., Farber, N. M., Cheung, A., Ninfa,E. G., Frey, A. Z., Gash, D. J., Chow, E. P., and et al. (1986). Isolation of the bovineand human genes for Mullerian inhibiting substance and expression of the humangene in animal cells. Cell 45, 685-698.

Celeste, A. J., Iannazzi, J. A., Taylor, R. C., Hewick, R. M., Rosen, V., Wang, E. A.,and Wozney, J. M. (1990). Identification of transforming growth factor b familymembers present in bone-inductive protein purified from bovine bone. Proc NatlAcad Sci USA 87, 9843-9847.

Chang, H., Huylebroeck, D., Verschueren, K., Guo, Q. X., Matzuk, M. M., and Zwijsen,A. (1999). Smad5 knockout mice die at mid-gestation due to multiple embryonicand extraembryonic defects. Development 126, 1631-1642.

Chang, H., Zwijsen, A., Vogel, H., Huylebroeck, D., and Matzuk, M. M. (2000). Smad5is essential for left-right assymmetry in mice. Dev Biol 219, 71-78.

Cheifetz, S., Bellón, T., Calés, C., Vera, S., Bernabeu, C., Massagué, J., and Letarte, M.(1992). Endoglin is a component of the transforming growth factor-b receptorsystem in human endothelial cells. J Biol Chem 267, 19027-19030.

Cheifetz, S., and Massagué, J. (1991). Isoform-specific transforming growth factor-bbinding proteins with membrane attachments sensitive to phosphatidylinositol-specific phospholipase C. J Biol Chem 266, 20767-20772.

Cheifetz, S., Weatherbee, J. A., Tsang, M. L.-S., Anderson, J. K., Mole, J. E., Lucas, R.,and Massagué, J. (1987). The transforming growth factor-b system, a complex patterof cross-reactive ligands and receptors. Cell 48, 409-415.

Chen, X., Rubock, M. J., and Whitman, M. (1996). A transcriptional partner for MADproteins in TGF-b signalling. Nature 383, 691-696.

Chen, X., Weisberg, E., Fridmacher, V., Watanabe, M., Naco, G., and Whitman, M.(1997). Smad4 and FAST-1 in the assembly of activin-responsive factor. Nature389, 85-89.

Chen, Y. G., Hata, A., Lo, R. S., Wotton, D., Shi, Y., Pavletich, N., and Massagué, J.(1998). Determinants of specificity in TGF-b signal transduction. Genes Dev 12,

Page 39: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

39

2144-2152.

Christian, J. L., and Nakayama, T. (1999). Can't get no SMADisfaction: Smad proteinsas positive and negative regulators of TGF-beta family signals. BioEssays 21, 382-390.

Conlon, F. L., Lyons, K. M., Takaesu, N., Barth, K. S., Kispert, A., Herrmann, B., andRobertson, E. J. (1994). A primary requirement for nodal in the formation andmaintenance of the primitive streak in the mouse. Development 120, 1919-1928.

Crawford, S. E., Stellmach, V., Murphy-Ullrich, J. E., Ribeiro, S. M. F., Lawler, J.,Hynes, R. O., Boivin, G. P., and Bouck, N. (1998). Thrombospondin-1 is a majoractivator of TGF-beta1 in vivo. Cell 93, 1159-1170.

Cui, Y., Jean, F., Thomas, G., and Christian, J. L. (1998). BMP-4 is proteolyticallyactivated by furin and/or PC6 during vertebrate embryonic development. EMBO J17, 4735-4743.

Cunningham, N. S., Paralkar, V., and Reddi, A. H. (1992). Osteogenin and recombinantbone morphogenetic protein 2B are chemotactic for human monocytes and stimulatetransforming growth factor b1 mRNA expression. Proc Natl Acad Sci USA 89,11740-11744.

Daniel, T. O., and Abrahamson, D. (2000). Endothelial signal integration in vascularassembly. Annu Rev Physiol 62, 649-671.

Datta, P. K., Chytil, A., Gorska, A. E., and Moses, H. L. (1998). Identification ofSTRAP, a novel WD domain protein in transforming growth factor-b signaling. JBiol Chem 273, 34671-34674.

Datta, P. K., and Moses, H. L. (2000). STRAP and Smad7 synergize in the inhibition oftransforming growth factor b signaling. Mol Cell Biol 20, 3157-3167.

Datto, M. B., Frederick, J. P., Pan, L., Borton, A. J., Zhuang, Y., and Wang, X.-F.(1999). Targeted disruption of Smad3 reveals an essential role in transforminggrowth factor b-mediated signal transduction. Mol Cell Biol 19, 2495-2504.

Datto, M. B., Yu, Y., and Wang, X.-F. (1995). Functional analysis of the transforminggrowth factor b responsive elements in the WAF1/Cip1/p21 promoter. J Biol Chem270, 28623-28628.

de Caestecker, M. P., Yahata, T., Wang, D., Parks, W. T., Huang, S., Hill, C. S., Shioda,T., Roberts, A. B., and Lechleider, R. J. (2000). The Smad4 activation domain(SAD) is a proline-rich, p300-dependent transcriptional activation domain. J BiolChem 275, 2115-2122.

De Larco, J. E., and Todaro, G. J. (1978). Growth factors from murine sarcoma virus-transformed cells. Proc Natl Acad Sci U S A 75, 4001-4005.

de Martin, R., Haendler, B., Hofer-Warbinek, R., Gaugitsch, H., Wrann, M.,Schlüsener, H., Seifert, J. M., Bodmer, S., Fontana, A., and Hofer, E. (1987).Complementary DNA for human glioblastoma-derived T cell suppressor factor, anovel member of the transforming growth factor-b gene family. EMBO J 6, 3673-3677.

Deguchi, Y. (1992). Spontaneous increase of transforming growth factor betaproduction by bronchoalveolar mononuclear cells of patients with systemic

Page 40: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

40

autoimmune diseases affecting the lung. Ann Rheum Dis 51, 362-365.

Dennler, S., Itoh, S., Vivien, D., ten Dijke, P., Huet, S., and Gauthier, J.-M. (1998).Direct binding of Smad3 and Smad4 to critical TGFb-inducible elements in thepromoter of human plasminogen activator inhibitor-type 1 gene. EMBO J 17, 3091-3100.

Derynck, R., Jarrett, J. A., Chen, E. Y., Eaton, D. H., Bell, J. R., Assoian, R. K.,Roberts, A. B., Sporn, M. B., and Goeddel, D. V. (1985). Human transforminggrowth factor-b complementary DNA sequence and expression in normal andtransformed cells. Nature 316, 701-705.

Derynck, R., Lindquist, P. B., Lee, A., Wen, D., Tamm, J., Graycar, J. L., Rhee, L.,Mason, A. J., Miller, D. A., Coffey, R. J., et al. (1988). A new type of transforminggrowth factor-b, TGF-b3. EMBO J 12, 3737-3743.

di Clemente, N., Wilson, C., Faure, E., Boussin, L., Carmillo, P., Tizard, R., Picard, J.-Y., Vigier, B., Josso, N., and Cate, R. (1994). Cloning, expression, and alternativesplicing of the receptor for anti-Müllerian hormone. Mol Endocrinol 8, 1006-1020.

Dickson, M. C., Martin, J. S., Cousins, F. M., Kulkarni, A. B., Karlsson, S., andAkhurst, R. J. (1995). Defective haematopoiesis and vasculogenesis in transforminggrowth factor-b1 knock out mice. Development 121, 1845-1854.

Dong, C., Li, Z., Alvarez Jr., R., Feng, X.-H., and Goldschmidt-Clermont, P. J. (2000).Microtubule binding to Smads may regulate TGFb activity. Mol Cell 5, 27-34.

Dubois, C. M., Laprise, M. H., Blanchette, F., Gentry, L. E., and Leduc, R. (1995).Processing of transforming growth factor beta1 precursor by human furinconvertase. J Biol Chem 270, 10618-10624.

Dudley, A. T., Lyons, K. M., and Robertson, E. J. (1995). A requirement for bonemorphogenetic protein-7 during development of the mammalian kidney and eye.Genes Dev 9, 2795-2807.

Ebisawa, T., Fukuchi, M., Murakami, G., Chiba, T., Tanaka, K., Imamura, T., andMiyazono, K. (2001). Smurf1 interacts with transforming growth factor-b type Ireceptor through Smad7 and induces receptor degradation. J Biol Chem 276, 12477-12480.

Ebner, R., Chen, R. H., Lawler, S., Zioncheck, T., and Derynck, R. (1993).Determination of type I receptor specificity by the type II receptors for TGF-beta oractivin. Science 262, 900-902.

Eppert, K., Scherer, S. W., Ozcelik, H., Pirone, R., Hoodless, P., Kim, H., Tsui, L.-C.,Bapat, B., Gallinger, S., Andrulis, I. L., et al. (1996). MADR2 maps to 18q21 andencodes a TGFb-regulated MAD-related protein that is functionally mutated incolorectal carcinoma. Cell 86, 543-552.

