REGULATION OF SKELETAL MUSCLE DEVELOPMENT

AND DIFFERENTIATION BY SKI

By

HONG ZHANG

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Thesis Advisor: Dr. Ed Stavnezer

Department of Biochemistry

CASE WESTERN RESERVE UNIVERSITY

January, 2009 CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES

We hereby approve the thesis/dissertation of

______Hong Zhang______candidate for the ______Doctor of Philosophy______degree *.

(signed)______David Samols, Ph.D.______

(chair of the committee)

______Clemencia Colmenares, Ph.D.______

______Nikki Harter, Ph.D.______

______Lynn Landmesser, Ph.D.______

______Ed Stavnezer, Ph.D.______

(date) _____October 27, 2008______

*We also certify that written approval has been obtained for any proprietary material contained therein.

ii TABLE OF CONTENTS

LIST OF TABLES ix

LIST OF FIGURES xii

ACKNOWLEDGEMENTS xiv

LIST OF ABBREVIATIONS xvi

ABSTRACT 1

INTRODUCTION 3

Structure and Protein-protein Interactions of Ski 3

Transcriptional Regulation by Ski 6

Expression Pattern of Ski 7

Biological Activities of Ski 8

Ski Superfamily 9

Overview of Embryonic Muscle Development 10

Genes Involved in Migration of Myogenic Progenitors 14

MRFs Functions in Muscle Development 15

Muscle Satellite Cells and Pax7 16

Transcriptional Regulation of Muscle-specific by Ski 17

Regulation of Myog Transcription 18

Conserved Network of Dach/Six/Eya 20

Our Research Focus 22

MATERIALS AND METHODS 27

Materials 27

Plasmids 27

iii In situ hybridization probes 29

PCR templates and primers 29

Reagents 35

Common reagents for RNA-related experiments 35

Genotyping 35

Immunohistochemistry 36

Whole mount in situ hybridization 37

Primary culture of satellite cell-derived myoblasts 39

DNA Cloning 40

Cell extract and Western blotting 41

Immunofluorescence 45

Chromatin immunoprecipitation (ChIP) 46

Immunoprecipitation 48

Methods 48

Mouse strains and embryos 48

Genomic DNA isolation 49

Genotyping 49

Histological analysis 50

Immunohistochemistry 51

Construction of plasmids for making RNA probes 52

Preparation of digoxigenin (DIG)-labeled RNA probes 53

Preparation of template DNA 54

In vitro transcription 54

iv Purification of labeled probes 55

Probe evaluation 55

Whole mount in situ hybridization 56

Primary culture of satellite cell-derived myoblasts 57

DNA Cloning 58

PCR amplification 58

Restriction enzyme digestion 59

Generation of blunt-end DNA 60

Dephosphorylation 60

Gel separation and DNA purification from gel slice 60

Ligation 61

Transformation 61

Construction of retroviral vectors 62

Tissue culture and transfection 63

Retroviral infection 64

Western blotting 64

Immunofluorescence 66

Quantitative realtime-PCR 67

Chromatin immunoprecipitation (ChIP) 70

Genomic DNA preparation 72

Reporter assays 73

Construction of luciferase reporters 73

Reporter assay 75

v Immunoprecipitation 76

RESULTS AND DISCUSSION 78

CHAPTER I: Ski Is Required For Normal Development of Embryonic

Progenitors of Mouse Hypaxial Muscle but Not For Fetal

Progenitors or Satellite Cells 78

Abstract 78

Introduction 80

Experimental Procedures 82

Mouse strains and embryos 82

Histological analysis 83

Immunohistochemistry 83

Whole mount in situ hybridization 84

Primary culture of satellite cell-derived myoblasts 85

Quantitative realtime-PCR 85

Results 86

Selective muscular hypoplasia of Ski-/- embryos 86

Dynamic expression of Ski 89

Determination and differentiation of myogenic progenitors

in Ski-/- embryos 92

Altered migration of myogenic progenitors in Ski-/- embryos 102

Normal specification of fetal/postnatal myogenic progenitors

in Ski-/- embryos 110

Discussion 115

vi Future Directions 119

Dissecting the roles of Ski during development 119

Mechanism underlying the regulation of Met by Ski 121

CHAPTER II: Ski Regulates Muscle Terminal Differentiation

by Transcriptional Activation of Myog in a Complex

with Six1 and Eya3 125

Abstract 125

Introduction 126

Materials and Methods 129

Construction of retroviral vectors 129

Tissue culture and transfection 130

Retroviral packaging and infection 130

Western blotting 131

Immunofluorescence 133

Realtime-PCR 134

Chromatin immunoprecipitation (ChIP) 134

Reporter assays 136

Co-immunoprecipitation 137

Results 138

Regulated over-expression of Ski stimulates terminal differentiation 138

Generation of C2C12 clones with inducible knock-down of Ski expression 141

Ski knockdown is reversible and Dox-dose dependent 145

Impaired myotube formation in the absence of Ski 146

vii Loss of Ski inhibits commitment to terminal muscle differentiation 149

Ski occupies Myog regulatory region in differentiating myoblasts 153

MEF3 binding site is required for the activation of Myog regulatory

region by Ski 159

Ski associates with Eya3 and Six1 in differentiating muscle cells 162

The DHD of Ski is required for its association with Six1 and

its activation of Myog transcription 163

Discussion 169

Future Directions 173

Mechanism of Myog transcriptional regulation by Ski/Six1/Eya3 complex 173

Regulatory cascades of Ski during myogenic differentiation 174

The regulation of satellite cell behavior by Ski in vivo 174

DISCUSSION 176

The contradictions in this study 176

REFERENCES 179

viii LIST OF TABLES

Table 1. Expression vectors (including empty vectors)

Table 2. Retroviral vectors for knock-down of Ski

Table 3. Luciferase reporters

Table 4. Plasmids for making in situ probes

Table 5. Probes for whole mount in situ hybridization

Table 6. PCR primers for mouse genotyping:

Table 7. Primers for realtime-PCR

Table 8. PCR primers for generating Met fragments

Table 9. PCR primers for generating tTA fragment

Table 10. PCR primers for generating fragment encoding Eya3 ORF (No 3’UTR)

Table 11. PCR primers for generating huSKI fragment without DHD

Table 12. Synthetic DNA oligonucleotides targeting mouse Ski

Table 13. Common primers for generating DNA forms of shRNA inserts targeting

mouse Ski

Table 14. PCR primers for generating Myog regulatory regions for Myog-luciferase

constructs

Table 15. PCR primers for generating Myog-luciferase mutants

Table 16. PCR primers for ChIP assays

Table 17. DEPC-treated reagents

Table 18-1. DNA isolation buffer-1 (For isolation of DNA from yolk sac)

Table 18-2. DNA isolation buffer-2 (For isolation of DNA from tail)

Table 19-1. Fixation solution (Immunohistochemistry)

ix Table 19-2. Permeabilization solution (Immunohistochemistry)

Table 19-3. Blocking solution (Immunohistochemistry)

Table 20-1. STE buffer

Table 20-2. PBST

Table 20-3. Refixation solution (Whole mount in situ hybridization)

Table 20-4. Hybridization mix (Whole mount in situ hybridization)

Table 20-5. Solution I (Whole mount in situ hybridization)

Table 20-6. Solution II (Whole mount in situ hybridization)

Table 20-7. Solution III (Whole mount in situ hybridization)

Table 20-8. MAB stock (Whole mount in situ hybridization)

Table 20-9. MABT (Whole mount in situ hybridization)

Table 20-10. Blocking Solution (Whole mount in situ hybridization)

Table 20-11. NTMT (Whole mount in situ hybridization)

Table 21. Laminin-coating solution (Primary culture of satellite cell-derived

myoblasts)

Table 22-1. Lysis buffer (For isolate genomic DNA)

Table 22-2. DNA loading buffer

Table 22-3. TAE buffer

Table 22-4. SOB medium

Table 22-5. LB medium

Table 22-6. LB-Ampr plate

Table 23-1. RIPA lysis buffer (Cell extract)

Table 23-2. Protein loading buffer

x Table 23-3. 10% APS

Table 23-4. SDS-PAGE separating gel

Table 23-5. SDS-PAGE stacking gel

Table 22-6. TG-SDS (Gel running buffer)

Table 23-7. Transfer buffer

Table 23-8. TBS

Table 23-9. TBST

Table 23-10. Blocking buffer (For Western blot with most of the antibodies)

Table 23-11. Blocking buffer (For Western blot with G8 antibody)

Table 23-12. Assay buffer (Western blot)

Table 23-13. Stripping buffer

Table 24-1. PBS

Table 24-2. Fixation buffer (Immunofluorescence)

Table 24-3. Permeablization buffer (Immunofluorescence)

Table 24-4. Blocking buffer (Immunofluorescence)

Table 25-1. Lysis buffer (ChIP)

Table 25-2. TE buffer

Table 25-3. Blocking buffer (ChIP)

Table 25-4. Pre-blocked protein A beads (ChIP)

Table 25-5. High salt buffer (ChIP)

Table 25-6. Lithium salt buffer (ChIP)

Table 25-7. Elution buffer (ChIP)

Table 26. NETN buffer (Immunoprecipitation)

xi LIST OF FIGURES

Figure 1. Architecture of Ski Protein

Figure 2. Schematic representation of the genetic mechanisms underlying

during mouse development

Figure 3. Selective muscular hypoplasia in prenatal Ski-/- mice

Figure 4. Hypoplasia of limb muscle in Ski-/- embryos at mid and late gestation

Figure 5. Dynamic expression pattern of Ski during mouse development

Figure 6. Specification and activation of epaxial and hypaxial myogenesis

Figure 7. Normal commitment to terminal muscle differentiation in the forelimbs of Ski-

/- embryos

Figure 8. Normal levels of proliferation and in the myogenic progenitors or

muscle cells in the absence of Ski

Figure 9. Impaired migration of myogenic progenitors in Ski-/- embryos

Figure 10. Loss of Ski affects the expression of Met but not HGF/SF or Msx1

Figure 11. Loss of Ski doesn’t affect in vivo accumulation or in vitro differentiation of

fetal myogenic progenitors

Figure 12. Loss of Ski doesn’t affect in vivo accumulation or in vitro differentiation of

satellite cells

Figure 13. SKI stimulates differentiation of C2C12 myoblasts

Figure 14. Dox-regulated knock-down of Ski in C2C12 cells via shRNAs in the context

of miR30

Figure 15. Loss of Ski prevents terminal differentiation of C2C12 myoblasts

Figure 16. Loss of Ski blocks myogenic differentiation at an early stage

xii Figure 17. Loss of Ski reduces the expression of muscle-specific genes at both mRNA and protein levels.

Figure 18. Ski binds the endogenous Myog regulatory region upon differentiation

Figure 19. The MEF3 binding site is required for maximal activation of Myog promoter by Ski.

Figure 20. Ski interacts with Six1 and Eya3 but not with MyoD or

Figure 21. The DHD of Ski is required for its association with Six1 and its activation of

Myog transcription

xiii ACKNOWLEDGEMENTS

This work would not have been possible without the support and encouragement

of my advisor, Dr. Ed Stavnezer. His caring supervision, his enthusiastic involvement in

this project and his skilful guidance were of essence to this work. As an excellent mentor,

he always encouraged me to learn as much as possible, support me to be independent, and was always available when I needed his help or advice. He helped me come out of many frustrating moments during my PhD research and his enthusiasm about science always inspires me.

I would also like to thank Dr. Clemencia Colmenares for all her help and support as our collaborator and my committee member. She brought me to the world of

Developmental Biology and shared with me her expertise and research insight. The resources she provided were invaluable and greatly appreciated.

I would like to thank all my committee members, Dr. David Samols, Dr. Nikki

Harter, Dr. Clemencia Colmenares, Dr. Stephen O'Gorman and Dr. Lynn Landmesser, for their caring criticism and support of my research. Without their help and support, my research would not have been as complete as it has become. Special acknowledgement goes to Dr. David Samols for all the corrections and revisions he made on the thesis, which have been significantly useful in shaping it up to completion. I would also like to thank Dr. Lynn Landmesser for taking time out from her busy schedule to serve as my thesis examining member on a short notice.

I want to thank all the fellow students, postdoc and technicians in my lab and Dr.

Colmenares’ lab for their technical support and for sharing their experiences and insights.

xiv I would also like to thank all my friends in this department, for their caring and friendship which help me through tough times and enabled me to have a wonderful time along the way.

I have been accompanied and supported by so many people along the way toward this point of my life and it is my great pleasure to take this opportunity to thank all of them. I hope one day I could provide similar help and support for someone else as they strive to achieve their dreams.

Finally, I am greatly indebted to my family: to my parents who provide constant encouragement and love to me. Their dedication to careers and striving to do their best in both professional and personal life always encouraged me. I also like to thank my husband and my best friend, Yun Ding, for his understanding and unconditional support.

The courage, perseverance and sense of humors with which he responds to life’s challenges have given me strength and courage to face my own. It is to them that I dedicate this work.

This work was supported by grants from the National Institutes of Health (DE15198 to C.C and E.S, CA43600 to E.S, and HD30728 to C.C).

xv LIST OF ABBREVIATIONS

A/B Acrylamide/Bisacrylamide

ALP alkaline phosphatase

APS ammonium persulfate

β-Me beta-Mercaptoethanol bFGF basic fibroblast growth factor

BME basal medium eagle

BMP bone morphogenetic proteins

BSA bovine serum albumin

CHCl3 chloroform

ChIP chromatin immunoprecipitation

CIP calf intestinal alkaline phosphatase

Co-IP co-immunoprecipitation

Ct threshold cycle number

DAPI 4’, 6’-diamidino-2-phenylindole

DEA diethanolamine

DEPC diethyl pyrocarbonate

DIG digoxigenin

DM differentiation medium

DMEM Dulbecco’s Modified Eagle Medium

dNTP deoxyribonucleotide triphosphate

DOC deoxycholate

Dox doxycycline

xvi dsDNA double strand DNA

DTT 1, 4-Dithiothreitol dUTP deoxyUridine triphosphate

E18 embryonic stage 18

EDTA ethylenediaminetetraacetic acid

FBS fetal bovine serum

FITC fluorescein isothiocyanate

FLAG flag epitope

G418 geneticin

GM growth medium

GFP green fluorescent protein

H2O2 hydrogen peroxide

HDAC histone deacetylase hr hour

HRP horseradish peroxidase

HS horse serum

IgG immunoglobulin G

MHC myosin heavy chain min minute

MIR30 microRNA 30

NaF sodium fluoride

NP-40 Nonidet P-40

O.C.T. compound optimal cutting temperature compound

xvii ORF open reading frame

PAGE polyacrylamide gel electrophoresis

PBS phosphate-buffered saline

p.c. post-coitum

PFA paraformaldehyde

Puro puromycin

PVP polyvinylpyrrolidone

RLU relative light units

RNAi RNA interference

RT room temperature

RT-PCR real time RT-PCR

SAP shrimp alkaline phosphatase

SDM site directed mutagenesis

SDS sodium dodecyl sulfate sec second shRNA short hairpin RNA

TBS Tris-buffered saline

TEMED N’,N’,N’,N’-Tetramethylethylenediamine

TGF-β transforming growth factor β

TK promoter thymidine kinase promoter tRNA transfer RNA tTA tetracycline-controlled transactivator

U unit

xviii Regulation of Skeletal Muscle Development and Differentiation by Ski

Abstract By

HONG ZHANG

Ski is the most studied member of a family of proteins all sharing a conserved

Dachshund homology domain. It has been implicated in oncogenic transformation, myogenic conversion of avian embryo fibroblasts and also many aspects of vertebrate development, especially myogenesis. Ski-/- mice exhibit severe defects in skeletal muscle and die at birth, yet little is know about either the underlying mechanisms or the role of

Ski in adult muscle regeneration. In these studies, I used Ski knockout mice and C2C12 myoblast cultures to address these issues, respectively. Detailed analysis of Ski-/-

embryos revealed dramatically reduced hypaxial muscles but less affected epaxial

muscles. The reduced number of myogenic regulatory factor positive cells in Ski-/- mice

suggested an insufficient myogenic cell pool to support muscle formation. However, both

the dermomyotomal hypaxial progenitors and myotomal epaxial progenitors formed and

committed to myogenic fate appropriately. The hypaxial muscle defect in Ski-/- mice was

not caused by abnormal proliferation, terminal differentiation or apoptosis of the

myogenic cells either, but due to impaired migration of embryonic hypaxial progenitors.