Feng, X.-H., and Derynck, R. (1997). A kinase subdomain of transforming growthfactor-b (TGF-b) type I receptor determines the TGF-b intracellular signalingspecificity. EMBO J 16, 3912-3923.

Feng, X. H., Zhang, Y., Wu, R. Y., and Derynck, R. (1998). The tumour suppressorSmad4/DPC4 and transcriptional adaptor CBP/p300 are coactivators for Smad3 inTGF-b-induced transcriptional activation. Genes Dev 12, 2153-2163.

Folkman, J., and D'Amore, P. A. (1996). Blood vessel formation: what is its molecularbasis? Cell 87, 1153-1155.

Page 41: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

41

Francois, V., and Bier, E. (1995). Xenopus chordin and Drosophila short gastrulationgenes encode homologous proteins functioning in dorsal-ventral axis formation. Cell

80, 19-20.

Franzen, P., ten Dijke, P., Ichijo, H., Yamashita, H., Schulz, P., Heldin, C. H., andMiyazono, K. (1993). Cloning of a TGF beta type I receptor that forms aheteromeric complex with the TGF beta type II receptor. Cell 75, 681-692.

Gaddy-Kurten, D., Tsuchida, K., and Vale, W. (1995). Activins and the receptor serinekinase superfamily. Recent Prog Horm Res 50, 109-129.

Gajdusek, C. M., Luo, Z., and Mayberg, M. R. (1993). Basic fibroblast growth factorand transforming growth factor beta-1: synergistic mediators of angiogenesis invitro. J Cell Physiol 157, 133-144.

Galvin, K. M., Donovan, M. J., Lynch, C. A., Meyer, R. I., Paul, R. J., Lorenz, J. N.,Fairchild-Huntress, V., Dixon, K. L., Dunmore, J. H., Gimbrone, M. A. J., et al.(2000). A role for smad6 in development and homeostasis of the cardiovascularsystem. Nat Genet 24, 171-174.

Gatherer, D., Ten Dijke, P., Baird, D. T., and Akhurst, R. J. (1990). Expression of TGF-beta isoforms during first trimester human embryogenesis. Development 110, 445-460.

Germain, S., Howell, M., Esslemont, G. M., and Hill, C. S. (2000). Homeodomain andwinged-helix transcription factors recruit activated Smads to distinct promoterelements via a common Smad interaction motif. Genes Dev 14, 435-451.

Giri, S. N., Hyde, D. M., and Hollinger, M. A. (1993). Effect of antibody totransforming growth factor beta on bleomycin induced accumulation of lungcollagen in mice. Thorax 48, 959-966.

Gorelik, L., and Flavell, R. A. (2000). Abrogation of TGFbeta signaling in T cells leadsto spontaneous T cell differentiation and autoimmune disease. Immunity 12, 171-181.

Gougos, A., and Letarte, M. (1990). Primary structure of endoglin, an RGD-containingglycoprotein of human endothelial cells. J Biol Chem 265, 8361-8364.

Goumans, M. J., and Mummery, C. (2000). Functional analysis of the TGFbreceptor/Smad pathway through gene ablation in mice. Int J Dev Biol 44, 253-265.

Graff, J. M., Bansal, A., and Melton, D. A. (1996). Xenopus Mad proteins transducedistinct subsets of signals for the TGFb superfamily. Cell 85, 479-487.

Greenwel, P., Inagaki, Y., Hu, W., Walsh, M., and Ramirez, F. (1997). Sp1 is requiredfor the early response of a2(I) collagen to transforming growth factor-b1. J BiolChem 272, 19738-19745.

Gu, Z., Nomura, M., Simpson, B. B., Lei, H., Feijen, A., van den Eijnden-van Raaij, J.,Donahoe, P. K., and Li, E. (1998). The type I activin receptor ActRIB is required foregg cylinder organization and gastrulation in the mouse. Genes Dev 12, 844-857.

Gu, Z., Reynolds, E. M., Song, J., Lei, H., Feijen, A., Yu, L., He, W., MacLaughlin, D.T., van den Eijnden-van Raaij, J., Donahoe, P. K., and Li, E. (1999). The type Iserine/threonine kinase receptor ActRIA (ALK2) is required for the gastrulation ofthe mouse embryo. Development 126, 2551-2561.

Page 42: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

42

Hanafusa, H., Ninomiya-Tsuji, J., Masuyama, N., Nishita, M., Fujisawa, J., Shibuya, H.,Matsumoto, K., and Nishida, E. (1999). Involvement of the p38 mitogen-activatedprotein kinase pathway in transforming growth factor-b-induced gene expression. J

Biol Chem 274, 27161-27167.

Hanai, J.-i., Chen, L. F., Kanno, T., Ohtani-Fujita, N., Kim, W. Y., Guo, W.-H.,Imamura, T., Ishidou, Y., Fukuchi, M., Shi, M.-J., et al. (1999). Interaction andfunctional cooperation of PEBP2/CBF with Smads.

Synergistic induction of the immunoglobulin germline ca promoter. J Biol Chem 274,31577-31582.

Hanyu, A., Ishidou, Y., Ebisawa, T., Shimanuki, T., Imamura, T., and Miyazono, K.(2001). The N domain of Smad7 is essential for specefic inhibition of tranforminggrowth factor-b signaling. J Cell Biol 155, 1017-1027.

Harland, R. M. (1994). The transforming growth factor beta family and induction of thevertebrate mesoderm: bone morphogenetic proteins are ventral inducers. Proc NatlAcad Sci U S A 91, 10243-10246.

Hata, A., Lagna, G., Massague, J., and Hemmati-Brivanlou, A. (1998). Smad6 inhibitsBMP/Smad1 signaling by specifically competing with the Smad4 tumoursuppressor. Genes Dev 12, 186-197.

Hata, A., Seoane, J., Lagna, G., Montalvo, E., Hemmati-Brivanlou, A., and Massagué,J. (2000). OAZ uses distinct DNA- and protein-binding zinc fingers in separateBMP-Smad and Olf signaling pathways. Cell 100, 229-240.

Hayashi, H., Abdollah, S., Qiu, Y., Cai, J., Xu, Y. Y., Grinnell, B. W., Richardson, M.A., Topper, J. N., Gimbrone, M. A. J., Wrana, J. L., and Falb, D. (1997). The MAD-related protein Smad7 associates with the TGFb receptor and functions as anantagonist of TGFb signaling. Cell 89, 1165-1173.

He, W. W., Gustafson, M. L., Hirobe, S., and Donahoe, P. K. (1993). Developmentalexpression of four novel serine/threonine kinase receptors homologous to theactivin/transforming growth factor-beta type II receptor family. Dev Dyn 196, 133-142.

Heldin, C.-H., Miyazono, K., and ten Dijke, P. (1997). TGF-b signalling from cellmembrane to nucleus through SMAD proteins. Nature 390, 465-471.

Hemmati-Brivanlou, A., and Melton, D. A. (1992). A truncated activin receptor inhibitsmesoderm induction and formation of axial structures in Xenopus embryos. Nature359, 609-614.

Hocevar, B. A., Brown, T. L., and Howe, P. H. (1999). TGF-beta induces fibronectinsynthesis through a c-Jun N-terminal kinase-dependent, Smad4-independentpathway. Embo J 18, 1345-1356.

Hogan, B. L. M. (1996). Bone morphogenetic proteins: multifunctional regulators ofvertebrate development. Genes Dev 10, 1580-1594.

Hoodless, P. A., Haerry, T., Abdollah, S., Stapleton, M., O'Connor, M. B., Attisano, L.,and Wrana, J. L. (1996). MADR1, a MAD-related protein that functions in BMP2signaling pathways. Cell 85, 489-500.

Howell, M., Itoh, F., Pierreux, C. E., Valgeirsdóttir, S., Itoh, S., ten Dijke, P., and Hill,C. S. (1999). Xenopus Smad4b is the co-Smad component of developmentally-

Page 43: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

43

regulated transcription factor complexes responsibe for induction of earlymesodermal genes. Dev Biol 214, 354-369.

Hsing, A. Y., Kadomatsu, K., Bonham, M. J., and Danielpour, D. (1996). Regulation ofapoptosis induced by transforming growth factor-beta1 in nontumourigenic and

tumourigenic rat prostatic epithelial cell lines. Cancer Res 56, 5146-5149.

Hsu, D. R., Economides, A. N., Wang, X., Eimon, P. M., and Harland, R. M. (1998).The Xenopus dorsalizing factor gremlin identifies a novel family of secreted proteinsthat antagonize BMP activities. Mol Cell 1, 673-683.

Hua, X. X., Miller, Z. A., Wu, G., Shi, Y. G., and Lodish, H. F. (1999). Specificity intransforming growth factor beta-induced transcription of the plasminogen activatorinhibitor-1 gene: Interactions of promoter DNA, transcription factor muE3, andSmad proteins. Proc Natl Acad Sci USA 96, 13130-13135.

Hungerford, J. E., and Little, C. D. (1999). Developmental biology of the vascularsmooth muscle cell: building a multilayered vessel wall. J Vasc Res 36, 2-27.