Surprisingly, the normal distribution of fetal/postnatal myogenic progenitors in Ski-/-

mice suggested different effects of Ski on the behaviors of embryonic and fetal/postnatal

myogenic progenitors. In addition, although not affecting the terminal differentiation of

embryonic myogenic cells, Ski was necessary for that of adult satellite-cell derived

C2C12 myoblasts as evidenced by impaired myotube formation and reduced induction of

genes essential for myogenic differentiation in the absence of Ski. This function was

1 mainly mediated by Ski’s ability to form a complex with Six1 and Eya3 and activate

Myog transcription through a MEF3 site. It is important in the future to further study mechanisms underlying the contrasting effects of Ski on embryonic, fetal and adult muscle development, to investigate how the association of Ski with Six1/Eya3 is triggered upon differentiation and to identify the transcriptional machinery mediates Ski’s action. The data presented here not only add a new aspect to the understanding of myogenic progenitor migration and hypaxial muscle development but also provide a starting point to achieve the regulation of muscle formation and regeneration through the action of Ski.

2 INTRODUCTION

Structure and Protein-protein Interactions of Ski

The v-ski originated in the avian Sloan Kettering retroviruses (1-3). Its cellular homologue c-ski was later found to be an evolutionarily conserved in species ranging from Xenopus, Zebrafish, chicken to mouse and human (reviewed in (4)).

The c-ski protein is a nuclear protein of 750 residues, while v-ski corresponds to its

residues 21–441(5). The N-terminal region of c-Ski protein, encoded by the first of

the gene, is the most highly conserved segment which includes an alanine-rich region

(34-45), a -rich region (53-76), several α helices and β turns (91-175), and several

- rich domains characteristic of zinc-finger like motif (202-295) (Figure

1A) (6-8). Mutational analysis suggests that except for the proline-rich region, these predicted motifs are all important for the biological activities of c-Ski (6). Two distinct domains composed of these motifs have been implicated in mediating important protein- protein interactions of c-Ski. The more conserved one is the Dachshund homology domain (DHD), which defines the Ski gene family including Ski, SnoN, Dach, Fussel-15,

Fussel-18 and Corl (Figure 1B) (7, 9-15).The DHD contains mixed helix turn helix motifs and bears structural features characteristic of the winged-helix/forkhead class of

DNA binding proteins (7). However, in Ski, this domain doesn’t function as a DNA binding domain, but facilitates interactions of Ski with Smad2/3, the nuclear hormone co-repressor N-CoR, Skip and the α (RARα) (16-24). A of this region completely eliminated Ski’s ability to repress transcription and transform chicken embryo fibroblasts (CEFs) (25, 26). The second conserved region is the zinc-finger like motif which is highly homologous to the SAND domain (Sp100,

3

4 Fig. 1. Architecture of Ski protein. (A) A schematic representation of c-Ski and v-Ski

proteins. Regions known to participate in protein-protein interactions are indicated by

colored boxes. Symbols indicate: Ala, alanine-rich region; Pro, proline-rich region;

AH/BT, mixed α helices and β turns; Zn, putative zinc fingers; TR, tandem repeat; LZ,

zipper motif. The locations of these conserved regions within Ski protein are

indicated above the bar in number. Domains required for binding to Ski protein partners

are also indicated in the figure. DHD, Dachshund homology domain; SAND, Sp100,

AIRE-1, NucP41/75 and DEAF-1 domain. Arrowheads indicate protein partners that

interact with Ski. (B) Sequence alignment of SKI-DHD, v-Ski-DHD, Sno-DHD and

DACHbox-N. Identical residues are indicated by a blue background and similar residues

are shown by a yellow background. The secondary structure of Ski-DHD is indicated

above the sequence by blue arrows for β strands and red coils for α helices. T indicates β

turns.

5 AIRE-1, NucP41/75 and DEAF-1 domain)(27). The SAND domain usually mediates

DNA binding, but in Ski, it mediates interaction of Ski with Smad4, Fhl2 and MeCP2 (8,

28, 29).

The C-terminal region of c-ski is missing in v-ski and contains a tandem repeat

and a motif that mediates homo-dimerization and hetero-dimerization with

the closely-related protein, SnoN (6, 30). Although lacking these C-terminal domains, v-

ski can dimerize with c-ski and retain all the biological activity of c-ski.

Transcriptional Regulation by Ski

Although Ski does not bind DNA directly (31), it has been shown to interact with

several transcription factors to modulate transcription as either a co-activator or co-

repressor depending on its DNA-binding partner.

By binding to Smads, Ski negatively regulates transforming growth factor-β

(TGF-β) (21-24) and bone morphogenic protein (BMPs) signaling pathways (28), both of which have been implicated in many aspects of vertebrate development including mesoderm patterning, myogenesis, bone formation, and organogenesis (32) .

Binding of TGF-β or BMPs to their respective receptor serine/threonine kinases

results in phosphorylation of the receptor-regulated Smad proteins (R-Smad). The R-

Smads are Smad2 and Smad3 for TGF-β signaling, and Smad1, Smad5 and Smad8 for

BMP signaling (33). The phosphorylated R-Smads then form complexes with a common

mediator Smad (Co-Smad), Smad4, and translocate into the nucleus. There they bind to

the Smad-binding element (SBE) present in TGF-β/BMP-responsive promoters, interact

with other transcription factors and activate transcription. Through the interaction with

6 Smad2, Smad3 and Smad4, Ski is recruited to DNA in a TGF-β-dependent manner and represses the activation of TGF-β target genes (21-24, 34). Ski also can interact with

Smad4 in a BMP-dependent manner to block BMP signaling (28), thereby influencing

neural specification in the Xenopus and Zebrafish embryos and inhibiting BMP2-induced

osteoblast differentiation in murine bone marrow stromal cells (28, 35, 36).

In addition to the Smad proteins, Ski has also been reported to interact with

several other proteins including N-CoR, mSin3A, RAR, Glioblastoma-3 (Gli3),

(Rb), the methyl-CpG-binding protein MeCP2 and Skip (8, 16-18,

28, 29, 37, 38). These interactions contribute to many aspects of the biological activity of

Ski.

Ski becomes a component of co-repressor complexes by directly binding to N-

coR and mSin3A and these interactions have been implicated in the transcriptional

repression of Mad, Rb and thyroid by Ski (16, 37). When associated

with RAR, Ski recruits N-CoR to repress RAR-induced transcription from retinoic acid

response elements and blocks RAR-induced differentiation of acute promyelocytic

leukemia cells (18). However, N-CoR recruitment is not involved in Smad co-repression

(37). In human cells, Ski activates Wnt/β -catenin signaling through its

interaction with DRAL/FHL2 (8). Ski has also been reported to bind to full length Gli3

and inhibit sonic hedgehog (Shh)-induced Gli3 target gene transcription (38).

Expression Pattern of Ski

Ski is expressed in virtually all adult and embryonic tissues but at low levels (39,

40) and has also been detected in human tumor cell lines derived from melanoma, breast

7 cancer, prostate cancer and carcinomas of the esophagus, thyroid, vulva, lung, stomach

and epidermoid (10, 41). However, the signaling pathways or molecules that regulate the

transcription, translation and stability of Ski during embryogenesis and in adult tissues are

largely unknown.

Ski is expressed dynamically during embryonic development. It is expressed in

the neural tube, migrating cells and dermomyotome at embryonic day 9.5

(E9.5)(39, 40). At E12.5, its mRNA expression peaks in skeletal muscle and drops to a

low level by E15.5. At E14-E16, Ski mRNAs are detected at high levels in the brain (39).

In adult mice, Ski is expressed in brain and lung at the highest levels, in stomach, ovaries, heart, diaphragm and spleen at intermediate levels, and in liver, small intestine, uterus, skeletal muscle and skin at relatively low levels (40).

Biological Activities of Ski

v-ski was discovered by virtue of its ability to cause anchorage-independent growth and morphological transformation in avian embryo fibroblasts(2, 3); a property it

shares with c-ski (25, 42) and is correlated with increased SKI levels in tumor cell lines.

Surprisingly, loss of one copy of Ski increased susceptibility to tumorigenesis in mice

(43). In addition, Ski-deficient mouse embryonic fibroblasts (MEFs) exhibited increased cell death, whereas overexpression of Ski protected the cells from apoptosis (44).

Consistent with its expression pattern during embryonic development, Ski has been implicated in biological processes including muscle formation and craniofacial morphogenesis. Both forms of ski have been shown to induce myogenic differentiation in non-muscle avian embryo fibroblasts by activation of the myogenic regulatory factor

8 (MRF) genes, MyoD and (Myog) (25, 42). In transgenic mice expressing v-ski or c-ski cDNAs, hypertrophy but not hyperplasia of type IIb fast skeletal muscle fibers was observed (45, 46). The involvement of the endogenous gene in muscle development was demonstrated by the observation that Ski-null mice showed a marked decrease in skeletal muscle mass (47). Ski-/- mice also suffered from exencephaly in the 129 background and severe midline facial clefting in the C57BL6/J background (48). In addition, these mice displayed other craniofacial and skeletal abnormalities including facial flattening, a broad forehead, malformation of the eye, a depressed nasal bridge and digit abnormalities. These phenotypes are reminiscent of those found in a human 1p36 deletion syndrome, caused by a hemizygous deletion in the short arm of

1(48). Interestingly, human SKI gene has been mapped to chromosome 1p36 and is included in the minimally deleted region in this syndrome (48). Thus, it is possible that loss of SKI is responsible for some of the observed craniofacial phenotypes in individuals with 1p36 deletion syndrome.

Ski Superfamily

The Ski gene family also includes another member, SnoN (ski related novel oncogene) which shares a high degree of sequence identity with the Ski protein in the N-terminal region(10). SnoN forms heterodimers with Ski (49) and interacts with many proteins which have been shown to interact with Ski, such as Smad2 and Smad3, Smad4 and N-CoR (4, 27). Similar to Ski, SnoN also negatively regulates

TGF-β signaling (43) and overexpression of sno in chicken embryo fibroblasts induced oncogenic transformation and myogenic differentiation, although less effectively than c-

9 ski (50). Despite all of the similarities to Ski, SnoN seems to have a very different role in

embryonic development: one SnoN-null line showed embryonic lethality before E3.5,

while the others only showed a defect in T cell activation (9, 43).

Recently, Ski and SnoN have been classified as part of a superfamily with the

retinal determination proteins, Dach1 and Dach2, based on the conserved DHD domain

(7, 51-54). Drosophila dachshund (dac) is involved in the cell-fate determination in the

eyes, and in muscle and leg development(55-61). Vertebrate Dach1 and Dach2 are also

expressed in the limb, brain and eye(11, 13, 62, 63), but targeted disruption of mouse

Dach1 resulted in postnatal lethality without any obvious abnormality in these major

Dach1-expressing organs(64).

The mammalian Dach and Drosophila dac proteins share two evolutionary conserved regions (Dachbox-N and Dachbox-C) which are implicated in protein-protein

interaction (Figure 1A)(7). The N-termini DHD of vertebrate Ski and SnoN proteins

share 28% identity with human DACHbox-N domain, while the C-terminal

oligomerization domain of Ski and SnoN exhibits much weaker homology to the

DACHbox-C domain. The conserved N-terminal DHD domain may mediate common

protein-protein interactions of these Ski superfamily members which are likely to share

not only structural, but also functional properties.

Overview of Embryonic Muscle Development

During post gastrulation vertebrate development, unsegmented paraxial

mesoderm condenses and segments in a rostral to caudal direction to form epithelial

on both sides of the neural tube (as reviewed in (65-67)). In response to the

10 extrinsic patterning signals from the adjacent epidermis, neural tube and notochord, the somites differentiate into sclerotome ventrally and dermomyotome dorsally. The sclerotome undergoes an epithelial-mesenchymal transition and differentiates further to give rise to vertebrae and ribs. The dermomyotome maintains its epithelial morphology and subdivides into medial and lateral compartments. Cells from the medial dermomyotome detach and ingress underneath the dermomyotome to form the myotome, which will give rise to epaxial muscles, such as deep back muscle. They are the first cells to differentiate into muscle during embryonic development. The lateral lip of the dermomyotome provides the myogenic progenitors of hypaxial muscles, such as muscles of appendicular, ventrolateral bodywall, tongue, and diaphragm. These progenitors are distinguishable by the expression of the paired box genes, Pax3 and Pax7. At the limb and cervical level, after an epithelial-mesenchymal transition, these myogenic progenitors delaminate from the lip of the dermomyotome and undergo long-range migration to form the tongue, diaphragm and appendicular muscles. At the thoracic and abdominal level, dermomyotome retains its epithelial organization and hypaxial myogenic progenitors don’t migrate but extends ventrally toward the midline to form the ventrolateral bodywall muscle (Figure 2). During muscle development in the limb, two distinct populations of myoblasts, embryonic and fetal myoblasts, temporally contribute to two waves of myogenesis, termed primary and secondary myogenesis, respectively(68, 69). Between

E11.5 and E15.5, the primary (embryonic) myoblasts fuse into primary myotubes as a scaffold for secondary myogenesis, while secondary (fetal) myoblasts remain proliferative. From E15.5, secondary myoblasts progressively enter terminal differentiation and form the final patterning of the limb muscle.

11

12 Fig. 2. Schematic representation of the genetic mechanisms underlying myogenesis in the

different myogenic compartments and at different times (adapted from Birchmeier and

Brohmann, 2000). At E9.5, the epaxial myogenic progenitors are committed to differentiation in the myotome (1), while the hypaxial myogenic progenitors start to

release from the dermomyotome (2). Several genes, such as Pax3 and Met, control this

process. Once these progenitors arrive to their targets sites, they are committed to myogenic lineage, undergo differentiation (3) and finally form hypaxial muscles (4).

Symbols indicate: NT, neural tube; DM, dermomyotome; Myo, myotome; FL, forelimb.

13 A fraction of the myogenic progenitors that migrate into muscle-forming regions

do not commit to terminal differentiation (as reviewed as (70, 71)). In late development these cells are distinguished by expression of Pax7 but not Pax3 and some of them serve

as a reservoir of fetal muscle progenitors. Others become satellite cells, so named

because of their location under the basal lamina of muscle fibers, and comprise the

lineage-restricted progenitor cells of post-natal skeletal muscle.