Imamura, T., Takase, M., Nishihara, A., Oeda, E., Hanai, J., Kawabata, M., andMiyazono, K. (1997). Smad6 inhibits signalling by the TGF-b superfamily. Nature389, 622-626.

Isaacs, J. T. (1987). Development and characteristics of the available animal modelsystems for the study of prostatic cancer. Prog Clin Biol Res 239, 513-576.

Ishisaki, A., Yamato, K., Nakao, A., Nonaka, K., Ohguchi, M., ten Dijke, P., andNishihara, T. (1998). Smad7 is an activin-inducible inhibitor of activin-inducedgrowth arrest and apoptosis in mouse B cells. J Biol Chem 273, 24293-24296.

Itoh, S., Landström, M., Hermansson, A., Itoh, F., Heldin, C.-H., Heldin, N.-E., and tenDijke, P. (1998). Transforming growth factor b1 induces nuclear export of inhibitorySmad7. J Biol Chem 273, 29195-29201.

Janknecht, R., Wells, N. J., and Hunter, T. (1998). TGF-b-stimulated cooperation ofsmad proteins with the coactivators CBP/p300. Genes Dev 12, 2114-2119.

Jonson, T., Gorunova, L., Dawiskiba, S., Andrén-Sandberg, Å., Stenman, G., ten Dijke,P., Johansson, B., and Höglund, M. (1999). Molecular analyses of the 15q and 18qSMAD genes in pancreatic cancer. Genes, Chrom and Cancer 24, 62-71.

Josso, N., Cate, R. L., Picard, J. Y., Vigier, B., di Clemente, N., Wilson, C., Imbeaud,S., Pepinsky, R. B., Guerrier, D., Boussin, L., and et al. (1993). Anti-mullerianhormone: the Jost factor. Recent Prog Horm Res 48, 1-59.

Kaartinen, V., Voncken, J. W., Shuler, C., Warburton, D., Bu, D., Heisterkamp, N., andGroffen, J. (1995). Abnormal lung development and cleft palate in mice lackingTGF-beta3 indicates defects of epithelial-mesenchymal interaction. Nature Genet11, 415-421.

Kardassis, D., Papakosta, P., Pardali, K., and Moustakas, A. (1999). c-Jun transactivatesthe promoter of the human p21WAF1/Cip1 gene by acting as a superactivator of theubiquitous transcription factor Sp1. J Biol Chem 274, 29572-29581.

Kato, S., Ueda, S., Tamaki, K., Fujii, M., Miyazono, K., ten Dijke, P., Morimatsu, M.,and Okuda, S. (2001). Ectopic expression of Smad7 inhibits transforming growthfactor-b responses in vascular smooth muscle cells. Life Sci 69, 2641-2652.

Page 44: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

44

Kawate, S., Ohwada, S., Hamada, K., Koyama, T., Takenoshita, S., Morishita, Y., andHagiwara, K. (2001). Mutational analysis of the Smad6 and Smad7 genes inhepatocellular carcinoma. Int J Mol Med 8, 49-52.

Kavsak, P., Rasmussen, R. K., Causing, C. G., Bonni, S., Zhu, H., Thomsen, G. H., andWrana, J. L. (2000). Smad7 binds to Smurf2 to form an E3 ubiquitin ligase that

targets the TGFb receptor for degradation. Mol Cell 6, 1365-1375.

Kerr, J. F., Wyllie, A. H., and Currie, A. R. (1972). Apoptosis: a basic biologicalphenomenon with wide-ranging implications in tissue kinetics. Br J Cancer 26, 239-257.

Khalil, N., Bereznay, O., Sporn, M., and Greenberg, A. H. (1989). Macrophageproduction of transforming growth factor beta and fibroblast collagen synthesis inchronic pulmonary inflammation. J Exp Med 170, 727-737.

Khalil, N., Whitman, C., Zuo, L., Danielpour, D., and Greenberg, A. (1993). Regulationof alveolar macrophage transforming growth factor-beta secretion by corticosteroidsin bleomycin-induced pulmonary inflammation in the rat. J Clin Invest 92, 1812-1818.

Kingsley, D. M. (1994). The TGF-b superfamily: new members, new receptors, andnew genetic tests of function in different organisms. Genes Dev 8, 133-146.

Klagsbrun, M., and D'Amore, P. A. (1991). Regulators of angiogenesis. Annu RevPhysiol 53, 217-239.

Kleeff, J., Ishiwata, T., Maruyama, H., Friess, H., Truong, P., Büchler, M. W., Falb, D.,and Korc, M. (1999). The TGF-beta signaling inhibitor Smad7 enhancestumourigenicity in pancreatic cancer. Oncogene 18, 5363-5372.

Koenig, B. B., Cook, J. S., Wolsing, D. H., Ting, J., Tiesman, J. P., Correa, P. E., Olson,C. A., Pecquet, A. L., Ventura, F., Grant, R. A., and al., e. (1994). Characterizationand cloning of a receptor for BMP-2 and BMP-4 from NIH 3T3 cells. Mol Cell Biol14, 5961-5974.

Kretzschmar, M., Doody, J., and Massagué, J. (1997a). Opposing BMP and EGFsignalling pathways converge on the TGF-b family mediator Smad1. Nature 389,618-622.

Kretzschmar, M., Doody, J., Timokhina, I., and Massagué, J. (1999). A mechanism ofrepression of TGFb/Smad signaling by oncogenic Ras. Genes Dev 13, 804-816.

Kretzschmar, M., Liu, F., Hata, A., Doody, J., and Massagué, J. (1997b). The TGF-bfamily mediator Smad1 is phosphorylated directly and activated functionally by theBMP receptor kinase. Genes Dev 11, 984-995.

Kulkarni, A. B., Huh, C.-G., Becker, D., Geiser, A., Lyght, M., Flanders, K. C.,Roberts, A. B., Sporn, M. B., Ward, J. M., and Karlsson, S. (1993). Transforminggrowth factor b1 null mutation in mice causes excessive inflammatory response andearly death. Proc Natl Acad Sci USA 90, 770-774.

Kyprianou, N., and Isaacs, J. T. (1988). Activation of programmed cell death in the ratventral prostate after castration. Endocrinology 12, 357-364.

Kyprianou, N., and Isaacs, J. T. (1989). Expression of transforming growth factor-betain the rat ventral prostate during castration-induced programmed cell death. MolEndocrinol 3, 1515-1522.

Page 45: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

45

Labbe, E., Letamendia, A., and Attisano, L. (2000). Association of Smads withlymphoid enhancer binding factor 1/T cell-specific factor mediates cooperativesignaling by the transforming growth factor-beta and wnt pathways. Proc Natl AcadSci U S A 97, 8358-8363.

Lagna, G., Hata, A., Hemmati-Brivanlou, A., and Massagué, J. (1996). Partnershipbetween DPC4 and SMAD proteins in TGF-b signalling pathways. Nature 383, 832-

836.

Landström, M., Damber, J. E., and Bergh, A. (1994). Prostatic tumour regrowth afterinitially successful castration therapy may be related to a decreased apoptotic celldeath rate. Cancer Res 54, 4281-4284.

Landström, M., Eklöv, S., Colosetti, P., Nilsson, S., Damber, J.-E., Bergh, A., and Funa,K. (1996). Estrogen induces apoptosis in a rat prostatic adenocarcinoma: associationwith an increased expression of TGF-b1 and its type-I and type-II receptors. Int JCancer 67, 573-579.

Landström, M., Heldin, N.-E., Bu, S., Hermansson, A., Itoh, S., ten Dijke, P., andHeldin, C.-H. (2000). Smad7 mediates apoptosis induced by transforming growthfactor b in prostatic carcinoma cells. Curr Biol 10, 535-538.

Larsson, J., Goumans, M. J., Sjostrand, L. J., van Rooijen, M. A., Ward, D., Leveen, P.,ten Dijke, P., Mummery, C. L., and Karlsson, S. (2001). Abnormal angiogenesis butintact hematopoietic potential in TGF-beta type I receptor-deficient mice. Embo J20, 1663-1673.

Lawson, K. A., Dunn, N. R., Roelen, B. A., Zeinstra, L. M., Davis, A. M., Wright, C.V., Korving, J. P., and Hogan, B. L. (1999). Bmp4 is required for the generation ofprimordial germ cells in the mouse embryo. Genes Dev 13, 424-436.

Lechleider, R. J., Ryan, J. L., Garrett, L., Eng, C., Deng, C., Wynshaw-Boris, A., andRoberts, A. B. (2001). Targeted mutagenesis of Smad1 reveals an essential role inchorioallantoic fusion. Dev Biol 240, 157-167.

Letterio, J. J., and Roberts, A. B. (1996). Transforming growth factor-beta1-deficientmice: Identification of isoform-specific activities in vivo. J Leukocyte Biol 59, 769-774.

Letterio, J. J., and Roberts, A. B. (1998). Regulation of immune responses by TGF-b.Annu Rev Immunol 16, 137-161.

Li, C.-Y., Suardet, L., and Little, J. B. (1995). Potential role of WAF1/Cip1/p21 as amediator of TGF-b cytoinhibitory effect. J Biol Chem 270, 4971-4974.

Li, D. Y., Sorensen, L. K., Brooke, B. S., Urness, L. D., Davis, E. C., Taylor, D. G.,Boak, B. B., and Wendel, D. P. (1999). Defective angiogenesis in mice lackingendoglin. Science 284, 1534-1537.

Li, J. M., Datto, M. B., Shen, X., Hu, P. P. C., Yu, Y., and Wang, X. F. (1998). Sp1, butnot Sp3, functions to mediate promoter activation by TGF-b through canonical Sp1binding sites. Nucleic Acids Res 26, 2449-2456.

Liberati, N. T., Datto, M. B., Frederick, J. P., Shen, X., Wong, C., Rougier-Chapman, E.M., and Wang, X.-F. (1999). Smads bind directly to the Jun family of AP-1transcription factors. Proc Natl Acad Sci USA 96, 4844-4849.

Page 46: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

46

Lin, H. Y., Wang, X.-F., Ng-Eaton, E., Weinberg, R. A., and Lodish, H. F. (1992).Expression cloning of the TGF-b type II receptor, a functional transmembraneserine/threonine kinase. Cell 68, 775-785.

Lin, L. F., Doherty, D. H., Lile, J. D., Bektesh, S., and Collins, F. (1993). GDNF: a glialcell line-derived neurotrophic factor for midbrain dopaminergic neurons. Science260, 1130-1132.

Lin, X., Liang, M., and Feng, X. H. (2000). Smurf2 is a ubiquitin E3 ligase mediatingproteasome-dependent degradation of Smad2 in transforming growth factor-b

signaling. J Biol Chem 275, 36818-36822.

Liu, F., Ventura, F., Doody, J., and Massagué, J. (1995). Human type II receptor forbone morphogenic proteins (BMPs): Extension of the two-kinase receptor model tothe BMPs. Mol Cell Biol 15, 3479-3486.

Lo, R. S., Chen, Y. G., Shi, Y., Pavletich, N. P., and Massagué, J. (1998). The L3 loop:a structural motif determining specific interactions between SMAD proteins andTGF-b receptors. EMBO J 17, 996-1005.

Lo, R. S., and Massagué, J. (1999). Ubiquitin-dependent degradation of TGF-b-activated Smad2. Nature Cell Biol 1, 472-478.

Lo, R. S., Wotton, D., and Massague, J. (2001). Epidermal growth factor signaling viaRas controls the Smad transcriptional co-repressor TGIF. Embo J 20, 128-136.

Lopez-Casillas, F., Wrana, J. L., and Massague, J. (1993). Betaglycan presents ligand tothe TGF beta signaling receptor. Cell 73, 1435-1444.

Luo, G., Hofmann, C., Bronckers, A. L., Sohocki, M., Bradley, A., and Karsenty, G.(1995). BMP-7 is an inducer of nephrogenesis, and is also required for eyedevelopment and skeletal patterning. Genes Dev 9, 2808-2820.

Luo, K., Stroschein, S. L., Wang, W., Chen, D., Martens, E., Zhou, S., and Zhou, Q.(1999). The Ski oncoprotein interacts with the Smad proteins to repress TGFbsignaling. Genes Dev 13, 2196-2206.

Lux, A., Attisano, L., and Marchuk, D. A. (1999). Assignment of transforming growthfactor beta1 and beta3 and a third new ligand to the type I receptor ALK-1. J BiolChem 274, 9984-9992.

Lyons, K., Graycar, J. L., Lee, A., Hashmi, S., Lindquist, P. B., Chen, E. Y., Hogan, B.L. M., and Derynck, R. (1989). Vgr-1, a mammalian gene related to Xenopus Vg-1,is a member of the transforming growth factor b gene superfamily. Proc Natl AcadSci USA 86, 4554-4558.

Macias-Silva, M., Abdollah, S., Hoodless, P. A., Pirone, R., Attisano, L., and Wrana, J.L. (1996). MADR2 is a substrate of the TGFbeta receptor and its phosphorylation isrequired for nuclear accumulation and signaling. Cell 87, 1215-1224.

Macias-Silva, M., Hoodless, P. A., Tang, S. J., Buchwald, M., and Wrana, J. L. (1998).Specific activation of Smad1 signaling pathways by the BMP7 type I receptor,ALK2. J Biol Chem 273, 25628-25636.

Madisen, L., Webb, N. R., Rose, T. M., Marquardt, H., Ikeda, T., Twardzik, D.,Seyedin, S., and Purchio, A. F. (1988). Transforming growth factor-beta 2: cDNAcloning and sequence analysis. DNA 7, 1-8.

Page 47: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

47

Madri, J. A., Pratt, B. M., and Tucker, A. M. (1988). Phenotypic modulation ofendothelial cells by transforming growth factor-b depends upon the composition andorganization of the extracellular matrix. J Cell Biol 106, 1375-1384.

Mahony, D., and Gurdon, J. B. (1995). A type 1 serine/threonine kinase receptor thatcan dorsalize mesoderm in Xenopus. Proc Natl Acad Sci U S A 92, 6474-6478.

Massague, J. (1990). The transforming growth factor-beta family. Annu Rev Cell Biol6, 597-641.

Massague, J. (1996). Neurotrophic factors. Crossing receptor boundaries. Nature 382,29-30.

Massague, J. (1998). TGF-beta signal transduction. Annu Rev Biochem 67, 753-791.

Massague, J., and Weis-Garcia, F. (1996). Serine/threonine kinase receptors: mediatorsof transforming growth factor beta family signals. Cancer Surv 27, 41-64.

Massagué, J., and Wotton, D. (2000). Transcriptional control by the TGF-b/Smadsignaling system. EMBO J 19, 1745-1754.

Masuyama, N., Hanafusa, H., Kusakabe, M., Shibuya, H., and Nishida, E. (1999).Identification of two Smad4 proteins in Xenopus. Their common and distinctproperties. J Biol Chem 274, 12163-12170.

Mather, J. P. (1996). Follistatins and a 2-macroglobulin are soluble binding proteins forinhibin and activin. Horm Res 45, 207-210.

Mathews, L. S., and Vale, W. W. (1991). Expression cloning of an activin receptor, apredicted transmembrane serine kinase. Cell 65, 973-982.

Mathews, L. S., Vale, W. W., and Kintner, C. R. (1992). Cloning of a second type ofactivin receptor and functional characterization in Xenopus embryos. Science 255,1702-1705.

Matzuk, M. M. (1995). Functional analysis of mammalian members of the transforminggrowth factor-beta superfamily. Trends Endocrinol Metab 6, 120-127.

Mazars, A., Lallemand, F., Prunier, C., Marais, J., Ferrand, N., Pessah, M., Cherqui, G.,and Atfi, A. (2001). Evidence for a role of the JNK cascade in Smad7-mediatedapoptosis. J Biol Chem 276, 36797-36803.

McPherron, A. C., Lawler, A. M., and Lee, S. J. (1997). Regulation of skeletal musclemass in mice by a new TGF-beta superfamily member. Nature 387, 83-90.

Mehler, M. F., Mabie, P. C., Zhang, D., and Kessler, J. A. (1997). Bone morphogeneticproteins in the nervous system. Trends Neurosci 20, 309-317.

Merwin, J. R., Anderson, J. M., Kocher, O., Van Itallie, C. M., and Madri, J. A. (1990).Transforming growth factor beta 1 modulates extracellular matrix organization andcell-cell junctional complex formation during in vitro angiogenesis. J Cell Physiol142, 117-128.

Mishina, Y., Suzuki, A., Ueno, N., and Behringer, R. R. (1995). Bmpr encodes a type Ibone morphogenetic protein receptor that is essential for gastrulation during mouseembryogenesis. Genes Dev 9, 3027-3037.

Monteleone, G., Kumberova, A., Croft, N. M., McKenzie, C., Steer, H. W., andMacDonald, T. T. (2001). Blocking Smad7 restores TGF-beta1 signaling in chronicinflammatory bowel disease. J Clin Invest 108, 601-609.

Page 48: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

48

Moriguchi, T., Kuroyanagi, N., Yamaguchi, K., Gotoh, Y., Irie, K., Kano, T.,Shirakabe, K., Muro, Y., Shibuya, H., Matsumoto, K., et al. (1996). A novel kinasecascade mediated by mitogen-activated protein kinase kinase 6 and MKK3. J BiolChem 271, 13675-13679.

Moustakas, A., and Kardassis, D. (1998). Regulation of the human p21/WAF1/Cip1promoter in hepatic cells by functional interactions between Sp1 and Smad familymembers. Proc Natl Acad Sci USA 95, 6733-6738.

Munger, J. S., Harpel, J. G., Giancotti, F. G., and Rifkin, D. B. (1998a). The integrinavb6 binds and activates latent TGFb1: A mechanism for regulating pulmonaryinflammation and fibrosis. Cell 96, 319-328.