Genes Involved in Migration of Myogenic Progenitors

The delamination and migration of myogenic progenitors from somites are well- studied processes controlled by several genes including Pax3 (72-77), Met (78, 79), its ligand hepatocyte growth factor/scatter factor (HGF/SF) (78, 79) and Lbx1 (80, 81). Pax3

encodes a paired and homeodomain containing that is expressed in

unsegmented paraxial mesoderm first (82) (E8.0) and then throughout the (E9.0)

(83). Later on (after E9.0), its expression is restricted to the dermomyotome and then

concentrated in the verntrallateral part of the dermomyotome where the hypaxial

myogenic progenitors arise (84, 85). Pax3 is required for myogenic progenitor’s survival

and the initiation of migration. Pax3 mutant Splotch mice showed a severe defect in all

migratory myogenic progenitor-derived hypaxial muscles, but the epaxial muscles and

the hypaxial muscles derived from the non-migratory myogenic progenitors were less

affected. Further analysis revealed that this defect resulted from a failure of the myogenic

progenitors to delaminate from the dermomyotome and migrate to their target sites (72-

77). Analysis of Splotch mice also revealed markedly reduced Met expression consistent

with an in vitro study that showed transcriptional regulation of Met by Pax3 in cultured

14 cells (73, 76, 86-88). These studies place Pax3 developmentally upstream of Met. Met

encodes a tyrosine kinase receptor and is expressed in the ventral lip of the

dermomyotome. Activation of Met by its ligand HGF/SF mediates epithelial-

mesenchymal transition of dermomyotome cells to generate migratory myogenic progenitors and guides their subsequent migration into limb buds (78, 79). In distal regions, such as limb buds, BMPs function to maintain the viability and proliferative potential of Pax3-positive migratory progenitors (89). In the null mutants of either Met or

HGF/SF, the myogenic progenitors failed to delaminate and migrate out of the dermomyotome and exhibited muscle defects similar to Splotch mice (78, 79). The Lbx1

gene, which encodes a homeodomain transcription factor and is expressed exclusively in

the migratory hypaxial myogenic progenitors, also acts downstream of Pax3 since its

expression is no longer detected in the Splotch mutant (87). In Lbx1-/- mice, myogenic

progenitors failed to migrate laterally into the limb but only some of the limb muscles

were affected, suggesting that Lbx1 is required for the lateral, but not ventral migration of

myogenic progenitors(80, 81).

MRFs Functions in Muscle Development

Although the migratory myogenic progenitors are destined to form skeletal

muscle, they don’t express myogenic regulatory factor genes (Myf5, MyoD, Myog and

MRF4) (67). Targeted inactivation of these genes in mice demonstrated that they are

essential for the skeletal muscle formation in the mouse and play partially overlapping

roles in determination and differentiation of skeletal muscle (as reviewed in(67, 90-92)).

The epaxial progenitors and non-migratory hypaxial progenitors commit to the myogenic

15 program by turning on Myf5 in the dorsal medial dermomyotome and the myotome at

E8.0, followed by the activation of Myog and MyoD expression at E9.0 and E10.0, respectively(93). During their delamination and migration, the migratory hypaxial progenitors do not express the MRFs, but Pax3 and Pax7 (94). Once they reach their destinations, their determination to the muscle lineage is marked by turning off Pax3/7 and expressing the Myf5 and subsequently Myog and MyoD. Previous studies suggested that Myf5 and MyoD are required for myogenic determination while myogenin and

MRF4 act later to control terminal differentiation. Ablation of either Myf5 or MyoD appears to have little or no effect on the muscle formation(95-97), while MyoD-/-/Myf5-/- double knockout mice lack myoblasts and fail to form skeletal muscle (98), suggesting that MyoD and Myf5 play redundant roles in myogenic lineage determination. Disruption of Myog gene in mice blocked myoblast terminal differentiation and fusion, causing a severe deficiency of skeletal muscle formation (99-101). Targeted disruption of MRF4 in mice only causes a mild muscle phenotype compared to myogenin knockout mice (102).

Moreover, genetic studies indicated that the myogenic determination and/or terminal differentiation of satellite cells are also compromised by the absence of the MRF genes,

Myf5, MyoD or Myog (103-107).

Muscle Satellite Cells and Pax7

Muscle satellite cells are responsible for the postnatal growth, repair and maintenance of skeletal muscle (70). Quiescent satellite cells are located underneath the basal lamina of mature skeletal muscle. After being activated by exercise or injury, they undergo multiple rounds of cell division and either undergo differentiation and fuse with

16 new or existing myofibers or return to satellite cell status. Their differentiation is preceded by turning off Pax7 and upregulating MyoD and Myog, while their quiescent status is supported by Pax7 (as reviewed in (70, 71)).

Pax7 is also a paired and homeodomain containing transcription factor and is expressed in somite and adult satellite cells (108-110). It has been shown that the function of Pax7 partially overlaps with the Pax3 gene to keep myogenic progenitors in a proliferative state during embryonic muscle development. Initially, cell culture and electron microscopic analysis suggested that there was a complete lack of muscle satellite cells and severely compromised regeneration in Pax7-/- mice (111, 112). However, recently more sensitive fluorescence-activated cell sorting analysis revealed that the satellite cells can still be detected in these mice, and it is their maintenance or proliferation that appears to be defective (113). These data suggest a role for Pax7 in satellite cell proliferation and survival, but not specification.

Transcriptional Regulation of Muscle-specific Genes by Ski

Previous studies indicated that the regulation of myogenesis by Ski might be mediated by its ability to activate the regulatory elements of muscle-specific genes, such as myosin light chain 1/3 (MLC1/3), muscle creatine kinase (MCK)(114) and most

importantly, Myog (115, 116). The regulatory regions of both MLC1/3 and MCK contain

potent muscle-specific enhancer elements which are active only in differentiated

myotubes. Detailed analysis of the MLC1/3 and MCK enhancers has identified binding

sites for multiple regulatory proteins, including the myogenic regulator factor, MyoD and

myocyte enhancer factor 2 (MEF2) (114). Stimulation of the MLC1/3 enhancer by Ski

17 was mediated by MyoD. However, the transcriptional activity of the MLC1/3 enhancer was reduced when c-ski was cotransfected with MyoD into non-muscle cells (114).

The ability of Ski to activate endogenous Myog expression almost certainly underlies its ability to induce muscle differentiation in vitro and muscle hypertrophy in transgenic mice, in vivo (25, 42, 45). These findings suggest that Ski plays a role in terminal muscle differentiation; consistent with the observation that during mouse development, Ski expression in skeletal muscle is upregulated after the commitment to myogenic differentiation (39).

Regulation of Myog Transcription

The expression of Myog is strictly controlled, both temporally and spatially, during embryonic development. Studies in transgenic mice have shown that a 133-bp regulatory element immediately upstream of the Myog promoter was sufficient for the complete recapitulation of the temporal and spatial expression pattern of Myog during embryogenesis (117-119). Several binding sites for transcription factors are present in this region, including binding sites for MyoD and MEF2 (119-121). of either site in the myogenin promoter had little effect on LacZ expression, but the mutation of both impaired its expression, suggesting that these two cis-acting elements are important for Myog transcription (117, 122). An in vitro study has shown that the transactivateion of Myog promoter by Ski is mediated through the MyoD/MEF2 complex (115). However, no direct interactions of Ski with these two proteins have been reported.

Recent work has shown that the regulatory elements governing the activation of

Myog are more complex than previously appreciated. Another evolutionarily conserved

18 motif, the MEF3 site (consensus sequence TCAGGTT) is also present in the 133 bp

Myog enhancer and a mutation of this site abolished correct expression of a Myog-LacZ

transgene during embryogenesis (123). The MEF3 site is also found in many other

skeletal muscle-specific regulatory regions and has been shown to mediate the

transcriptional regulation of the muscle-specific cardiac troponin C and aldolase A

promoters both in vitro (124, 125) and in vivo (126, 127). Two skeletal-muscle specific

members of the Six family, Six1 and Six4, bind to the MEF3 element and transactivate

Myog transcription (123).

In vitro studies have revealed that terminal differentiation of myoblasts proceeds

through a highly ordered sequence of events. These cells express MyoD while

proliferating but when growth stimuli are removed, they initiate Myog expression,

followed by the induction of the cyclin-dependent kinase (cdk) inhibitor p21 and

irreversible withdrawal from the cell cycle. Subsequently, these post-mitotic myocytes express muscle-specific contractile proteins such as myosin heavy chain (MHC) and

finally fuse into multinucleated myotubes (128). This process is governed mainly by two families of transcription factors, the MRFs and MEF2, which function in a cooperative manner to activate myogenic-specific (129-131).

All members in MRF gene family including MyoD, Myf5, myogenin and MRF4, share a highly conserved basic region and adjacent helix-loop-helix motif (bHLH) (as reviewed in (91, 92, 129)). These motifs mediate their heterodimerzation with E2A (E12 and E47) and binding to a conserved consensus DNA sequence CANNTG, known as E box. E boxes are present in the regulatory regions of many muscle specific genes. In myoblast cell culture, MyoD and Myf5 function as markers for the committed myogenic

19 state. They are expressed in both proliferating myoblasts and myotube while Myog

expression strictly coincides with the terminal differentiation. Ectopic overexpression of

any one of the MRF family members is capable of inducing expression of muscle-

specific genes and activation of myogenic differentiation, even in non-muscle cells.

The MEF2 family belongs to the superfamily of MADS (MCM1-agamous

deficient-)-box transcription factors and includes MEF2A, MEF2B,

MEF2C and MEF2D (as reviewed in (91, 92, 129)). MEF2A, MEF2B, and MEF2D are

expressed in cardiac muscle, smooth muscle and non-muscle cell types, whereas MEF2C

is restricted to skeletal muscle, brain, and spleen. MEF2 family members share two

highly conserved domains, MADS and MEF2, which mediate DNA binding and

dimerization, respectively. Genetic analysis has demonstrated that members of the MEF2

family are essential for terminal muscle differentiation both in vivo and in vitro (132).

MEF2 proteins can directly bind an A+T-rich element found in the promoters and

enhancers of many muscle specific genes (133). They can also be recruited by bHLH

factors to synergistically regulate the transcription of muscle-specific genes. In addition, the induction of MEF2 by Myog and MyoD during myogenesis suggests that it may function within a regulatory pathway controlled by MRF proteins.

Conserved Network of Dach/Six/Eya

The vertebrate Six family includes sine oculis homeodomain-containing

transcription factors from Six1 to Six6 (134). In vertebrate embryos, Six1 and Six4

expression have been detected in dermomyotome, myotome, limb bud mesenchyme and in migrating myogenic progenitors, suggesting their involvement in muscle development

20 (135, 136). However, Six4-/- mice exhibited no major developmental defects, while Six1-

/- mice had defects in many organs, including craniofacial and muscle deficiencies (135,

136). In addition, Six1-/-Six4-/- mice show an aggravation of phenotype previously

observed in the Six1-/- mice (135). All Six family proteins have two domains in common:

a homeodomain that mediates DNA binding and a SIX-domain that mediates their

interactions with other proteins, such as eyes absent, Eya (134, 137). Six and Eya

function as a complex in a cooperative manner, in which Six mediates DNA binding and

Eya mediates transcriptional activation (138).

There are three members in the mammalian Eya family (Eya1-3) (139-142). The

N-terminal domains of Eya proteins are variable in sequence and mediate transcriptional

activation (139). However, all the Eya proteins share a highly conserved C-terminal domain, known as the EYA domain (55, 61, 137). The EYA domain carries phosphatase activity and also mediates interactions of Eya proteins with the SIX domain of Six proteins (137) and with the Dachbox-C domain of Dach proteins (55, 61). Although Eya has no apparent DNA-binding activity, it can be translocated from the cytoplasm to the nucleus by Six proteins and serves as a co-activator of Six in transcriptional regulation

(138). Eya also can reverse Dach co-repressor function and activate transcription by recruiting other co-activators (58).

Drosophila sine oculis (so), eyes absent (eya) and dac are co-expressed in multiple organs, including eye, somite and muscle (11, 13, 51, 53, 54, 60, 139). Their highly conserved vertebrate homologues Six, Eya and Dach/Ski/SnoN (7, 11, 13) are also co-expressed in these organs (142). Drosophila sine oculis (so) has been shown to act synergistically with eyes absent (eya) and dachshund (dac) by direct protein-protein

21 interactions. Similar interactions underlie the synergism of their mammalian homologues,

Six, Eya and Dach (11, 55-58, 134, 137, 139, 143). This evolutionarily conserved

regulatory network of Eya/Six/Dach has been shown to drive Drosophila eye morphogenesis(55, 57, 60, 137, 144-147) (51) }(11, 13, 54), regulate myogenesis in

chicken somite culture and in the chick limb and activate transcription of reporters

containing the Myog MEF3 site (56, 58).

Our Research Focus

The previous studies of the molecular mechanisms underlying the biological activities of Ski focused on its involvement in tumorigenesis. However, although both in

vivo and in vitro studies clearly implicated Ski in the regulation of myogenesis, our

knowledge is still fragmentary and the underlying mechanisms have not yet been fully

characterized. This study will focus on investigating the molecular mechanisms by which

Ski regulates myogenesis in greater detail.

The major aim of the first part of my study is to identify the cellular and

molecular mechanisms by which the Ski-null mutation leads to the skeletal muscle defects observed in Ski -/- mice.

The previous work in Dr. Colmenares’ lab suggested that Ski-/- mouse embryos exhibited a dramatic reduction of skeletal muscle mass and this defect was not caused by impaired terminal differentiation (47). However, this conclusion was only based on

Northern analyses of RNA prepared from embryos at E13.5~E16.5. Here we will expand our horizon and apply more developmental analyses to investigate the possible mechanisms underlying the regulation of embryonic muscle development by Ski. In

22 general, lineage determination, cell proliferation, cell death and terminal differentiation

all play important roles in organogenesis, including muscle development (89, 148).

Interestingly, Ski has been previously implicated in these biological processes in certain non-muscle cells (25, 42, 44). This led us to hypothesize that Ski might also regulate lineage determination, cell proliferation, cell death or terminal differentiation in muscle cells and that the muscle defect in Ski-null mice might be caused by interference with these processes during muscle development.

A deficiency in myogenic determination at early stages could lead to an insufficient myogenic progenitor pool to undergo differentiation therefore resulting in reduced muscle mass. In the same way, increased apoptosis or defective proliferation of

Ski-/- muscle cells could also cause a dramatic reduction in the number of myogenic progenitors at early stages or myoblasts at later stages. In addition, previous published study only examined terminal differentiation of myoblasts in Ski-/- embryos at

E13.5~E16.5 (47). However, differentiation deficiencies, as either premature or blocked

differentiation, could occur to not only myoblasts at late stages (E11.5~E18.5) but also to

myogenic progenitors at early stages (E9.0~E11.5). Premature differentiation of

myogenic cells occurs at the expense of cell expansion and depletes muscle cell pool,

whereas blocked differentiation results in a lack of myotube formation even with a

sufficient number of myogenic cells. Finally, numerous studies have indicated that

impaired migration of myogenic progenitors to their target sites selectively affects

hypaxial muscle formation (72-81, 87), which led us to consider the possible involvement

of Ski in cell migration.

23 In the present study, muscle formation in Ski-/- embryos was reexamined in greater detail with additional developmental and molecular analyses and molecular markers in order to identify the underlying cause of the muscle defect observed in Ski-/-

embryos. We considered the possibilities that the defect was caused by abnormal fate

determination, cell proliferation, cell survival, terminal muscle differentiation, or

impaired migration of embryonic myogenic progenitors.

The major aim of the second part of my study was to understand not only the role

of Ski in the terminal differentiation of satellite cells-derived C2C12 myoblasts but to

determine how Ski regulates this process.

Although Ski has been implicated in myogenesis by both in vivo and in vitro

studies (25, 42, 45, 47, 48, 149), there are many aspects that have not been investigated or

fully characterized yet. Previous studies only investigated the involvement of endogenous

Ski in differentiation of embryonic myogenic cells, but not that of adult satellite cells

since Ski-/- mice die at birth. However, considering that these satellite cells are the source

for adult muscle regeneration, it is of great interest to determine whether Ski regulates their differentiation. In addition, although overexpression of Ski has been reported to stimulate myogenesis in quail fibroblasts by inducing MyoD and Myog expression (25), the molecular mechanisms underlying the transactivation of these genes by Ski haven’t been fully understood yet. Also very little is known about whether Ski could also control terminal differentiation of determined muscle cells by a similar mechanism. These unresolved questions are all critical if we are to understand the role of Ski in myogenesis.