Munger, J. S., Harpel, J. G., Giancotti, F. G., and Rifkin, D. B. (1998b). Interactionsbetween growth factors and integrins: Latent forms of transforming growth factor-

beta are ligands for the integrin alphavbeta1. Mol Biol Cell 9, 2627-2638.

Munger, J. S., Harpel, J. G., Gleizes, P. E., Mazzieri, R., Nunes, I., and Rifkin, D. B.(1997a). Latent transforming growth factor-beta: structural features and mechanismsof activation. Kidney Int 51, 1376-1382.

Munger, J. S., Harpel, J. G., Gleizes, P. E., Mazzieri, R., Nunes, I., and Rifkin, D. B.(1997b). Latent transforming growth factor-beta: Structural features andmechanisms of activation. Kidney Int 51, 1376-1382.

Nakamura, T., Takio, K., Eto, Y., Shibai, H., Titani, K., and Sugino, H. (1990). Activin-binding protein from rat ovary is follistatin. Science 247, 836-838.

Nakao, A., Afrakhte, M., Morén, A., Nakayama, T., Christian, J. L., Heuchel, R., Itoh,S., Kawabata, M., Heldin, N.-E., Heldin, C.-H., and ten Dijke, P. (1997).Identification of Smad7, a TGFb-inducible antagonist of TGF-b signalling. Nature389, 631-635.

Nakao, A., Fujii, M., Matsumura, R., Kumano, K., Saito, Y., Miyazono, K., andIwamoto, I. (1999). Transient gene transfer and expression of Smad7 preventsbleomycin-induced lung fibrosis in mice. J Clin Invest 104, 5-11.

Nakao, A., Miike, S., Hatano, M., Okumura, K., Tokuhisa, T., Ra, C., and Iwamoto, I.(2000). Blockade of transforming growth factor beta/Smad signaling in T cells byoverexpression of Smad7 enhances antigen-induced airway inflammation andairway reactivity. J Exp Med 192, 151-158.

Nakayama, T., Berg, L. K., and Christian, J. L. (2001). Dissection of inhibitory Smadproteins: both N- and C- terminal domains are necessary for full activities ofXenopus Smad6 and Smad7. Mech Dev 100, 251-262.

Nakayama, T., Cui, Y., and Christian, J. L. (2000). Regulation of BMP/Dpp signalingduring embryonic development. Cell Mol Life Sci 57, 943-956.

Nishihara, A., Hanai, J.-i., Okamoto, N., Yanagisawa, J., Kato, S., Miyazono, K., andKawabata, M. (1998). Role of p300, a transcriptional coactivator, in signalling ofTGF-b. Genes Cells 3, 613-623.

Nishita, M., Hashimoto, M. K., Ogata, S., Laurent, M. N., Ueno, N., Shibuya, H., andCho, K. W. (2000). Interaction between Wnt and TGF-beta signalling pathwaysduring formation of Spemann's organizer. Nature 403, 781-785.

Page 49: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

49

Nishitoh, H., Ichijo, H., Kimura, M., Matsumoto, T., Makishima, F., Yamaguchi, A.,Yamashita, H., Enomoto, S., and Miyazono, K. (1996). Identification of type I andtype II serine/threonine kinase receptors for growth/differentiation factor-5. J BiolChem 271, 21345-21352.

Nohno, T., Ishikawa, T., Saito, T., Hosokawa, K., Noji, S., Wolsing, D. H., andRosenbaum, J. S. (1995). Identification of a human type II receptor for bonemorphogenetic protein-4 that forms differential heteromeric complexes with bonemorphogenetic protein type I receptors. J Biol Chem 270, 22522-22526.

Nomura, M., and Li, E. (1998). Smad2 role in mesoderm formation, left-right patterningand craniofacial development. Nature 393, 786-790.

Ogawa, Y., Schmidt, D. K., Dasch, J. R., Chang, R.-J., and Glaser, C. B. (1992).Purification and characterization of transforming growth factor-b2.3 and -b1.2heterodimers from bovine bone. J Biol Chem 267, 2325-2328.

Oh, S. P., and Li, E. (1997). The signaling pathway mediated by the type IIB activinreceptor controls axial patterning and lateral asymmetry in the mouse. Genes Dev11, 1812-1826.

Oh, S. P., Seki, T., Goss, K. A., Imamura, T., Yi, Y., Donahoe, P. K., Li, L., Miyazono,K., ten Dijke, P., Kim, S., and Li, E. (2000). Activin receptor-like kinase 1modulates transforming growth factor-b1 signaling in the regulation of angiogenesis.Proc Natl Acad Sci USA 97, 2626-2631.

Onichtchouk, D., Chen, Y. G., Dosch, R., Gawantka, V., Dellus, H., Massagué, J., andNiehrs, C. (1999). Silencing of TGF-beta signalling by the pseudoreceptor BAMBI.Nature 401, 480-485.

Oshima, M., Oshima, H., and Taketo, M. M. (1996). TGF-b receptor type II deficiencyresults in defects of yolk sac hematopoiesis and vasculogenesis. Dev Biol 179, 297-302.

Padgett, R. W., Wozney, J. M., and Gelbart, W. M. (1993). Human BMP sequences canconfer normal dorsal-ventral patterning in the Drosophila embryo. Proc Natl AcadSci U S A 90, 2905-2909.

Pardali, E., Xie, X.-Q., Tsapogas, P., Itoh, S., Arvanitidis, K., Heldin, C.-H., ten Dijke,P., Grundström, T., and Sideras, P. (2000). Smad and AML proteins synergisticallyconfer transforming growth factor b1 responsiveness to human germ-line IgA genes.J Biol Chem 275, 3552-3560.

Patil, S., Wildey, G. M., Brown, T. L., Choy, L., Derynck, R., and Howe, P. H. (2000).Smad7 is induced by CD40 and protects WEHI 231 B-lymphocytes fromtransforming growth factor-beta -induced growth inhibition and apoptosis. J BiolChem 275, 38363-38370.

Pelton, R. W., Dickinson, M. E., Moses, H. L., and Hogan, B. L. (1990). In situhybridization analysis of TGF beta 3 RNA expression during mouse development:comparative studies with TGF beta 1 and beta 2. Development 110, 609-620.

Pelton, R. W., Nomura, S., Moses, H. L., and Hogan, B. L. (1989). Expression oftransforming growth factor beta 2 RNA during murine embryogenesis. Development106, 759-767.

Page 50: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

50

Pepper, M. S. (1997). Transforming growth factor-beta: vasculogenesis, angiogenesis,and vessel wall integrity. Cytokine Growth Factor Rev 8, 21-43.

Pepper, M. S., Belin, D., Montesano, R., Orci, L., and Vassalli, J. D. (1990).Transforming growth factor-beta 1 modulates basic fibroblast growth factor-inducedproteolytic and angiogenic properties of endothelial cells in vitro. J Cell Biol 111,743-755.

Pepper, M. S., Vassalli, J. D., Orci, L., and Montesano, R. (1993). Biphasic effect oftransforming growth factor-beta 1 on in vitro angiogenesis. Exp Cell Res 204, 356-363.

Persson, U., Izumi, H., Souchelnytskyi, S., Itoh, S., Grimsby, S., Engström, U., Heldin,C.-H., Funa, K., and ten Dijke, P. (1998). The L45 loop in type I receptors for TGF-b family members is a critical determinant in specifying Smad isoform activation.FEBS Lett 434, 83-87.

Piccolo, S., Sasai, Y., Lu, B., and De Robertis, E. M. (1996). Dorsoventral patterning inXenopus: Inhibition of ventral signals by direct binding of chordin to BMP-4. Cell86, 589-598.

Piek, E., Franzén, P., Heldin, C.-H., and ten Dijke, P. (1997). Characterization of a 60-kDa cell surface-associated transforming growth factor-b binding protein that caninterfere with transforming growth factor-b receptor binding. J Cell Physiol 173,447-459.

Piek, E., Westermark, U., Kastemar, M., Heldin, C.-H., van Zoelen, E. J., Nistér, M.,and ten Dijke, P. (1999). Expression of transforming-growth-factor (TGF)-breceptors and Smad proteins in glioblastoma cell lines with distinct responses toTGF-b1. Int J Cancer 80, 756-763.

Proetzel, G., Pawlowski, S. A., Wiles, M. V., Yin, M. Y., Boivin, G. P., Howles, P. N.,Ding, J. X., Ferguson, M. W. J., and Doetschman, T. (1995). Transforming growthfactor-beta3 is required for secondary palate fusion. Nature Genet 11, 409-414.

Pulaski, L., Landström, M., Heldin, C.-H., and Souchelnytskyi, S. (2001).Phosphorylation of Smad7 at Ser-249 does not interfere with its inhibitory role intransforming growth factor-b-dependent signaling but affects Smad7-dependenttranscriptional activation. J Biol Chem 276, 14344-14349.

Puri, M. C., Rossant, J., Alitalo, K., Bernstein, A., and Partanen, J. (1995). The receptortyrosine kinase TIE is required for integrity and survival of vascular endothelialcells. Embo J 14, 5884-5891.

Raftery, L. A., Twombly, V., Wharton, K., and Gelbart, W. M. (1995). Genetic screensto identify elements of the decapentaplegic signaling pathway in Drosophila.Genetics 139, 241-254.

Rajah, R., Valentinis, B., and Cohen, P. (1997). Insulin like growth factor (IGF)-bindingprotein-3 induces apoptosis and mediates the effects of transforming growth factor-beta1 on programmed cell death through a p53- and IGF-independent mechanism. JBiol Chem 272, 12181-12188.

Risau, W. (1997). Mechanisms of angiogenesis. Nature 386, 671-674.Roberts, A. B., Anzano, M. A., Lamb, L. C., Smith, J. M., and Sporn, M. B. (1981).

New class of transforming growth factors potentiated by epidermal growth factor:isolation from non-neoplastic tissues. Proc Natl Acad Sci U S A 78, 5339-5343.

Page 51: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

51

Roberts, A. B., Kim, S.-J., Noma, T., Glick, A. B., Lafyatis, R., Lechleider, R.,Jakowlew, S. B., Geiser, A., O'Reilly, M. A., Danielpour, D., and B., S. M. (1991).Multiple forms of TGF-b: Distinct promoters and differential expression. CibaFound Sympo 157, 7-28.

Roberts, A. B., and Sporn, M. B. (1989). Regulation of endothelial cell growth,architecture, and matrix synthesis by TGF-beta. Am Rev Respir Dis 140, 1126-1128.

Roberts, A. B., and Sporn, M. B. (1990). The transforming growth factor-bs. In PeptideGrowth Factors and Their Receptors, Part I, M. B. Sporn, and A. B. Roberts, eds.(Berlin, Springer-Verlag), pp. 419-472.

Roberts, A. B., and Sporn, M. B. (1992). Differential expression of the TGF-b isoformsin embryogenesis suggests specific roles in developing and adult tissues. MolReprod Dev 32, 91-98.

Roberts, A. B., and Sporn, M. B. (1993). Physiological actions and clinical applicationsof transforming growth factor-b (TGF-b). Growth Factors 8, 1-9.

Rodriguez, C., Chen, F., Weinberg, R. A., and Lodish, H. F. (1995). Cooperativebinding of transforming growth factor (TGF)-b2 to the types I and II TGF-breceptors. J Biol Chem 270, 15919-15922.

Rodriguez Esteban, C., Capdevila, J., Economides, A. N., Pascual, J., Ortiz, A., andIzpisua Belmonte, J. C. (1999). The novel Cer-like protein Caronte mediates theestablishment of embryonic left-right asymmetry. Nature 401, 243-251.

Rosenzweig, B. L., Imamura, T., Okadome, T., Cox, G. N., Yamashita, H., ten Dijke,P., Heldin, C. H., and Miyazono, K. (1995). Cloning and characterization of ahuman type II receptor for bone morphogenetic proteins. Proc Natl Acad Sci U S A92, 7632-7636.

Rydén, M., Imamura, T., Jörnvall, H., Belluardo, N., Neuveu, I., Trupp, M., Okadome,T., ten Dijke, P., and Ibáñez, C. F. (1996). A novel type I receptor serine-threoninekinase predominantly expressed in the adult central nervous system. J Biol Chem271, 30603-30609.

Röijer, E., Morén, A., ten Dijke, P., and Stenman, G. (1998). Assignment of the Smad7gene (MADH7) to human chromosome 18q21.1 by fluorescence in situhybridization. Cytogenet Cell Genet 81, 189-190.

Saharinen, J., Taipale, J., Monni, O., and Keski-Oja, J. (1998). Identification andcharacterization of a new latent transforming growth factor-beta-binding protein,LTBP-4. J Biol Chem 273, 18459-18469.

Saijoh, Y., Adachi, H., Sakuma, R., Yeo, C. Y., Yashiro, K., Watanabe, M., Hashiguchi,H., Mochida, K., Ohishi, S., Kawabata, M., et al. (2000). Left-right asymmetricexpression of lefty2 and nodal is induced by a signaling pathway that includes thetranscription factor FAST2. Mol Cell 5, 35-47.

Sampath, T. K., Rashka, K. E., Doctor, J. S., Tucker, R. F., and Hoffmann, F. M.(1993). Drosophila transforming growth factor beta superfamily proteins induceendochondral bone formation in mammals. Proc Natl Acad Sci U S A 90, 6004-6008.

Page 52: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

52

Sanford, L. P., Ormsby, I., Gittenberger-de Groot, A. C., Sariola, H., Friedman, R.,Boivin, G. P., Cardell, E. L., and Doetschman, T. (1997). TGFb2 knockout micehave multiple developmental defects that are non-overlapping with other TGFbknockout phenotypes. Development 124, 2659-2670.

Sano, Y., Harada, J., Tashiro, S., Gotoh-Mandeville, R., Maekawa, T., and Ishii, S.(1999). ATF-2 is a common nuclear target of Smad and TAK1 pathways intransforming growth factor-b signaling. J Biol Chem 274, 8949-8957.

Sato, T. N., Tozawa, Y., Deutsch, U., Wolburg-Buchholz, K., Fujiwara, Y., Gendron-Maguire, M., Gridley, T., Wolburg, H., Risau, W., and Qin, Y. (1995). Distinct rolesof the receptor tyrosine kinases Tie-1 and Tie-2 in blood vessel formation. Nature376, 70-74.

Savage, C., Das, P., Finelli, A. L., Townsend, S. R., Sun, C.-Y., Baird, S. E., andPadgett, R. W. (1996). Caenorhabditis elegans genes sma-2, sma-3, and sma-4define a conserved family of transforming growth factor b pathway components.Proc Natl Acad Sci USA 93, 790-794.

Schiffer, M., Bitzer, M., Roberts, I. S., Kopp, J. B., ten Dijke, P., Mundel, P., andBöttinger, E. P. (2001). Apoptosis in podocytes induced by TGF-b and Smad7. JClin Invest 108, 807-816.

Schmid, P., Cox, D., Bilbe, G., Maier, R., and McMaster, G. K. (1991). Differentialexpression of TGF beta 1, beta 2 and beta 3 genes during mouse embryogenesis.Development 111, 117-130.

Schrewe, H., Gendron-Maguire, M., Harbison, M. L., and Gridley, T. (1994). Micehomozygous for a null mutation of activin beta B are viable and fertile. Mech Dev47, 43-51.

Segarini, P. R., D.M., R., and Seyedin, S. M. (1989). Binding of transforming growthfactor-b to cell surface proteins varies with cell type. Mol Endocrinol 3, 261-272.

Sekelsky, J. J., Newfeld, S. J., Raftery, L. A., Chartoff, E. H., and Gelbart, W. M.(1995). Genetic characterization and cloning of Mothers against dpp, a generequired for decapentaplegic function in Drosophila melanogaster. Genetics 139,1347-1358.

Shen, X., Hu, P. P., Liberati, N. T., Datto, M. B., Frederick, J. P., and Wang, X. F.(1998). TGF-b-induced phosphorylation of Smad3 regulates its interaction withcoactivator p300/CREB-binding protein. Mol Biol Cell 9, 3309-3319.

Shi, Y. G., Wang, Y. F., Jayaraman, L., Yang, H. J., Massagué, J., and Pavletich, N. P.(1998). Crystal structure of a Smad MH1 domain bound to DNA: insights on DNAbinding in TGF-b signaling. Cell 94, 585-594.

Shibuya, H., Iwata, H., Masuyama, N., Gotoh, Y., Yamaguchi, K., Irie, K., Matsumoto,K., Nishida, E., and Ueno, N. (1998). Role of TAK1 and TAB1 in BMP signaling inearly Xenopus development. EMBO J 17, 1019-1028.

Shibuya, H., Yamaguchi, K., Shirakabe, K., Tonegawa, A., Gotoh, Y., Ueno, N., Irie,K., Nishida, E., and Matsumoto, K. (1996). TAB1: An activator of the TAK1MAPKKK in TGF-b signal transduction. Science 272, 1179-1182.

Shirakabe, K., Yamaguchi, K., Shibuya, H., Irie, K., Matsuda, S., Moriguchi, T., Gotoh,Y., Matsumoto, K., and Nishida, E. (1997). TAK1 mediates the ceramide signaling

Page 53: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

53

to stress-activated protein kinase/c-Jun N-terminal kinase. J Biol Chem 272, 8141-8144.

Shovlin, C. L., and Letarte, M. (1999). Hereditary haemorrhagic telangiectasia andpulmonary arteriovenous malformations: issues in clinical management and reviewof pathogenic mechanisms. Thorax 54, 714-729.

Shull, M. M., Ormsby, I., Kier, A. B., Pawlowski, S., Diebold, R. J., Yin, M., Allen, R.,Sidman, C., Proetzel, G., Calvin, D., et al. (1992). Targeted disruption of the mousetransforming growth factor-b1 gene results in multifocal inflammatory disease.Nature 359, 693-699.