We hypothesize that Ski regulates the terminal differentiation of satellite cells by regulating the transcription of critical muscle-specific genes. As a transcriptional co-

24 regulator, this activity of Ski is mediated by its interaction with certain DNA-binding partners on the regulatory regions of these target genes. The published study which suggested a cooperation between Ski and MyoD:MEF2 complex only relied on reporter assays and did not demonstrate an interaction between Ski and either of these two proteins (115). Because of that, the possible involvement of MyoD and MEF2 in the transcriptional regulation of Myog by Ski will be re-examined. In addition, recent studies have implicated other cis-elements in the transcriptional regulation of Myog promoter including a MEF3 binding site (117, 119-121, 123, 150). The observation that a Ski family member, Dach, synergized with Six1 and Eya3 to regulate the transcription driven by this cis-element led us to investigate whether Ski also influences this element through a similar cooperation with Six1 and Eya3. Furthermore, since the interaction of mammalian Dach with Six protein is mediated by the evolutionarily conserved DHD motif (55, 151), we tested the possibility that through this conserved domain (7, 11, 13),

Ski interacts with Six and Eya proteins to regulate Myog expression and thereby control commitment of myogenic cells to terminal differentiation (56).

Since Ski-/- mice die prior to birth and techniques to manipulate embryonic myogenic cells are not yet available (47), we were not able to address these issues using the Ski knockout mouse model. Therefore, C2C12 myoblasts were used. C2C12 myoblast is widely-used as an in vitro model to study the behavior of adult satellite cells. These cells were originally obtained through serial passage of myoblasts isolated from the muscles of adult mice shortly after injury and are capable of proliferation and differentiation (152). Inducible gain-of-function and loss-of-function of Ski in C2C12 myoblasts were generated to examine the affect of Ski on terminal differentiation and to

25 identify the downstream direct target genes of Ski. Additionally, the occupancy of Ski on the regulatory region of these genes in the chromatin context was examined and the cis- element(s) and DNA-binding partner(s) which mediate the transcriptional regulation of these genes by Ski were investigated.

26 MATERIALS AND METHODS

1 Materials

1.1 Plasmids

Table 1 Expression vectors (including empty vectors)

Plasmid name Vector (cut) Insert (cut)

LNIT-huSKI LNITX (PmeI) pSHHSKIN1D (BamHIB-ClaIB) Encodes full-length huSKI pCSA pCSA-MyoD (EcoRI) self ligation pCSA-huSKI pCSA (EcoRV) pCDNA3-huSKI (BamHIB-EcoRV) Encodes full-length huSKI

RSV-huSKIEE Encodes full-length c-ski

RSV-delta7EE Deletion of bases 1745~2012 of huSKI

RSV-huSKIDHDEE RSV-SKIEE (AfeI-AscI) PCR product (StuI-AscI) (see Table 11) pCMX-flag-Six1 pCMX-flag (PstIB) pCDNA-Six1 (BamHIB-XhoIB) Encodes full-length mSix1 pCMX-flag-Eya3 pCMX-flag (EcoRI-NheI) PCR product (EcoRI-NheI) (see Table 10) (No 3’UTR) Encodes full-length mEya3 without 3’ UTR pCMX-flag-Mef2c From H-Y. Kao’s lab Encodes full-length Mef2c pCSA-MyoD From M.L. Harter’s lab Encodes full-length MyoD

Table 2 Retroviral vectors for knock-down of Ski

Plasmid name Vector (cut) Insert (cut)

TMP-tTA TMP (EcoRV-BstXI) PCR product (EcoRV-BstXI) (see Table 9)

27 TMP-tTA-mSki1145 TMP-tTA (EcoRI/XhoI) PCR product (EcoRI-XhoI) (see Table 12, 13)

TMP-tTA-mSki1819 TMP-tTA (EcoRI/XhoI) PCR product (EcoRI-XhoI) (see Table 12, 13)

TMP-tTA-mSki1311 TMP-tTA (EcoRI/XhoI) PCR product (EcoRI-XhoI) (see Table 12, 13)

TMP-tTA-mSki977 TMP-tTA (EcoRI/XhoI) PCR product (EcoRI-XhoI) (see Table 12, 13)

Table 3 Luciferase reporters

Myog184 pGL3-Basic (HindIII-SmaI) PCR product (HindIII-StuI) (see Table 14)

Myog1102 pGL3-Basic (HindIII-SacI) PCR product (HindIII-SacI) (see Table 14)

Myog184-MEF2m Myog184 MEF2 site mutant (see Table 15)

Myog184-MEF3m Myog184 MEF3 site mutant (see Table 15)

Myog184E1m Myog184 E1 box mutant (see Table 15)

Myog184E1E2m Myog184E1m E1 and E2 boxes mutant (see Table 15)

Table 4 Plasmids for making in situ hybridization probes

Plasmid name Vector (cut) Insert (cut)

Bluescribe-mMet1837 Bluescribe-Myf5 (HindIII-EcoRI) PCR product (HindIII-EcoRI)

Bluescribe-mMet459 Bluescribe-Myf5 (HindIII-EcoRI) PCR product (HindIII-EcoRI) pVZIIα-mMet167 pVZIIα-MyoD (EcoRI) PCR product(EcoRI)

BluescriptSK-mSkiSE0.35 pBluescript II SK (EcoRV) pmSkiCDS1 (SmaI)

BluescriptSK-mSkiSE0.7 pBluescript II SK (EcoRV) pmSkiCDS1 (SmaI) pVZIIα-mSkiPE0.35 pVZIIα-mMet167 (EcoRI-SmaI) pmSkiCDS1(EcoRI-PvuII) pVZIIα-mSkiSE0.45 pVZIIα-mMet167 (EcoRI-SmaI) pmSkiCDS1 (EcoRI-SmaI)

28 Bluescribe-mHGF07HE Bluescribe-Myf5 (HindIII-EcoRI) pBS-mHGF07 (HindIII-EcoRI) pVZIIα-mHGF07PE PVZIIα-MyoD (PstI-EcoRI) pBS-mHGF07 (PstI-EcoRI)

Bluescribe-mHGF14HE Bluescribe- Myf5(EcoRI-HindIII) pBS-mHGF14 (EcoRI-HindIII)

1. 2 In situ hybridization probes

Table 5 Probes for whole mount in situ hybridization

Probe Plasmid antisense T3/T7 sense T3/T7

MyoD (1342-1785) pVZIIα-MyoD MluI T3 HindIII T7

Myf5 (15-326) Bluescribe-Myf5 HindIII T7 EcoRI T3

Myogenin (250bp) mMYBS1A HindIII T7 EcoRI T3

Pax3 (1156-1675) BluescriptKS-Pax3 HindIII T7 BamHI T3

Ski (969-1320) BluescriptSK-mSkiSE0.35 HindIII T3 EcoRI T7

Ski (263-969) BluescriptSK-mSkiSE0.7 HindIII T3 EcoRI T7

Ski (1773-2123) pVZIIα-mSkiPE0.35 EcoRI T3 HindIII T7

Ski (1320-1773) pVZIIα-mSkiSE0.45 HindIII T7 EcoRI T3

Met (1837-2347) Bluescribe-mMet1837 HindIII T7 EcoRI T3

Met (459-757) Bluescribe-mMet459 HindIII T7 EcoRI T3

Met (167-757) pVZIIα-mMet167 NotI T3 HindIII T7

HGF (871-1098) Bluescribe-mHGF07HE HindIII T7 EcoRI T3

HGF (397-652) pVZIIα-mHGF07PE HindIII T7 EcoRI T3

HGF (40-588) Bluescribe-mHGF14HE EcoRI T3 HindIII T7

Msx1 (300bp) BluescriptSKII-Msx1 EcoRI T7 Asp718 T3

1. 3 PCR templates and primers

Table 6 PCR primers for mouse genotyping

29 PCR primers Sequence

Primer 1 Ski 2980 (forward) 5’-GGGGAGACCATCTCTTGTTTCG-3’

Primer 4 Ski 4258 (reverse) 5’-GACTTTGAGGATCTCCAGCTGG-3’

Old primers (which generate an extra “neo alone” band)

Neo Primer 2 (forward) 5’-GGAGAGGCTATTCGGCTATGAC-3’

Neo Primer 3 (reverse) 5’-CGCATTGCATCAGCCATGATGG-3’

New neo primers

Neo Primer 2 Complement 5’-GTCATAGCCGAATAGCCTCTCC-3’

Neo Primer 3 Complement 5’-CCATCATGGCTGATGCAATGCG-3’

Table 7 Primers for realtime-PCR

Genes Sequence

GAPDH (GenBank NM_008656) Forward: 5'-CATGGCCTTCCGTGTTCCTACC-3'

Reverse: 5'-GATGCCTGCTTCACCACCTTCTT-3'

Myf5 (GenBank NM_008656) Forward: 5'-TATTACAGCCTGCCGGGACA-3'

Reverse: 5'-CTGCTGTTCTTTCGGGACCA-3'

Pax7 (GenBank AF254422) Forward: 5'-CCACCCACCTACAGCACCAC-3'

Reverse: 5'-GCTGTGTGGACAGGCTCACG-3'

Ski (GenBank NM_011385) Forward: 5'-TGCAGTGTCTGCGAGTGAGAAAG-3'

Reverse: 5'-GCAGGGTGGACTGTTCACAACCT-3'

Myog (GenBank NM_031189) Forward: 5'-CCAGTGAATGCAACTCCCACAG-3'

Reverse: 5'-TGGACGTAAGGGAGTGCAGATTG-3' p21 (GenBank NM_007669) Forward: 5'-AGTCAGGCGCAGATCCACAG-3'

Reverse: 5'-ACGGGACCGAAGAGACAACG-3'

MyoD (GenBank M84918) Forward: 5'-CTCTGATGGCATGATGGATTACA-3'

30 Reverse: 5'-CTCGACACAGCCGCACTCTT-3'

Table 8 PCR primers for generating Met in situ hybridization probes

Template cDNA synthesized from total RNA extracted from mouse liver

PCR primers Sequence

167-757 forward 5’-CCCAAGCTTCCTACACGGCCATCATATTT-3’

167-757 reverse 5’-CGAAGGCATGTATGTACTTTATGG-3’

459-757 forward 5’-CCCAAGCTTCATCCAGTCTGAGGTCCACT-3’

459-757 reverse 5’-CGAAGGCATGTATGTACTTTATGG-3’

1837-2347 forward 5’-CCCAAGCTTAAGCGAGAGCACGACAAA-3’

1837-2347 reverse 5’-CACCCACTTCATGCACATCT-3’

Table 9 PCR primers for generating tTA fragment

Template LNITX

Primers Sequence tTA-L 5'-TATATAATATCTTCCTTTGAAAAACACGATGA-3' tTA-R 5'-TATATAGATATCTACCCACCGTACTCGTCAA-3'

Table 10 PCR primers for generating fragment encoding Eya3 ORF for pCMX-flag-Eya3 (No 3’UTR)

Sequence for Mus musculus Eya3 mRNA refer to GenBank NM_010166

Template pCDNA-Eya3

Primers Sequence

Eya3PCR-L 5'-CATGAATTCCAGGAACCAAGAGAACAGACT-3'

31 Eya3PCR-R 5'-CATGCTAGCTCAGAGGAAGTCAAGCTCTAA-3'

Table 11 PCR primers for generating huSKI DHD deletion

Sequence for human SKI mRNA refer to GenBank NM_003036

Template RSV-huSKIEE

Primers Sequence

hkDHD-L 5'- AAAAGGCCTGCAAGAAGGAGCTGG-3'

hkDHD-R 5'- TGCCATAGTCGAATTTCTCC-3'

Table 12 Synthetic DNA oligonucleotides for shRNA targeting mouse Ski

Sequence for Mus musculus Ski mRNA refer to GenBank AF435852

Underlines indicate the sequences of the target sites: sense strands are shown in

lowercase and anti-sense strands are shown in uppercase.

TMP-tTA-mSki1145 oligo 1145 (1145-1163)

TGCTGTTGACAGTGAGCGCGcagtgtctgcgagtgagaaATAGTGAAGCCACAGATG TATTTCTCACTCGCAGACACTGCATGCCTACTGCCTCGGA

TMP-tTA-mSki1819 oligo 1819 (1819-1837)

TGCTGTTGACAGTGAGCGCGcgcaatttgaggaaagagaTTAGTGAAGCCACAGATG TAATCTCTTTCCTCAAATTGCGCTTGCCTACTGCCTCGGA

TMP-tTA-mSki977 oligo 977 (977-995)

TGCTGTTGACAGTGAGCGACcagcttccataagacccaaATAGTGAAGCCACAGATG TATTTGGGTCTTATGGAAGCTGGGTGCCTACTGCCTCGGA

TMP-tTA-mSki1311 oligo 1311 (1311-1329)

TGCTGTTGACAGTGAGCGacgggcaccagagcctcttactTAGTGAAGCCACAGATGT

32 AAGTAAGAGGCTCTGGTGCCCGGTGCCTACTGCCTCGGA

Table 13 Common Primers for generating DNA forms of shRNA inserts

targeting mouse Ski

Primers Sequence

5’ primer for MIR30 5'-CAGAGGCTCGAGAAGGTATATTGCTGTTGACAGTGAGCG-3'

3’ primer for MIR30 5'-CGCGGCGAATTCCGAGGCAGTAGGCA-3'

Table 14 PCR primers for generating Myog regulatory regions for Myog-

luciferase constructs (Myog184 and Myog1102)

Template: genomic DNA from C57BL/6 mouse tail

Sequence for Mus musculus Myog gene refer to GenBank M95800

DNA fragments containing Myog regulatory region (-1102~+18)

Primers Sequence

Myog1102CR-L 5'-CAAGGAACTGAAGGGGTCTG-3'

Myog1102CR-R 5'-GCTTGTTCCTGCCACTGG-3'

DNA fragments containing Myog regulatory region (-184~+18)

Primers Sequence

Myog184CR-L 5'-AGAAATGAAAACTAATCAAATTACAG-3'

Myog184CR-R 5'-GCTTGTTCCTGCCACTGG-3'

Table 15 PCR primers for generating Myog-luciferase mutants

Note: mutated bases are shown in lowercases.

33 Myog184-MEF2m

Template Myog184

Primers Sequence

MEF2(SDM)F 5'-TCAGGTTTCTGTGGCGTTGGCgggATTTATCTCTGGGTTCATGCC-3'

MEF2(SDM)R 5'-GGCATGAACCCAGAGATAAATcccGCCAACGCCACAGAAACCTGA-3'

Myog184-MEF3m

Template Myog184

Primers Sequence

MEF3(SDM)F1 5'-CAACAGCTTAGAGGGGGGCTCgaGggTCTGTGGCGTTGGCTATATTT-3'

MEF3(SDM)R1 5'-AAATATAGCCAACGCCACAGAccCtcGAGCCCCCCTCTAAGCTGTTG-3'

Myog184-E1m

Template Myog184

Primers Sequence

MyoG-E1-F2 5'-GTTTAAATGGCACCCAGtgGTaGGTGTGAGGGGCTGCGG-3'

MyoG-E1-R2 5'-CCGCAGCCCCTCACACCtACcaCTGGGTGCCATTTAAAC-3'

Myog184-E1E2m

Template Myog184-E1m

Primers Sequence

E2(SDM)F 5'-AAAGGAGAGGGAAGGGGAATtgCAaGTAATCCACTGGAAACGTCT-3'

E2(SDM)R 5'-AGACGTTTCCAGTGGATTACtTGcaATTCCCCTTCCCTCTCCTTT-3'

Table 16 PCR primers for ChIP assays

34 Sequence

Mouse Myog regulatory region (-69~+39) Forward: 5′-GGGCAAAAGGAGAGGGAAG-3′ (GenBank M95800)

Reverse: 5′-AGTGGCAGGAACAAGCCTT-3′

Non-promoter region (+1943~+2185 Forward: 5′-GTCAAGAACTGACTTAAGGCC-3′ downstream of the mouse Myog gene) (GenBank NW_001030662.1) Reverse: 5′-GACACTAGGAGAAGGGTGGAG-3′

Mouse Smad7 regulatory (-274~-142) Forward: 5’-TAGAAACCCGATCTGTTGTTTGCG-3’ (GenBank NW_001030635.1)

Reverse: 5’-CCTCTGCTCGGCTGGTTCCACTGC-3’

1. 4 Reagents

1.4.1 Common reagents for RNA-related experiments

Table 17 DEPC-treated reagents

DEPC-treated H2O Add 1ml DEPC to 1000ml of Milli-Q H2O, mix well and let it set

at room temperature overnight. The next day, autoclave for 30min

and store at room temperature.

DEPC-treated PBS Add 1ml DEPC to 1000ml of 1×PBS, mix well and let it set at

room temperature overnight. The next day, autoclave for 30min

and store at room temperature.

1.4.2 Genotyping

Table 18-1 DNA isolation buffer-1 (For isolation of DNA from yolk sac)

KCl 50mM

35 Tris-Cl (pH8.3) 10mM

MgCl2 2.0mM

Gelatin 0.1mg/ml

NP-40 0.45%

Tween-20 0.45%

Proteinase K 500µg/ml

Table 18-2 DNA isolation buffer-2 (For isolation of DNA from tail)

Tris-Cl (pH 7.5) 100mM

EDTA (pH 8.0) 5mM

SDS 0.2%

NaCl 200mM

Proteinase K 0.2μg/ml

1.4.3 Immunohistochemistry

Table 19-1 Fixation solution

PBS 1×

PFA 4%

Table 19-2 Permeabilization solution

PBS 1×

Triton X-100 0.1%

Table 19-3 Blocking solution

36 PBS 1×

BSA 1%

Heat inactivated goat serum 3%

Tween 20 0.01%

1.4.4 Whole mount in situ hybridization

Table 20-1 STE buffer

Tris-Cl (pH8.0) 10mM

NaCl 10mM

EDTA (pH 8.0) 1mM

Add DEPC-treated H2O

Table 20-2 PBST

PBS 1×

Tween 20 0.1%

Table 20-3 Refixation solution

PBST 1×

PFA 4%

Glutaraldehyde 0.2%

Table 20-4 Hybridization mix

Formamide 50%

37 20×SSC (pH4.5 with citric acid) 5×

SDS 1%

Yeast tRNA (20mg/ml) 50µg/ml

Heparin (50mg/ml) 50µg/ml

Table 20-5 Solution I

Formamide 50%

20×SSC (pH4.5 with citric acid) 4×

SDS 1%

Table 20-6 Solution II

Tris-Cl (pH7.5) 0.01M

NaCl 0.5M

Tween 20 0.01%

Table 20-7 Solution III

Formamide 50%

20×SSC (pH4.5 with citric acid) 2×

Table 20-8 MAB stock (2×)

Maleic Acid 200mM

NaCl 300mM

Ajust pH to 7.5 using NaOH

38 Table 20-9 MABT

MAB stock (2×) 1×

Tween 0.1%

Levamisole 2mM

Table 20-10 Blocking solution

MAB stock (2×) 1×

Tween 0.1%

Levamisole 2mM

BMB blocking reagent (Roche) 2%

Incubate at 65°C for 30min to let the blocking reagent dissolve completely and adjust to pH7.5 with NaOH. Cool to room temperature and add heat-inactivated goat serum to

20% final concentration.

Table 20-11 NTMT

Tris-Cl (pH9.5) 100mM

MgCl2 50mM

NaCl 100mM

Tween 20 0.1%

Levamisole 2mM

1.4.5 Primary culture of satellite cell-derived myoblasts

Table 21 Laminin-coating solution

Tris-Cl (pH7.2) 50mM

39 NaCl 0.15M

Laminin 1µg/cm2

1.4.6 DNA Cloning

Table 22-1 Lysis buffer (For isolation of genomic DNA)

Tris-Cl (pH8.0) 100mM

NaCl 200mM

EDTA (pH8.0) 10mM

SDS 0.5%

Proteinase K 0.1mg/ml

Table 22-2 DNA loading buffer (10×) (For agarose gels)

Glycerol 50%

EDTA (pH8.0) 100mM

SDS 1%

Bromphenol Blue 0.4%

Xylene cyanol 0.4%

Table 22-3 TAE buffer (10×)

Tris 400mM

Acetate 400mM

EDTA (pH8.0) 10mM

Table 22-4 SOB medium

40 BACTO-Tryptone 20g

BACTO-Yeast Extract 5.5g

1M NaCl 10ml

1M KCl 10ml

Add H2O to 1L, autoclave

2+ Add 10ml of 2M Mg (1M MgCl2 and 1M

MgSO4) and 10 ml 2M Glucose before use

Table 22-5 LB medium

BACTO-Tryptone 10g

BACTO-Yeast Extract 5g

NaCl 10g

Add H2O to 1L, autoclave

Table 22-6 LB-Ampr plate

BACTO-Tryptone 10g

BACTO-Yeast Extract 5g

NaCl 10g

Add H2O to 1L, adjust pH to 7.0 with 2N NaOH and add 20g Agar

After autoclave, cool to 55°C and add Ampicillin to 100μg/ml

1.4.7 Cell extracts and Western blotting

Table 23-1 RIPA lysis buffer

41 Tris-Cl (pH 8.0) 50mM

NaCl 100mM

EDTA (pH8.0) 1mM

Glycerol 10%

NP-40 0.2%

Sodium orthovanadate 0.1mM

Sodium pyrophosphate 5mM

NaF 1mM

Complete protease inhibitor cocktails (Roche)

Table 23-2 Protein loading buffer (2×)

1M Tris-Cl (pH6.8) 1ml

Glycerol 2ml

10% SDS 4ml

1% Bromophenol blue 2ml

Add H2O to 10ml

Add 2% β-Me before use

Table 23-3 10% APS

0.1g ammonium persulfate (APS)

Add H2O to 1.0 ml

Ideally use fresh. Can store at -20°C

42 Table 23-4 SDS-PAGE separating gel (6%, 20ml)

H2O 8.2ml

1M Tris-Cl (pH8.9) 7.6ml

10% SDS 0.2ml

29:1 A/B 4ml

10% APS 0.2ml

TEMED 20µl

Table 23-5 SDS-PAGE stacking gel (4%, 5ml)

H2O 7.4ml

1M Tris-Cl (pH6.8) 1.3ml

10% SDS 0.1ml

29:1 A/B 1.3ml

10% APS 0.1ml

TEMED 10µl

Table 23-6 TG-SDS (Gel running buffer, 10×)

Tris 30g

Glycine 144g

SDS 10g

Add H2O to 1L

Table 23-7 Transfer buffer (10×)

43 Tris 60.6g (25mM)

Glycine 242.4g (192mM)

Add H2O to 4L

Before use, dilute 10× Transfer buffer to 1× and add Methanol to 12.5% final concentration

Table 23-8 TBS (10×)

Tris 24.2g

NaCl 80g

Add H2O to 1L and adjust pH to 7.6

Table 23-9 TBST

TBS 1×

Tween 20 0.1%

Table 23-10 Blocking buffer (For Western blot with most of the antibodies)

TBS 1×

Tween 20 0.1%

Nonfat dry milk 5%

Table 23-11 Blocking buffer (For Western blot with anti-Ski G8 antibody)

PBS 1×

Casein (technical grade) 0.4%

44 PVP (MW 40,000) 1%

EDTA (pH8.0) 10mM

Tween 0.2%

Dissolve at 65°C for 2hr and cool to about 37°C

Add Sodium azide to 0.0.2% and adjust pH to 7.4

Table 23-12 Assay buffer

DEA 100mM

MgCl2 1mM

Sodium azide 0.02%

Adjust pH to 10.0

Table 23-13 Stripping buffer

Tris-Cl (pH 6.8) 62.5mM

SDS 2%

β-Me 100mM

1.4.8 Immunofluorescence

Table 24-1 PBS (10×)

Na2HPO4 (anhydrous) 10.9g

NaH2PO4 (anhydrous) 3.2g

NaCl 90g

Add H2O to 1L and adjust pH to 7.2

45 Table 24-2 Fixation buffer

PFA 3.7%

PBS 1×

Table 24-3 Permeablization buffer

PBS 1×

Goat serum 10%

Triton-X 100 1%

Table 24-4 Blocking buffer

PBS 1×

Goat serum 10%

Tween 20 0.1%

1.4.9 Chromatin immunoprecipitation (ChIP)

Table 25-1 Lysis buffer

Tris-Cl (pH8.0) 50mM

EDTA 2mM

NaCl 150mM

Triton X-100 1%

DOC 0.1%

SDS 0.1%

Table 25-2 TE buffer

46 Tris-Cl (pH8.0) 10mM

EDTA 1mM

Table 25-3 Blocking buffer

Salmon testes DNA 1mg/ml

BSA 10mg/ml sodium azide 0.05%

TE 1×

Table 25-4 Pre-blocked protein A beads

Wash 50% Slurry protein A beads (Repligen) twice with cold PBS and TE buffer.

Resuspend beads in blocking buffer to 33% slurry and incubate for at least one day at 4°C

Table 25-5 High salt buffer

Tris-Cl (pH8.0) 50mM

EDTA (pH8.0) 2mM

NaCl 500mM

Triton X-100 1%

DOC 0.1%

SDS 0.1%

Table 25-6 Lithium salt buffer

47 Tris-Cl (pH8.0) 20mM

EDTA 1mM

LiCl 250mM

NP-40 0.5%

DOC 0.5%

Table 25-7 Elution buffer

Tris-Cl (pH8.0) 10mM

EDTA (pH8.0) 5mM

SDS 1%

1.4.10 Immunoprecipitation

Table 26 NETN buffer

Tris-Cl (pH8.0) 20mM

NaCl 100mM

EDTA (pH8.0) 1mM

NP-40 0.1%

Glycerol 10%

DTT 1mM complete protease inhibitors cocktails (Roche)

2 Methods

2. 1 Mouse strains and embryos

48 Generation of C57BL/6 mice with a null mutation of Ski has been described (47,

48). Embryos obtained by heterozygous mating were staged by counting the appearance

of the vaginal plug as day 0.5p.c. and in some cases by determining the number of

somites. Yolk sac or tail DNA was isolated and genotyped by PCR.

2. 2 Genomic DNA isolation

For embryos at E9.0~E12.5, yolk sac were dissected and rinsed thoroughly in

PBS, placed in appropriate amount of DNA isolation buffer-1 (E10.0~E10.5=50µl;

E11.0~E11.5=100µl; E12.0~E12.5=150µl) and incubated overnight at 55°C. The lysate was cleaned by centrifuge at 6,000rpm for 30sec and the supernatant was used for genotyping.

For embryos older than E12.5, mouse tail were dissected and rinsed thoroughly in

PBS, placed into appropriate amount of DNA isolation buffer-2 (E12.5~E15.5=300µl;

E15.5~E18.5=500µl) and incubated overnight at 55°C. The lysate was cleaned by centrifuge at 6,000rpm for 30sec and the supernatant was used for genotyping.

2. 3 Genotyping

PCR reactions using Platinum Taq DNA polymerase (Invitrogen) was set up as follows:

PCR Buffer 1×

MgCl2 1.5mM

dNTPs 200µM each

PCR enhancer 1×

49 Platinum Taq 2.5U

Primers 1~4 200nM each

Genomic DNA 1.5µl

Add sterile Milli-Q H2O to 50µl

The PCR program was:

Step 1 (initial melting) 94°C =4min

Step 2 (amplification) 94°C =30sec

60°C =1min

72°C =1min 30sec

Repeated for 32 cycles

Step 3 (final extension) 72°C =9min

PCR reactions with DNA isolated from Ski+/+, Ski+/- and Ski-/- embryos were used as positive controls and that with H2O only was used as negative control. PCR products were resolved in a 1.0% agrose gel and visualized by ethidium bromide- staining.

2. 4 Histological analysis

Dissected embryos were fixed overnight at 4°C with 3.7% formaldehyde in PBS, embedded in paraffin and sectioned at 12µm. Sections were stained with hematoxylin/eosin and photographed under bright-field illumination with an Olympus

BX50 microscope equipped with a Polaroid PDMC2 digital camera and software. Figures were assembled using Photoshop CS (Adobe).

50 2. 5 Immunohistochemistry

Embryos were dissected in PBS and fixed with 4% paraformaldehyde (PFA) in

PBS at 4°C for varying times depending on the stage. The times were E9.5=1hr;

E10.5=2hr; E11.5=3hr; E12.5=4hr; E13.5~E15.5=overnight and E16.5~E18.5=overnight

(cut into three pieces). The fixed embryos were equilibrated in PBS with increasing concentrations of sucrose (10%, 20% and 30%) at 4°C, then 1:1 30%sucrose/OCT mixture for 1hr at room temperature with gentle rocking and finally embedded in OCT compound (Tissue-Tek, Miles Inc., Elkhart, Illinois, USA). The embryos were positioned as desired at the bottom center of embedding molds (Polysciences) and the molds were quickly placed in pulverized dry ice until the OCT compound was completely frozen.

Embedded samples were stored at -80°C until ready for sectioning. Cryostat sections were cut at 10μm, mounted on Superfrost plus slides (Fisher) and air dried overnight.

Immunofluorescence on equivalent sections from sibling Ski-/- and control embryos was performed as described previously (153). Sections were washed three times in PBS, permeabilized in 0.1%Triton X-100/PBS for 20min and washed again in PBS.

Afterwards, sections were incubated in blocking solution for 10min and with primary antibodies in blocking solution overnight at 4°C. After another three washes in PBS, sections were incubated with secondary antibodies in blocking solution for 45min at RT, washed again in PBS and mounted in aqueous Vectashield (with DAPI).

For immunofluorescence, the following primary antibodies at the indicated dilutions were used: Pax3 (monoclonal, 1/200; Developmental Studies Hybridoma Bank),

Pax3 (rabbit polyclonal, 1/200; kindly provided by Martyn D. Goulding, The Salk

Institute) (80), Lbx1 (polyclonal, 1/200; kindly provided by Martyn D. Goulding, The

51 Salk Institute, CA) (80), Myogenin (F5D, monoclonal, 1/200; Santa Cruz), Myosin heavy

chain (MF20,monoclonal, 1/400; Developmental Studies Hybridoma Bank), Desmin

(D33, monoclonal, 1/400; Dako), phospho-Histone H3(Ser 10) (rabbit polyclonal, 1/200;

Cell Signaling), activated Caspase-3 (Asp 175) (rabbit polyclonal, 1/200; Cell Signaling)

and Pax7 (monoclonal, 1/200; Developmental Studies Hybridoma Bank). Secondary

antibodies were AlexaFluor 488 and 594 (1/300, Molecular Probes). Fluorescent images

were obtained with Olympus BX50 microscope equipped with a PDMC2 Polaroid digital

camera and software. Figures were assembled using Photoshop CS (Adobe).

Experiments were performed on at least 6 sections from 4 or more normal and

mutant siblings with each antibody. Statistical significance (set at P<0.05) of differences

between mutants and controls was assessed using Student’s test.