Sirard, C., de la Pompa, J. L., Elia, A., Itie, A., Mirtsos, C., Cheung, A., Hahn, S.,Wakeham, A., Schwartz, L., Kern, S. E., et al. (1998). The tumour suppressor geneSmad4/Dpc4 is required for gastrulation and later for anterior development of themouse embryo. Genes Dev 12, 107-119.

Song, J., Oh, S. P., Schrewe, H., Nomura, M., Lei, H., Okano, M., Gridley, T., and Li,E. (1999). The type II activin receptors are essential for egg cylinder growth,gastrulation, and rostral head development in mice. Dev Biol 213, 157-169.

Souchelnytskyi, S., Nakayama, T., Nakao, A., Morén, A., Heldin, C.-H., Christian, J.L., and ten Dijke, P. (1998). Physical and functional interaction of murine andXenopus Smad7 with bone morphogenetic protein receptors and transforminggrowth factor-b receptors. J Biol Chem 273, 25364-25370.

Souchelnytskyi, S., Tamaki, K., Engström, U., Wernstedt, C., ten Dijke, P., and Heldin,C.-H. (1997). Phosphorylation of Ser465 and Ser467 in the C terminus of Smad2mediates interaction with Smad4 and is required for transforming growth factor-bsignaling. J Biol Chem 272, 28107-28115.

Souchelnytskyi, S., ten Dijke, P., Miyazono, K., and Heldin, C.-H. (1996).Phosphorylation of Ser165 in TGF-b type I receptor modulates TGF-b1-inducedcellular responses. EMBO J 15, 6231-6240.

Spencer, F. A., Hoffmann, F. M., and Gelbart, W. M. (1982). Decapentaplegic: a genecomplex affecting morphogenesis in Drosophila melanogaster. Cell 28, 451-461.

Sun, Y., Liu, X. D., Ng-Eaton, E., Lodish, H. F., and Weinberg, R. A. (1999). SnoN andSki protooncoproteins ave rapidly degraded in response to transforming growthfactor beta signaling. Proc Natl Acad Sci USA 96, 12442-12447.

Suri, C., Jones, P. F., Patan, S., Bartunkova, S., Maisonpierre, P. C., Davis, S., Sato, T.N., and Yancopoulos, G. D. (1996). Requisite role of Angiopoietin-1, a ligand forthe TIE2 receptor, during embryonic angiogenesis. Cell 87, 1171-1180.

Taipale, J., and Keski-Oja, J. (1997). Growth factors in the extracellular matrix. Faseb J11, 51-59.

Takatsu, Y., Nakamura, M., Stapleton, M., Danos, M. C., Matsumoto, K., O'Connor, M.B., Shibuya, H., and Ueno, N. (2000). TAK1 participates in c-Jun N-terminal kinasesignaling during Drosophila development. Mol Cell Biol 20, 3015-3026.

ten Dijke, P., Hansen, P., Iwata, K. K., Pieler, C., and Foulkes, J. G. (1988).Identification of another member of the transforming growth factor type b genefamily. Proc Natl Acad Sci USA 85, 4175-4179? 4715-4719?

ten Dijke, P., Ichijo, H., Franzen, P., Schulz, P., Saras, J., Toyoshima, H., Heldin, C. H.,and Miyazono, K. (1993). Activin receptor-like kinases: a novel subclass of cell-

Page 54: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

54

surface receptors with predicted serine/threonine kinase activity. Oncogene 8, 2879-2887.

ten Dijke, P., Yamashita, H., Ichijo, H., Franzen, P., Laiho, M., Miyazono, K., andHeldin, C. H. (1994a). Characterization of type I receptors for transforming growthfactor-beta and activin. Science 264, 101-104.

ten Dijke, P., Yamashita, H., Sampath, T. K., Reddi, A. H., Estevez, M., Riddle, D. L.,Ichijo, H., Heldin, C. H., and Miyazono, K. (1994b). Identification of type Ireceptors for osteogenic protein-1 and bone morphogenetic protein-4. J Biol Chem269, 16985-16988.

Thomsen, G. H. (1996). Xenopus mothers against decapentaplegic is an embryonicventralizing agent that acts downstream of the BMP-2/4 receptor. Development 122,2359-2366.

Topper, J. N., Cai, J., Qiu, Y., Anderson, K. R., Xu, Y.-Y., Deeds, J. D., Feeley, R.,Gimeno, C. J., Woolf, E. A., Tayber, O., et al. (1997). Vascular MADs: Two novelMAD-related genes selectively inducible by flow in human vascular endothelium.Proc Natl Acad Sci USA 94, 9314-9319.

Topper, J. N., DiChiara, M. R., Brown, J. D., Williams, A. J., Falb, D., Collins, T., andGimbrone, M. A., Jr. (1998). CREB binding protein is a required coactivator forSmad-dependent, transforming growth factor-b transcriptional responses inendothelial cells. Proc Natl Acad Sci USA 95, 9506-9511.

Tsuchida, K., Sawchenko, P. E., Nishikawa, S. I., and Vale, W. W. (1996). Molecularcloning of a novel type I receptor serine/threonine kinase for the TGFbetasuperfamily from rat brain. Mol Cell Neurosci 7, 467-478.

Tsuji, K., Ito, Y., and Noda, M. (1998). Expression of thePEBP2alphaA/AML3/CBFA1 gene is regulated by BMP4/7 heterodimer and itsoverexpression suppresses type I collagen and osteocalcin gene expression inosteoblastic and nonosteoblastic mesenchymal cells. Bone 22, 87-92.

Tsukazaki, T., Chiang, T. A., Davison, A. F., Attisano, L., and Wrana, J. L. (1998).SARA, a FYVE domain protein that recruits Smad2 to the TGFb receptor. Cell 95,779-791.

Ulloa, L., Doody, J., and Massagu, J. (1999). Inhibition of transforming growth factor-b/SMAD signalling by the interferon-gamma/STAT pathway. Nature 397, 710-713.

Urness, L. D., Sorensen, L. K., and Li, D. Y. (2000). Arteriovenous malformations inmice lacking activin receptor-like kinase-1. Nat Genet 26, 328-331.

Waldrip, W. R., Bikoff, E. K., Hoodless, P. A., Wrana, J. L., and Robertson, E. J.(1998). Smad2 signaling in extraembryonic tissues determines anterior-posteriorpolarity of the early mouse embryo. Cell 92, 797-808.

Vale, W., Hsueh, A., Rivier, C., and Yu, J. (1990). The inhibin/activin family ofhormones and growth factors. In Peptide Growth Factors and Their Receptors, PartII, M. B. Sporn, and A. B. Roberts, eds. (Berlin, Springer-Verlag), pp. 211-248.

Wang, D., Kanuma, T., Mizumuma, H., Ibuki, Y., and Takenoshita, S. (2000). Mutationanalysis of the Smad6 and Smad7 gene in human ovarian cancers. Int J Oncol 17,1087-1091.

Page 55: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

55

Wang, X.-F., Lin, H. Y., Ng-Eaton, E., Downward, J., Lodish, H. F., and Weinberg, R.A. (1991). Expression cloning and characterization of the TGF-b type III receptor.Cell 67, 797-805.

Warburton, D., and Lee, M. K. (1999). Current concepts on lung development. CurrOpin Pediatr 11, 188-192.

Vargesson, N., and Laufer, E. (2001). Smad7 misexpression during embryonicangiogenesis causes vascular dilation and malformations independently of vascularsmooth muscle cell function. Dev Biol 240, 499-516.

Vassalli, A., Matzuk, M. M., Gardner, H. A., Lee, K. F., and Jaernisch, R. (1994).Activin/inhibin beta B subunit gene disruption leads to defects in eyeliddevelopment and female reproduction. Genes Dev 8, 414-427.

Weeks, D. L., and Melton, D. A. (1987). A maternal mRNA localized to the vegetalhemisphere in Xenopus eggs codes for a growth factor related to TGF-b. Cell 51,861-867.

Weinstein, M., Yang, X., Li, C., Xu, X., Gotay, J., and Deng, C. X. (1998). Failure ofegg cylinder elongation and mesoderm induction in mouse embryos lacking thetumour supressor Smad2. Proc Natl Acad Sci USA 95, 9378-9383.

Westergren-Thorsson, G., Hernnas, J., Sarnstrand, B., Oldberg, A., Heinegard, D., andMalmstrom, A. (1993). Altered expression of small proteoglycans, collagen, andtransforming growth factor-beta 1 in developing bleomycin-induced pulmonaryfibrosis in rats. J Clin Invest 92, 632-637.

Wharton, K. A., Thomsen, G. H., and Gelbart, W. M. (1991). Drosophila 60A gene,another transforming growth factor b family member, is closely related to humanbone morphogenetic proteins. Proc Natl Acad Sci USA 88, 9214-9218.

Wieser, R., Wrana, J. L., and Massagué, J. (1995). GS domain mutations thatconstitutively activate TbR-I, the downstream signaling component in the TGF-breceptor complex. EMBO J 14, 2199-2208.