2. 6 Construction of plasmids for making RNA probes

To generate plasmids for making Met probes, total RNA was extracted from mouse liver using an RNAeasy kit (Qiagen) and cDNA was generated using reverse

transcriptase SuperScript™ III (Invitrogen) with random hexamer primers according to the manufacturer’s instructions. PCR reactions using Pfx DNA polymerase to amplify

Met fragments were set up as following:

Pfx PCR Buffer 1×

dNTPs 300µM each

MgSO4 1mM

Primers 300nM each

cDNA 4µl

52 Pfx (5U/µl) 1.25U

PCR enhancer 0.5×

Add sterile Milli-Q H2O to 50µl

The PCR program was set up as following:

Step 1 (initial melting) 94°C =5min

Step 2 (amplification) 94°C =15sec

55°C =30 sec

68°C =1min 15sec

Repeated for 30 cycles

Step 3 (final extension) 68°C =7min

PCR products were digested with restriction enzymes and cloned into vectors as described in section 2.

To generate plasmids for making Ski probes, fragments containing different regions of the Ski gene were generated by digestion of pmSkiCDS1 with restriction enzymes and inserted into appropriate vectors (see Table 4).

To generate plasmids for making HGF/SF probes, fragments containing different regions of the HGF/SF gene were generated by digestion of pBS-mHGF07 and pBS- mHGF14 with restriction enzymes and inserted into appropriate vectors (see Table 4 and

8).

2. 7 Preparation of digoxigenin (DIG)-labeled RNA probes

53 DIG-labeled RNA probes were prepared from linearized plasmids with DIG RNA

Labeling Mixture (Roche) and T3 (Ambion) or T7 RNA polymerase (Roche) according

to the instructions provided by the manufacturers.

2.7.1 Preparation of template DNA

~20µg DNA plasmid was linearized by digestion overnight at 37°C with a

restriction enzyme that leaves a 5’ overhang. Afterwards, the digestion products were resolved in a 0.8% agrose gel to make sure the digestion was complete. Digested DNA was purified by phenol chloroform extraction and ethanol precipitation. Digested DNA was diluted in RNase-free H2O (Qiagen) to a final volume of 200µl and mixed with 1

volume (200µl) of Phenol/CHCl3/soamyl alcohol (USB). After centrifugation at

12,500rpm, 4°C for 10min, the top aqueous phase was mixed with 1/10 volume of

RNase-free 3M sodium acetate and 2 volume of cold 100% ethanol and kept at -80°C for

15min. DNA was pelleted by spinning at the 12,500rpm, 4°C for 15min and was washed with cold 70% ethanol. Finally, the DNA pellet was air dried and dissovled in ~25µl

RNase-free H2O.

2.7.2 In vitro transcription

In vitro transcription was set up as following:

Linearized plasmid 1µg

Transcription Buffer (with DTT) (Ambion) 1×

DIG RNA Labeling Mix 1×

RNase OUT inhibitor (40U/µl)(invitrogen) 1µl

T7/T3 RNA Polymerase (20U/µl) 2µl

DEPC-treated H2O to 20µl

54 The reaction mixture was incubated for 2hr at 37°C and treated with 1µl of DNase

I (Ambion, 2U/µl) for 30min at 37°C.

Probes used in this study are for Pax3 (75); Msx1 (154); Myog (155), MyoD (155) and Myf5 (156). Met probes were subcloned fragments of bases 167-757 and 1837-2347 of murine Met cDNA (88). Mouse Ski probes were four subclones of bases 263-969, 969-

1320, 1320-1773 and 1773-2123 of Ski cDNA (NM_011385). HGF/SF (NM_010427) probes were three separate clones of bases 322-578, 871-1068 and 1837-2347. All antisense riboprobes revealed reproducible hybridization patterns whereas their corresponding sense probes showed no significant signals.

2.7.3 Purification of labeled RNA probes

Free nucleotide in the DIG-labled probes was romoved using a Spin Column

(Amasham ProbeQuant G-50 column). A ProbeQuant G50 column was vortexed and the bottom closure was snapped off. The cap was loosen ¼ turn and the column was centrifuged into an empty RNase-free Eppendorf tube at 3000 rpm for 1min. The in vitro transcription products were diluted to 50µl by adding DEPC-treated STE buffer, added to the center of the rinse bed and collected in a new RNase-free Eppendorf tube by centrifugation at 3000 rpm for 2 min. Probes were stored at -80 °C.

2.7.4 RNA Probe evaluation

2µl of 1/10 and 1/100 dilutions of the newly synthesized probe and of a set of serial diluted DIG-labeled control RNA (Roche, 100ng/µl) (1/10~1/105) were spotted on a nylon hybridization membrane (Hybond). After the spot dried completely, probes were cross-linked twice onto the membrane at UV 245nm. The membrane was incubated with blocking solution for 15min and then with alkaline phosphatase-conjugated anti-

55 digoxigenin antibody (ALP-conjugated anti-DIG antibody, 1/5000 dilution in blocking

solution) for 30min with gentle rocking. After washes in MABT, the membrane was

equilibrated in NTMT for 2min and incubated with BM purple (Roche) for 10min in the

dark to develop the signal. The signals from the prepared probes were compared to those

of standards to estimate the concentration.

2. 8 Whole mount in situ hybridization

Sibling embryos of different genotypes, but similar size, were analyzed in

parallel. Whole mount in situ hybridization was performed according to a protocol

described by Joyner (157). Embryos were dissected in DEPC-treated PBS, fixed in 4%

PFA overnight at 4°C, washed twice in PBS, dehydrated through 25%, 50%, 75%, and

100% methanol series in PBST (5min each) and stored at -20°C in 100% methanol.

Prior to use, embryos were rehydrated through the reversed methanol series back

to PBST. Embryos were bleached in 6% H2O2/PBST for 30 min at room temperature,

washed three times in PBST, treated with 1µg/ml or 10µg/ml proteinase K in PBST at room temperature for varying time, depending on the stage. The times and proteinase K concentrations used here were E7.5=1µg/ml for 1min; E8.5=1µg/ml for 2min;

E9.5=10µg/ml for 2~3min; E10.5=10µg/ml for 5min; E11.5=10µg/ml for 8min and for each additional day of development, incubation with 10µg/ml proteinase K for 1min was added. The proteinase K treatment was stopped by incubation with 2mg/ml glycine in

PBST for 10min. Embryos were washed three times in PBST and refixed in 4%

PFA/0.2% glutaraldehyde/PBST for 20min at room temperature. After another two washes in PBST, embryos were equilibrated in 1:1 PBST/hybridization mix and then

56 hybridization mix. Afterwards, embryos were incubated with hybridization mix for

20min at room temperature and then for at least 1hr at 70°C. Hybridized was carried out

with 1µg/ml DIG-labeled RNA probe overnight at 70°C. The next day, embryos were

washed through Solution I (70°C, 2×30min), 1:1 SolutionI/Solution II (70°C, 10min),

Solution II (room temperature, 3×5min), 100µg/ml RNase A in solution II (37°C, 30min),

Solution II (room temperature, 5min) and finally Solution III (65°C, 2×30min).

Afterwards, embryos were washed twice in MABT buffer at room temperature, incubated

with blocking solution for at least 3hr at RT and with ALP-conjugated anti-DIG antibody

(1/2000 dilution in blocking solution) overnight at 4°C. Embryos were then washed five

times in MABT 1hr each at room temperature and overnight at 4°C. The next day,

embryos were incubated in NTMT for 30min and stained with BM purple plus 2mM

Levamisole in the dark. After the signals were fully developed, stained embryos were

refixed in 4% PFA/PBST overnight. Stained embryos were dehydrated through PBST,

25%, 50%, 75% and 100% ethanol series and photographed under Zeiss KL1500 LCD

illumination, using a Zeiss stemi SV11 stereomicroscope equipped with a Zeiss AxioCam

color digital camera and Zeiss Axiovision software version 2.0. Afterwards, embryos were rehydrated back to PBST and stored in 70% glycerol/PBST at 4°C. Figures were assembled using Photoshop CS (Adobe). Experiments were performed on 3 or more normal and mutant siblings with each probe.

2. 9 Primary culture of satellite cell-derived myoblasts

Satellite cell-derived myoblasts were isolated from pooled limb muscles of E15.5 or E18.5 embryos by enzymatic dissociation (152) with minor modifications. Limb

57 muscles were dissected and the skin and bone were removed. Muscle tissues were

digested in 2:1 0.05% trypsin-EDTA (GIBCO-BRL) and 100U/ml collagenase III

(GIBCO-BRL) in Hank’s BSS for 4hr (E15.5) or overnight (E18.5) at 4°C. The next day,

most of the enzyme solution was removed, leaving the tissue in ~300µl enzyme solution

at 37°C for 10min (E15.5) or 20min (E18.5). The tissues were further suspended in

serum-free medium and separated into uniform cell suspension by vigorous pipetting.

After the tissue debris was removed by filtration through a cell filter (GIBCO-BRL), the

cell suspension was preplated on standard culture dishes for 1hr to get rid of the fibroblasts. The unattached myoblasts were counted and plated on culture dishes coated

with 1µg/cm2 laminin (GIBCO-BRL) at room temperature for at least 1hr.

Growth medium (GM) was Dulbecco's modified Eagle's medium (DMEM) plus

20% fetal bovine serum (FBS), 0.1mM BME, 10ng/ml human recombinant basic fibroblast growth factor (GIBCO BRL), and 2µg/ml gentamycin (GIBCO BRL).

Differentiation medium (DM) was DMEM plus 2% heat-inactivated horse serum and

2µg/ml gentamycin. GM was changed every 12hr and DM every second day.

Cells were cultured to 60% confluence in GM and switched into DM for the indicated time periods. Cells were fixed in 4% PFA and differentiated myoblasts were detected with anti-MHC (MF20, 1/100) and FITC-conjugated secondary antibodies.

Nuclei were stained with DAPI. Images were taken from representative areas and about

200 total cells/area were counted.

2. 10 DNA Cloning

2.10.1 PCR amplification

58 To generate DNA fragments for subcloning, PCR reactions using Platinum Pfx DNA polymerase (Invitrogen) were set up as follows:

Pfx PCR Buffer 1×

dNTPs 300µM each

MgSO4 1mM

Primers 300nM each

Plasmid DNA 300ng

Pfx DNA polymerase 1.25U

PCR enhancer 0.5×

Add sterile Milli-Q H2O to 50µl

PCR program was:

Step 1 (initial melting) 94°C =2min

Step 2 (amplification) 94°C =15sec

55°C =30sec

68°C=2min

Repeated for 30 cycles

Step 3 (final extension) 68°C =7min

2.10.2 Restriction enzyme digestion

To generate the vectors and inserts for cloning, DNA was digested as follows:

Plasmid DNA/PCR product 1~5µg

NEB buffer 1×

BSA (if needed) 1×

Restriction enzyme (NEB) 30 U

59 Add sterile Milli-Q H2O to 30µl

The reaction was incubated at the appropriate temperature for 2hr.

2.10.3 Generation of blunt-end DNA

To generate blunt ends, the enzyme-digested DNA was subjected to the following reaction:

Enzyme-digested DNA 1µg

T4 DNA polymerase buffer 1×

BSA 50µg/ml

dNTP 200µM each

T4 DNA polymerase (NEB) 1U

The reaction was incubated at 12°C for 15min, resolved in 0.8% agarose gel and the product was purified by a gel purification kit (Qiagen).

2.10.4 Dephosphorylation

To avoid self-ligation, enzyme-digested vector with two compatible ends was subjected to shrimp alkaline phosphatase (SAP, Roche) treatment as follows:

Enzyme-digested DNA (3~5kb) 1µg

SAP buffer 1×

SAP enzyme 1U

The reaction was incubated at 37°C for 10min (sticky end) or 60min (blunt end).

Afterward, the enzyme was heat-inactivated at 65°C for 20min and the dephosphorylated product was used directly for ligation.

2.10.5 Agarose gel separation and DNA purification from gel slices

60 DNA fragments were mixed with DNA loading buffer and separated on 0.8%

agarose gels (for fragments larger than 500bps) or 1.5% agarose gel (for fragments

smaller than 500bps). DNA fragments were visualized by ethidium bromide-staining.

DNA in the gel slices were recovered using a gel purification kit (Qiagen) following the

manufacture’s instructions. The concentration of recovered DNA fragments was

estimated by running an aliquot on a gel in comparison with a 1kb DNA ladder (NEB).

2.10.6 Ligation

In ligation reactions, 200~500ng vector DNA was used and the molar ratio of

insert to vector was around 3:1. The ligation reaction was carried out with T4 DNA ligase

(Roche) as follows:

Insert and Vector 8µl

T4 DNA ligation buffer 1µl

T4 DNA ligase 1µl

The reaction was incubated overnight at 4°C.

2.10.7 Bacterial Transformation

40µl competent cells were transformed with 5µl of the ligation products by

incubation on ice for 30min, heat-shock at 42°C for 1min (XL1-blue supercompetent

cells) or 2min (home made DH5α competent cells) followed by incubation on ice for

another 2min. The transformed cells were then cultured in 1ml SOC medium for 60min at

37°C with shaking (180rpm). Afterwards, cells were pelleted with a brief spin in a

microfuge and resuspended in 100µl LB medium. 50µl of cells were plated on LB-Ampr plates and kept at 37°C for 16~18hr. Clones were picked and cultured in LB medium at

61 37°C. The following day, DNA was extracted using a DNA Miniprep kit (Eppendorf) and

analyzed by enzymatic diagnosis or further verified by DNA sequencing.

2.11 Construction of retroviral vectors LNIT-huSKI and TMP-tTA-mSki

The replication-defective retroviral vector LNITX was kindly provided by Dr. F.

Gage and was described earlier (158). A DNA fragment containing the entire coding

region of the human SKI cDNA was excised from the plasmid pSHHSKIN1D using

BamHI and ClaI, blunt-ended with T4 DNA polymerase, and inserted into the LINTX

vector at the PmeI site to produce LNIT-huSKI.

The replication-defective self-inactivating retroviral vector TMP-tTA was

modified from the TMP vector (kindly provided by Dr. Scott Lowe) (159-161) by

replacing its GFP gene with the tetracycline transactivator gene tTA from the LNITX

vector (158). The sequences of shRNAs targeting mouse Ski were chosen using RNAi

Codex (http://codex.cshl.edu/scripts/newmain.pl) and are designated by their positions in

the mouse Ski cDNA sequence (GenBank AF435852): mSki1145 (bases 1145-1163),

mSki1819 (bases 1819-1837) and mSki977 (bases 977-995), respectively. DNA forms of

these shRNA inserts were generated by PCR amplification of 97 base synthetic

oligonucleotides (Fisher Operon) using Pfu DNA polymerase (Invitrogen) and a common

set of primers (5'-cagaggctcgagaaggtatattgctgttgacagtgagcg-3' and 5'-

cgcggcgaattccgaggcagtaggca-3'). The PCR products were subsequently digested with

XhoI and EcoRI and inserted between these sites within TMP-tTA vector to generate

TMP-tTA-mSki1145, TMP-tTA-mSki1819 and TMP-tTA-mSki977. Clones containing these shRNA-encoding inserts were sequence-verified.

62 The PCR reaction was as follows:

Pfx PCR buffer 2×

dNTP 300µM

MgSO4 1mM

Primer 300nM each

PCR enhancer 0.5×

Pfx DNA polymerase 0.5 U

DNA template 60ng

Add sterile Milli-Q H2O to 50µl

The PCR program was:

Step 1 (initial melting) 94°C =5min

Step 2 (amplification) 94°C =45sec

68°C =1min15sec

Repeated for 25 cycles

Step 3 (final extension) 68°C =7min

2.12 Tissue culture and transfection

Proliferating mouse C2C12 myoblasts were maintained in growth medium (GM) consisting of Dulbecco's modified Eagle's medium (DMEM) (Gibco-BRL) supplemented with 20% fetal bovine serum (FBS, Atalantic Biosolution), 100μg/ml penicillin,

100units/ml streptomycin and 0.002% Fungizone (Gibco-BRL). To avoid spontaneous differentiation, cells were always kept in subconfluent (60-70%) conditions. Terminal differentiation was induced by switching subconfluent cell (80%) to differentiation

63 medium (DM) consisting of DMEM, 2% heat-inactivated horse serum plus antibiotics as in GM.

2.13 Retroviral infection

The retrovirus packaging cell line, PA317 (ATCC No. SD3443) was cultured in

DMEM containing 10% FBS, 100μg/ml penicillin and 100units/ml streptomycin. They

were transfected with retroviral constructs by using Fugene 6 (Roche) according to the

manufacturer's protocol. 24 hr after transfection, the medium was harvested and live cells

were removed by centrifugation at 1000×g for 10min. The retroviral-containing

supernatant was used to infect exponentially growing C2C12 cells at 25% confluence in

the presence of 10ng/ml polybrene (Sigma) to enhance infectivity. The infection was

repeated with freshly-harvested virus 12hr later. After an additional 12hr of culture,

C2C12 cells were switched to GM plus 2μg/ml doxycycline (Dox, an analog of

tetracycline) and antibiotics (G418 at 750μg/ml for LNITX-based constructs and

puromycin at 2ng/ml for TMP-tTA-based constructs). Two weeks later the resulting

individual colonies were isolated using cloning discs according to manufacturer’s protocol (PGC Scientifics, 62-6151-14) and expanded in GM with Dox and G418 or puromycin. After culture in GM minus Dox, clones were screened by western blot analysis for tet-regulated Ski overexpression or knockdown.

2.14 Western blotting

Whole-cell extracts were prepared from confluent C2C12 cells on 100 mm culture dishes (~3 × 106 cells) as follows: cells were washed twice with PBS, scrapped into

64 400µl of RIPA lysis buffer (supplemented with phosphatase and protease inhibitors

complete protease inhibitor cocktail (Roche)). After 10min incubation on ice, the

suspension was subjected to three freeze-thaw cycles and cell debris was removed by

centrifugation at 12,500 rpm for 15min at 4°C. Protein concentrations were determined

by Bradford assay (Bio-Rad) and equal amounts of proteins (20μg ~ 40μg) were boiled in protein loading buffer, separated by 6% SDS–PAGE and liquid transfered to Immobilon-

P membranes (0.45um, Millipore) at 100V for 3hr (small-size gel) or 4 hr (medium-size gel).

For most antibodies, the blots were washed for 10min in TBST, pre-blocked for

1hr at room temperature in blocking buffer and incubated with primary antibodies diluted in the blocking buffer overnight at 4°C. The membranes were then washed three times

with TBST, further incubated for 1hr at room temperature with secondary antibodies

conjugated to horseradish peroxidase (HRP) or alkaline phosphatase (ALP) followed by

thress washes with TBST. Signals were detected using either enhanced chemiluminescent

HRP substrate (Pierce) or CDP-Star® alkaline phosphatase chemiluminescent substrate

(1/50 dilution in assay buffer, Tropix) and exposure to HyBlot CL autoradiography film

(Denville Scientific).

When using anti-Ski G8 monoclonal antibody, the blots were pre-blocked

overnight at 4°C in blocking buffer and then incubated with primary antibody in the same

solution for 30min at room temperature. The membranes were washed twice with

blocking buffer and further incubated with ALP-conjugated secondary antibodies

(Sigma) for 30min at room temperature. Subsequently, the blots were washed with

65 blocking buffer then with assay buffer and the signal was detetected with CDP-Star® alkaline phosphatase chemiluminescent substrate as above.

The following primary antibodies were used for immunoblotting: Ski

(monoclonal antibody, 1/2000, G8, Learner Research Institute Hybridoma Core Facility or rabbit polyclonal antibody, 1/1000, H329, Santa Cruz); Myog (1/1000, F5D, Santa

Cruz), TFIIE-α (1/3000, C-17, Santa Cruz), MHC (1/1000, MF-20, Developmental

Studies Hybridoma Bank), MyoD (1/1000, 5.8A, Santa Cruz) and anti-Flag (1/5000, M2,

Sigma). The secondary antibodies used were: ALP-conjugated anti-mouse IgG (Fc

specific) and anti-rabbit IgG (whole molecule) (1/30000; Sigma); HRP-conjugated

Mouse IgG TrueBlotTM and Rabbit IgG TrueBlot™ (1/1000, eBioscience)

For reblotting the membrane with other antibodies, the membrane was stored in

plastic wrap at 4°C after immunodetection until stripping was performed. The membrane

was submerged in stripping buffer and incubated in a 70°C water bath for 30min with

occasional agitation. The membrane was then washed twice with TBST and Western blotting was performed as described above.

2.15 Immunofluorescence

LAB-TEK® eight-well chamber slides (Nunc) were coated with 1µg/cm2 laminin

(Invitrogen) at room temperature for at least 1hr. To assay the differentiation potential of cultures, C2C12 myoblasts (2×104) were seeded into each well in GM and switched to

DM at 80% confluence for 3 days. Cells were washed three times with PBS and fixed in

3.7% PFA/PBS for 30min at room temperature. After three washes with PBS, fixed cells were permeabilized with 10% goat serum/1% Triton-X 100/PBS for 10min at room

66 temperature, incubated with blocking buffer for 1hr at room temperature and then

incubated with primary antibodies diluted in blocking buffer overnight at 4°C.

Afterwards, cells were washed three times with PBST, incubated with secondary

antibodies for 1hr at room temperature, washed three with PBS again and mounted in

Vectashield aqueous mounting medium with DAPI (Vector Laboratories). Images were

obtained using Olympus BX50 upright fluorescence microscope equipped with a Polaroid

digital camera PDMC2, Polaroid PDMC2 software and fluorescent illumination. Images

were assembled into figures using Photoshop CS (Adobe).

The following primary antibodies were used for immunofluorescence: MyoD

(1/100; 5.8A, Santa Cruz); Myog (1/100; F5D, Santa Cruz), MHC (1/200; MF20;

Developmental Studies Hybridoma Bank) and p21 (1/100; SX118, BD Pharmingen).

Secondary antibodies were Alexa 488 or Alexa594-conjugated goat anti-mouse IgG

antibody (1/300, Molecular Probes). Control experiments performed with normal IgG as the primary antibody yielded no signal above the background.

Quantification was performed by counting at least 1000 DAPI-stained nuclei in

more than 10 random fields per culture plate. For MHC, the differentiation index =

Number of nuclei within MHC-stained multinucleate myotubes/total number of DAPI-

stained nuclei and the fusion index = the average number of nuclei per MHC-stained myotube. For nuclear proteins, the differentiation index=number of antibody-stained nuclei/total number of DAPI-stained nuclei. All experiments were performed in triplicate on three independent cultures and the standard deviation was calculated.

2.16 Quantitative realtime-PCR

67 RNA was isolated using a RNeasy kit (Qiagen) according to manufacture’s

instructions. Briefly, cells from a 100mm dish (~90% confluence) were washed with

DEPC-treated PBS and scrapped into 600µl RLT buffer. The lysate was homogenized by

vortexing, mixed with 1 volume 70% ethanol and RNA was collected by RNeasy spin column. After genomic DNA was removed by on-column DNase I digestion, RNA was eluted from the column in RNase-free water. cDNA was generated using reverse transcriptase SuperScript™ III (Invitrogen) with random hexamer primers according to the manufacturer’s instructions. RNA was incubated with hexamer and dNTPs as

follows:

Total RNA 5µg

Hexamer 200ng

dNTP 1mM

Incubate at 65 °C for 5min and placed on ice for another 2min

Afterwards, this mixture was incubate with

RT buffer 1×

MgCl2 2mM

DTT 10mM

RNase Inhibitor 40U

SuperScriptTM III RT 200U

The reverse transcription program was as follows:

25 °C =10min

50 °C =50min

85 °C =5min

68 Finally, the reaction was incubated with 1µl RNase H (2U/µl) at 37°C for 20min to degrade the RNA template.

Quantitative realtime-PCR (iCycler iQTM, Bio-Rad) was performed with 5-fold diluted RT reaction using SYBR® Green PCR Core Reagents (Applied Bioscience)

according to the manufacturers’ protocols.

The reaction was as follows:

SYBR Green Buffer 1×

MgCl2 3.5mM

dNTPs (with dUTP) 200µM each

Gold Taq DNA polymerase 0.625U

Primers 100nM each

Diluted cDNA 1µl

Add sterile Milli-Q H2O to 25µl

The PCR program was as follows:

Step 1 (initial melting) 94°C =10min

Step 2 (amplification) 94°C =15sec

60°C =20sec

72 °C=20sec

Repeated for 40 cycles

Step 3 (final extension) 72°C =5min

Step 4 (melting curve) starting from 72°C, the temperature was increased

by 0.5°C each cycle until 94°C (totally 44 cycles)

69 Analyses were performed in triplicate on RNA samples from three independent

experiments. Threshold cycles (Ct) of target genes were normalized against the

housekeeping gene Gapdh, and relative transcript levels were calculated from the Ct

values as Y=2-ΔCt, where Y is fold difference in amount of target gene versus Gapdh and

ΔCt=Ctx-CtGapdh.

2.17 Chromatin immunoprecipitation (ChIP)

C2C12 cells cultured in GM or DM for 36hr (~2×107 cells, 150mm plate) were

washed with cold PBS and cross-linked with 1% formaldehyde/PBS for 10min at room

temperature. Fixation was stopped by with the addition of 2.5ml 1.25M Glycine for 5min

at room temperature. Fixed cells were further washed with PBS, scrapped from the plates

and resuspended at 2×107 cells/ml in lysis buffer. Suspensions was sonicated for 8 cycles

of 0.5sec on/0.5sec off for a total of 1min to yield chromatin with an average DNA length of 200~500bp. The size of sonicated chromatin was analyzed as following: 20μl sonicated cells were mixed with 30μl H2O and 5μl 5M NaCl and kept in a thermal cycler at 65°C overnight to reverse the cross-link. Afterwards, samples were treated with 10µg

RNase A at 37°C for 30min and 50µg proteinase K at 45°C for 1.5 hr. After being mixed with 50μl Phenol/CHCl3/isoamyl alcohol and centrifuge at 12,500rpm for 5min, the

aqueous phase was resolved in a 1.5% agarose gel (30V for 45min to diffuse out the salt followed by 100V for 60min) and visualized by ethidium bromide staining.

Equal amounts of sheared chromatin from each sample (2×106 cells per assay)

were preabsorbed at 4°C for 1hr with 40μl 33% slurry of pre-blocked protein A beads

(Repligen). After pelleting the beads by centrifugation, the supernatant was incubated

70 overnight at 4°C with either 2µg rabbit polyclonal Ski antibody (H329, Santa Cruz) or

normal rabbit IgG (Santa Cruz). Antibody-chromatin complexes were captured by

incubation with 40µl 33% slurry pre-blocked Protein A beads for 1hr at 4°C. The beads

were washed sequentially in lysis buffer, high salt buffer, lithium salt buffer and TE

buffer. Complexes were eluted from the beads in 300μl of elution buffer.

The immunoprecipitated DNA and input DNA were mixed with 5M NaCl to final concentration of 0.2M and formaldehyde cross-linking was reversed by overnight incubation at 65°C. Afterwards, a 40µl sample was treated with 10µg RNase A and 50µg proteinase K and DNA was isolated by Phenol/CHCl3/isoamyl alcohol and ethanol

precipitation as described above. The aqueous phase was then incubated with 10µg

glycogen and 2.5 volume of cold 100% ethanol for 1hr at -80°C and DNA was pelleted

by centrifugation for 30min at 12,500rpm. DNA pellets were washed with room

temperature 70% ethanol; air dried and resuspended in 40µl of 10mM Tris-HCl pH8.5.

The optimal PCR cycle numbers were determined by realtime-PCR and 5% of purified

DNA was analyzed by regular PCR using HotStart-IT Taq Master mix (USB). For input

control, 10% of cross-linked chromatin was precleaned and purified as described above

and then assessed for PCR by using the same sets of primers.

The regular PCR reaction was set up as follows:

DNA from ChIP assay 2µl

Primers 100nM each

HotStart-ITTM Taq Master mix 1×

H2O to 25µl

The real-time PCR reaction was set up as follows:

71 DNA from ChIP assay 2µl

Primers 100nM each

HotStart-ITTM Taq Master mix 1×

Fluoresence Dye (Bio-rad) 10nM

SYBR Green (Bio-rad) 0.2×

H2O to 25µl

The PCR program was:

Step 1 (initial melting) 95°C =2min

Step 2 (amplification) 95°C =30sec

60°C =30sec

72 °C=60sec

Repeated for 35 cycles

Step 3 (final extension) 72°C =5min

25% of each reaction mixture was resolved on a 1.5% agarose gel (30V for 45min to diffuse out the salt followed by 100V for 60min) and visualized by ethidium bromide staining.

2.18 Genomic DNA preparation

Cells on 100mm plates were washed twice with PBS, scrapped into three

Eppendorf tubes and pelleted by centrifuge at 5,000rpm for 4min. Each tube of cells was incubated with 1ml of lysis buffer at 55°C with occasional agitation for 3hr and then kept at 55°C without agitation overnight. After being mixed with an equal volume of

Phenol/CHCl3/isoamyl alcohol and centrifuging at 12,500rpm for 10min, the aqueous

72 phase was mixed with 2.5 volume of cold 100% ethanol and DNA was pelleted at

12500rpm for 10min. After washing with 70% ethanol, the DNA pellet was air dry and

dissolved in 250µl TE at 37°C overnight.

2.19 Reporter assays

2.19.1 Construction of luciferase reporters

DNA fragments containing the full-length (-1102~+18, Genbank M95800) or

proximal (-184~+18, Genbank M95800) Myog regulatory region were amplified by PCR

using C57BL/6 genomic DNA as template.

The real-time PCR reaction was set up as follows:

Pfx Buffer 1×

dNTP mix 300µM

MgSO4 1mM

Primers 300nM each

Genomic DNA 360ng

Pfx DNA polymerase 1U

Enhancer 0.5×

Add H2O to 50µl

The PCR program was:

Step 1 (initial melting) 94°C =2min

Step 2 (amplification) 94°C =15sec

55°C =30sec

68 °C=2min

73 Repeated for 30 cycles

Step 3 (final extension) 68°C =7min

Myog184 luciferase reporter carrying the luciferase gene downstream of this

Myog regulatory region was generated by inserting HindIII-StuI digested PCR product

into the HindIII-SmaI site of the pGL3-Basic vector (Promega). Myog-luciferase

constructs with E box, MEF2 and MEF3 (Myog184-E1E2m, Myog184-

MEF2m and Myog184-MEF3m) were generated from Myog184 luciferase reporter using

the QuikChange® Site-Directed Mutagenesis Kit (Stratagene). Mutagenic primers were

designed using Stratagene’s QuikChange® Primer Design Program at

http://www.stratagene.com/qcprimerdesign. Ski expression vector pCDNA-huSki was

previously described (24).

The Site-Directed Mutagenesis PCR reaction was set up as follows:

DNA template (Myog184) 100ng

SDM PCR buffer 1×

Primer 125ng each

dNTP mix 125µM

Pfu DNA polymerase 2.5U

Add H2O to 50µl

The PCR program was:

Step 1 (initial melting) 95°C =30sec

Step 2 (amplification) 95°C =30sec

55°C =1min

68 °C=17min

74 Repeated for 18 cycles

After amplification, 1µl DpnI (10U/µl) was added to the PCR reaction and

incubated for 6hr at 37°C to digest the parental supercoiled dsDNA. 1µl of the DpnI-

treated DNA was transformed into XL1-Blue supercompetent cells and individual clones

were picked and expanded. DNA was extracted from these cells using a Miniprep Kit

(Eppendorf) and sequence-verified.