Vikkula, M., Boon, L. M., Carraway III, K. L., Calvert, J. T., Diamonti, A. J.,Goumnerov, B., Pasyk, K. A., Marchuk, D. A., Warman, M. L., Cantley, L. C., et al.(1996). Vascular dysmorphogenesis caused by an activating mutation in the receptortyrosine kinase TIE2. Cell 87, 1181-1190.

Vindevoghel, L., Lechleider, R. J., Kon, A., de Caestecker, M. P., Uitto, J., Roberts, A.B., and Mauviel, A. (1998). SMAD3/4-dependent transcriptional activation of thehuman type VII collagen gene (COL7A1) promoter by transforming growth factor b.Proc Natl Acad Sci USA 95, 14769-14774.

Winnier, G., Blessing, M., Labosky, P. A., and Hogan, B. L. (1995). Bonemorphogenetic protein-4 is required for mesoderm formation and patterning in themouse. Genes Dev 9, 2105-2116.

Wotton, D., Lo, R. S., Lee, S., and Massagué, J. (1999). A Smad transcriptionalcorepressor. Cell 97, 29-39.

Wozney, J. M., Rosen, V., Celeste, A. J., Mitsock, L. M., Whitters, M. J., Kriz, R. W.,Hewick, R. M., and Wang, E. A. (1988). Novel regulators of bone formation:Molecular clones and activities. Science 242, 1528-1534.

Page 56: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

56

Wrana, J. L., Attisano, L., Wieser, R., Ventura, F., and Massagué, J. (1994). Mechanismof activation of the TGF-b receptor. Nature 370, 341-347.

Xu, L., Chen, Y. G., and Massagué, J. (2000). The nuclear import function of Smad2 ismasked by SARA and unmasked by TGFb-dependent phosphorylation. Nat CellBiol 2, 559-562.

Yamaguchi, K., Nagai, S.-i., Ninomiya-Tsuji, J., Nishita, M., Tamai, K., Irie, K., Ueno,N., Nishida, E., Shibuya, H., and Matsumoto, K. (1999). XIAP, a cellular member ofthe inhibitor of apoptosis protein family, links the receptors to TAB1-TAK1 in theBMP signaling pathway. EMBO J 18, 179-187.

Yamaguchi, Y., Mann, D. M., and Ruoslahti, E. (1990). Negative regulation oftransforming growth factor-b by the proteoglycan decorin. Nature 346, 281-284.

Yamashita, H., ten Dijke, P., Franzen, P., Miyazono, K., and Heldin, C. H. (1994).Formation of hetero-oligomeric complexes of type I and type II receptors fortransforming growth factor-beta. J Biol Chem 269, 20172-20178.

Yamashita, H., ten Dijke, P., Huylebroeck, D., Sampath, T. K., Andries, M., Smith, J.,C., Heldin, C.-H., and Miyazono, K. (1995a). Osteogenic protein-1 binds to activintype II receptors and induces certain activin-like effects. J Cell Biol 130, 217-226.

Yamashita, H., ten Dijke, P., Huylebroeck, D., Sampath, T. K., Andries, M., Smith, J.C., Heldin, C. H., and Miyazono, K. (1995b). Osteogenic protein-1 binds to activintype II receptors and induces certain activin-like effects. J Cell Biol 130, 217-226.

Yancopoulos, G. D., Davis, S., Gale, N. W., Rudge, J. S., Wiegand, S. J., and Holash, J.(2000). Vascular-specific growth factors and blood vessel formation. Nature 407,242-248.

Yang, X., Li, C., Xu, X., Deng, C. (1998). The tumor suppressor SMAD4/DPC4 isessential for epiblast proliferation and mesoderm induction in mice. Proc Natl AcadSci USA 95, 3667-3672.

Yang, X., Castilla, L. H., Xu, X., Li, C., Gotay, J., Weinstein, M., Liu, P. P., and Deng,C. X. (1999a). Angiogenesis defects and mesenchymal apoptosis in mice lackingSMAD5. Development 126, 1571-1580.

Yang, X., Letterio, J. J., Lechleider, R. J., Chen, L., Hayman, R., Gu, H., Roberts, A. B.,and Deng, C. (1999b). Targeted disruption of SMAD3 results in impaired mucosalimmunity and diminished T cell responsiveness to TGF-beta. Embo J 18, 1280-1291.

Yi, S. E., Daluiski, A., Pederson, R., Rosen, V., and Lyons, K. M. (2000). The type IBMP receptor BMPRIB is required for chondrogenesis in the mouse limb.Development 127, 621-630.

Yingling, J. M., Datto, M. B., Wong, C., Frederick, J. P., Liberati, N. T., and Wang, X.F. (1997). Tumour suppressor Smad4 is a transforming growth factor b-inducibleDNA binding protein. Mol Cell Biol 17, 7019-7028.

Yokouchi, M., Wakioka, T., Sakamoto, H., Yasukawa, H., Ohtsuka, S., Sasaki, A.,Ohtsubo, M., Valius, M., Inoue, A., Komiya, S., and Yoshimura, A. (1999). APS, anadaptor protein containing PH and SH2 domains, is associated with the PDGFreceptor and c-Cbl and inhibits PDGF-indnced mitogenesis. Oncogene 18, 759-767.

Page 57: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

57

Yu, Q., and Stamenkovic, I. (2000). Cell surface-localized matrix metalloproteinase-9proteolytically activates TGF-beta and promotes tumour invasion and angiogenesis.Genes Dev 14, 163-176.

Zawel, L., Dai, J. L., Buckhaults, P., Zhou, S., Kinzler, K. W., Vogelstein, B., and Kern,S. E. (1998). Human Smad3 and Smad4 are sequence-specific transcriptionactivators. Mol Cell 1, 611-617.

Zhang, H., and Bradley, A. (1996). Mice deficient for BMP2 are nonviable and havedefects in amnion/chorion and cardiac development. Development 122, 2977-2986.

Zhang, K., Flanders, K. C., and Phan, S. H. (1995). Cellular localization of transforminggrowth factor-beta expression in bleomycin-induced pulmonary fibrosis. Am JPathol 147, 352-361.

Zhang, Y., Feng, X. H., and Derynck, R. (1998). Smad3 and Smad4 cooperate with c-Jun/c-Fos to mediate TGF-b-induced transcription. Nature 394, 909-913.

Zhang, Y., Feng, X.-H., Wu, R.-Y., and Derynck, R. (1996). Receptor-associated Madhomologues synergize as effectors of the TGF-b response. Nature 383, 168-172.

Zhao, J., and Buick, R. N. (1995). Regulation of transforming growth factor b receptorsin H-ras oncogene-transformed rat intestinal epithelial cells. Cancer Res 55, 6181-6188.

Zhao, J., Crowe, D. L., Castillo, C., Wuenschell, C., Chai, Y., and Warburton, D.(2000a). Smad7 is a TGF-beta-inducible attenuator of Smad2/3-mediated inhibitionof embryonic lung morphogenesis. Mech Dev 93, 71-81.

Zhao, J., Shi, W., Chen, H., and Warburton, D. (2000b). Smad7 and Smad6differentially modulate transforming growth factor beta -induced inhibition ofembryonic lung morphogenesis. J Biol Chem 275, 23992-23997.

Zhao, J. S., Lee, M., Smith, S., and Warburton, D. (1998). Abrogation of Smad3 andSmad2 or of Smad4 gene expression positively regulates murine embryonic lungbranching morphogenesis in culture. Dev Biol 194, 182-195.

Zhu, H., Kavsak, P., Abdollah, S., Wrana, J. L., and Thomsen, G. H. (1999). A Smadubiquitin ligase targets the BMP pathway and affects embryonic pattern formation.Nature 400, 687-693.

Zhu, Y., Richardson, J. A., Parada, L. F., and Graff, J. M. (1998). Smad3 mutant micedevelop metastatic colorectal cancer. Cell 94, 703-714.

Zimmerman, L. B., De Jesus-Escobar, J. M., and Harland, R. M. (1996). The SpemannOrganizer signal noggin binds and inactivates bone morphorgenetic protein 4. Cell86, 599-606.

Zwijsen, A., van Rooijen, M. A., Goumans, M.-J., Dewulf, N., Bosman, E. A., tenDijke, P., Mummery, C. L., and Huylebroeck, D. (2000). Expression of theinhibitory Smad7 in early mouse development and upregulation during embryonicvasculogenesis. Dev Dyn 218, 663-670.

Özkaynak, E., Rueger, D. C., Drier, E. A., Corbett, C., Ridge, R. J., Sampath, T. K., andOppermann, H. (1990). OP-1 cDNA encodes an osteogenic protein in the TGF-bfamily. EMBO J 9, 2085-2093.

Özkaynak, E., Schnegelsberg, P. N. J., Jin, D. F., Clifford, G. M., Warren, F. D., Drier,E. A., and Oppermann, H. (1992). Osteogenic protein-2: A new member of the

Page 58: 161534/FULLTEXT01.pdf · transcriptional induction as well as the basal promoter activity. Gene ablation of Smad proteins has revealed specific physiological and developmental roles.

58

transforming growth factor-b superfamily expressed early in embryogenesis. J BiolChem 267, 25220-25227.