2.19.2 Reporter assays

Cells (1×105 per well) were seeded in 12-well plates and transfected at 80% confluence using Lipofectamine 2000 (Invitrogen) with a combined total of 4µg expression vector and reporter plasmids DNAs. 18 h after transfection, cells were switched to DM and incubated for another 48hr prior to cell harvest. Transfected cells were washed twice with PBS and then lysis in 70μl of 1× Reporter Lysis Buffer

(Promega) with one round of freeze-thaw followed by incubation at room temperature for

20min with shaking. Cell debris was removed by centrifugation and the luciferase activity in the supernatant was determined by the dual luciferase reporter assay system

(Promega), according to the manufacturer’s instructions. A Renillar luciferase vector pGL3-TK-Renillar was used as an internal control. Briefly, in a 96-well white plate, 70μl cell lysate was incubated with equal volume of Dual-GloTM Luciferase Reagent for 10min

and the Firefly luciferase activity was measured. Afterwards, this mixture was incubated

with 1 volume of diluted Stop&Glo substrate (1/100 in Dual-GloTM Stop&Glo Reagent)

for another 10min and the Renilla luciferase activity was measured. Luciferase units were

determined using a MAXline microplate luminometer (Molecular Devices). The relative

light units (RLUs) were generated by normalizing firefly luciferase units to Renilla

75 luciferase units of the co-transfected pTK-Renilla-luc vector. Relative Response Ratio

was calculated by comparing RLUs of different samples. The experiments were done in

duplicate and the reported results represent at least three independent experiments.

2.20 Immunoprecipitation

C2C12 cells (~ 50% confluent, 100mm dishes) were transfected with expression

plasmids for pCMX-flag-Six1, pCMX-flag-Eya3 (No 3’ UTR), pCMX-flag-Mef2c,

pCSA-MyoD and full-length Ski or its mutants with Lipofectamine 2000. 18 h after

transfection, cells were switched into DM and cultured for 2 days. Cells were scrapped

from the plates, resuspended in 1ml of NETN buffer and sonicated 0.5sec on/0.5 sec off

for a total of 1 min. Cell debris were removed by centrifugation for 15 min at 10, 000rpm

at 4°C and lysate was preabsorbed with 30μl 50% slurry protein A beads for 1hr at 4 °C

followed by centrifugation. After the protein concentration was determined with a

Bradford protein assay (Bio-Rad), a fraction of the cell lysates were subjected to immunoblotting to detect the expression of the proteins of interest and, if needed, the amounts of lysates used for immunoprecipitation were adjusted accordingly.

Immunoprecipitation was carried out as following: ~200µg precleaned lysate was

incubated in parallel with either 4µg rabbit polyclonal anti-Ski (H329, Santa Cruz) or

normal rabbit IgG for 3 h at 4°C. The immunocomplexes were then incubated with 40μl

33% slurry protein A beads (Repligen) overnight at 4°C. Precipitates were washed 5

times in ice-cold NETN buffer, resuspended and released from the beads by boiling in

protein loading buffer for 10 min. The precipitated proteins were separated by 6% SDS–

PAGE and analyzed by Western blotting. If precipitating and primary Western blotting

76 antibodies were from the same species, either HRP-conjugated Mouse IgG TrueBlotTM or

Rabbit IgG TrueBlot™ was used as the secondary antibody accordingly.

77 CHAPTER I: Ski Is Required For Normal Development of Embryonic Progenitors

of Mouse Hypaxial Muscle but Not For Fetal Progenitors or Satellite Cells

This chapter has been submitted to Molecular and Cellular Biology

Abstract

The Ski proto-oncogene encodes a transcriptional co-regulator that stimulates

muscle differentiation in vitro and promotes muscle hypertrophy in vivo. Ski-/- fetal mice

exhibit dramatically reduced skeletal muscle mass. Here we show that the defect is most severe in hypaxial muscles which are derived from progenitor cells that originate in the somite-derived dermomyotome and migrate to distal sites of muscle development. In Ski-

/- embryos the dermomyotome develops normally as does the myotomal epaxial muscle that develops from non-migratory progenitors. Using the myogenic regulatory factors

(MRF) as markers, we found that myogenic determination (Myf5 and MyoD) and

commitment to differentiation (Myog) were initiated appropriately in the absence of Ski.

However, cells expressing MRF genes were reduced in number and restricted to more proximal locations in the limbs. This defect apparently resulted from impaired distal migration of embryonic myogenic progenitors (Pax3 positive) that accumulated at the body/limb bud junction and dorsoproximal region of the limb; not from a failure of myogenic cells to proliferate and differentiate or to excessive apoptosis. Defective migration was not due to reduced BMP signaling or altered expression of HGF/SF but was likely caused by reduced expression of Met, a downstream target of Pax3 and known

effecter of progenitor migration. Although they derive from the same pool of migratory

progenitors as their embryonic counterparts, we found that the number and distribution of

78 fetal myogenic progenitors and satellite cells were not affected by the lack of Ski. When cultured in vitro these Pax7-positive satellite cells gave rise to myoblasts that were fully capable of terminal muscle differentiation.

79 Introduction

The Ski proto-oncogene encodes a bifunctional transcriptional co-regulator that is ubiquitously expressed and participates in several key regulatory pathways. Ski does not bind DNA but associates with DNA binding transcription factors and mediates either transcriptional repression or activation. The best characterized role of Ski is as a co- repressor of the Smad proteins activated by transforming growth factor beta (TGFβ) or bone morphogenetic proteins (BMPs) (21-24). Ski is also a co-repressor of hormone receptors (16, 18, 37), Gli3 (38), MeCP2 (29), and GATA-1 (162). On the other hand, Ski acts as a transcriptional co-activator with NF-I (163), β-catenin (8) and the myogenic regulatory factor (MRF), MyoD (114).

Ski was shown to activate the muscle-specific enhancers of myosin light polypeptide 1 (Myl1), muscle creatine kinase (Ckm) and myogenin (Myog) in cooperation with MyoD (114-116). Co-activation of muscle-specific gene transcription by Ski has been shown to be related to its normal biological function by analyses of transgenic and knock-out mice (45, 47). The ability of Ski to activate endogenous Myog expression

(149) almost certainly underlies its ability to induce muscle differentiation in vitro (42) and muscle hypertrophy in transgenic mice, in vivo (45). These findings suggest that Ski plays a role in terminal muscle differentiation; consistent with the observation that during mouse development Ski expression in skeletal muscle is upregulated after the commitment to myogenic differentiation (39).

Ski has also been implicated in myogenic lineage determination by virtue of its ability to induce the transdifferentiation of fibroblasts into fully functional myoblasts

(42). This role for Ski was also suggested by its expression within the lateral regions of

80 the somites (dermomyotome) that contain the progenitors of skeletal muscle (39). These progenitors do not express muscle-specific genes, including the MRFs, but they are distinguishable by the expression of the paired box genes, Pax3 and Pax7 (70). Some of

the progenitors in the dermomyotome delaminate to form the myotome, which gives rise

to locally differentiating epaxial muscles of the trunk and back (65). Their determination

to the muscle lineage is marked by turning off Pax3/7 and expressing the MRF gene,

Myf5 (92). Hypaxial muscle on the other hand, is derived from progenitors within the

dermomyotome that delaminate and migrate to distal regions, such as the limb buds and

ventral trunk (65). They also commit to the myogenic program by turning off Pax3/7 and

turning on Myf5, MyoD and eventually, Myog expression (92). In Pax3 mutant mice,

delamination and migration of these progenitors are defective, leading to a loss of hypaxial muscle (72-77). Defects in hypaxial muscle formation arising from abnormal migration are also observed in mice lacking either receptor tyrosine kinase Met (78, 79), which acts downstream of Pax3 (73, 76, 86-88) or its ligand hepatocyte growth factor/scatter factor (HGF/SF) (78, 79). Activation of Met by HGF/SF mediates epithelial-mesenchymal transition of dermomyotome cells to generate migratory myogenic progenitors and guides their subsequent migration into limb buds. In distal regions, such as limb buds, BMPs function to maintain the viability and proliferative potential of Pax3-positive migratory progenitors (89).

A fraction of the progenitors that migrate into muscle-forming regions do not commit to terminal differentiation. In late development these cells are distinguished by expression of Pax7 but not Pax3. Some of them serve as a reservoir of fetal muscle progenitors while others become satellite cells, which are located under the basal lamina

81 of muscle fibers and comprise the lineage-restricted progenitor cells of post-natal skeletal

muscle (70). Studies of knockout mice indicate that the specification and function of these cells are defective in the absence of Pax3 or Pax7 and that their myogenic

determination and/or terminal differentiation are compromised by the absence of the

MRF genes, Myf5, MyoD or Myog (70).

Defective skeletal muscle development is also a feature of Ski-null mice (47).

These mice have several additional defects, some of which are highly dependent upon the

genetic background; exencephaly is a prominent feature in 129/S6 Ski-/- mice whereas

midline facial clefting occurs in a C57BL/6J background (48). However, both Ski-/-

congenic strains show hypoplasia and disorganization of several types of skeletal muscle including limb, diaphragm and tongue muscles (47, 48). The muscle phenotype of Ski-/-

mice is reminiscent of that in mice lacking Myog in that most muscles are present but

smaller than normal and with somewhat disorganized myofibers (99, 100). This

similarity, and the fact that Ski was shown to activate Myog transcription, led us to re-

assess Ski expression during early mouse development, to examine MRF gene expression

in Ski knockout mice, and subsequently to investigate BMP signaling and the expression

of Pax3, Pax7, Met, HGF/SF. Our findings reveal a role for Ski in the regulation of

embryonic but not fetal myogenic progenitor migration.

Experimental Procedures

Mouse strains and embryos

Generation of C57BL/6 mice with a null mutation of Ski has been described (47,

48). Embryos obtained by heterozygous mating were staged by counting the appearance

82 of the vaginal plug as embryonic day 0.5 (E0.5). Greater than 90% of C57BL/6 Ski-/-

embryos exhibit midline facial clefting by E12.5. In all comparative studies, sibling

embryos of similar somite number and/or crown-rump size were chosen and genotyped

by PCR using yolk sac or tail DNA.

Histological analysis

Dissected embryos were fixed overnight at 4°C with 3.7% formaldehyde in

phosphate buffered saline (PBS), embedded in paraffin and sectioned at 12 µm. Sections

were stained with hematoxylin/eosin and photographed under bright-field illumination

with an Olympus BX50 microscope equipped with a Polaroid PDMC2 digital camera and

software. Figures were assembled using Photoshop CS (Adobe).

Immunohistochemistry

Embryos were fixed for 1-4 hours at 4°C with 4% paraformaldehyde (PFA) in

PBS, equilibrated in sucrose/PBS at 4°C and embedded in OCT (Tissue-Tek, Miles Inc.,

Elkhart, Illinois, USA). 10μm cryostat sections were mounted on glass slides and immunofluorescence on equivalent sections from sibling Ski-/- and control embryos was performed as described (153). Stained sections were mounted in aqueous Vectashield

(with DAPI) and fluorescent images were obtained with an Olympus BX50 microscope equipped with a PDMC2 Polaroid digital camera. Figures were assembled using

Photoshop CS. Experiments involving cell counts were performed on 4 or more mutant and wild-type pairs with each antibody. Statistical significance (set at P<0.05) of differences between mutants and controls was assessed using Student’s t-test.

The following antibodies at the indicated dilutions were used: Pax3 monoclonal,

1/200, Myosin heavy chain (MHC) monoclonal (MF20), 1/400, and Pax7 monoclonal,

83 1/200 (Developmental Studies Hybridoma Bank); Pax3 rabbit polyclonal, 1/200, kindly

provided by M. Goulding of The Salk Institute (80); Myogenin monoclonal (F5D), 1/200

(Santa Cruz), Desmin monoclonal (D33), 1/400 (Dako); phospho-Histone H3 (Ser 10) rabbit polyclonal, 1/200, and activated Caspase-3 (Asp 175) rabbit polyclonal, 1/200

(Cell Signaling); AlexaFluor 488 and AlexaFluor 594 secondary antibodies, 1/300

(Molecular Probes).

Whole mount in situ hybridization

Embryos were dissected in diethyl pyrocarbonate treated PBS, fixed in 4% PFA at

4°C for 1- 4 hours, washed in PBS, dehydrated stepwise in 25%, 50%, and 75%

methanol in PBT (PBS plus 0.1% Tween 20) and stored at –20°C in 100% methanol.

Fixed embryos were rehydrated through the reversed methanol series back to PBT and

whole mount in situ hybridization was performed as described by Matise and Joyner

(157).

Stained embryos were refixed in 4% PFA/PBT, dehydrated through 25%, 50%,

75%, and 100% ethanol series in PBT and photographed under Zeiss KL1500 LCD

illumination, using a Zeiss stemi SV11 stereomicroscope equipped with a Zeiss AxioCam

color digital camera and Zeiss Axiovision software version 2.0. Figures were assembled

using Photoshop CS. At least three mutant and wild-type pairs were analyzed with each

probe. Control sense strand probes showed no significant signals.

Digoxigenin (DIG)-labeled RNA probes were transcribed from linearized

plasmids with DIG RNA Labeling Mixture (Roche) and T3 (Ambion) or T7 (Roche)

RNA polymerase. Probes for Pax3 (75); Msx1 (154); Myog, MyoD (155) and Myf5 (156)

have been described. Met probes were subcloned fragments of bases 167-757 and 1837-

84 2347 of murine Met cDNA (88). Mouse Ski probes were four subclones of bases 263-969,

969-1320, 1320-1773 and 1773-2123 of Ski cDNA (GenBank accession no.

NM_011385). HGF/SF (GenBank accession no.NM_010427) probes were four subclones of bases 322-578, 871-1068 and 1837-2347.

Primary culture of satellite cell-derived myoblasts and immunodetection of MHC

Satellite cell-derived myoblasts were isolated from pooled limb muscles of E15.5 and E18.5 embryos by enzymatic dissociation (152) with minor modifications. Limb muscles were digested in 0.0375% trypsin-EDTA (GIBCO-BRL), 33 unit/ml collagenase

III (GIBCO-BRL) in Hank’s BSS at 4°C overnight. After preplating the cell suspension on standard culture dishes, unattached myoblasts were counted and plated on culture dishes coated with laminin (1 µg/cm2) (GIBCO-BRL).

Growth medium (GM) was Dulbecco's modified Eagle medium (DMEM)

(GIBCO BRL) plus 20% fetal bovine serum (GIBCO BRL), 0.1 mM BME, 10 ng/ml human recombinant basic fibroblast growth factor (GIBCO BRL), and 2 µg/ml gentamycin (GIBCO BRL). Differentiation medium (DM) was DMEM plus 2% horse serum (GIBCO BRL) and 2 µg/ml gentamycin. GM was changed every 12 hour and DM every second day.

Cells were cultured to 60% confluence in GM then cultured in DM for the indicated time periods. Cells were fixed in 4% PFA and differentiated myoblasts were detected with anti-MHC (MF20, 1: 100) and FITC-conjugated secondary antibodies.

Nuclei were stained with DAPI. Images were taken from representative areas and about

200 total cells/area were counted.

Quantitative realtime-PCR

85 Isolation of RNA using RNeasy (Qiagen), reverse transcription using

SuperScript™ III (Invitrogen) and quantitative realtime-PCR (iCycler iQTM, Bio-Rad)

using SYBR® Green PCR Core Reagents (Applied Bioscience) were performed

according to the manufacturers’ protocols. The PCR primers were: Myf5 (GenBank accession no. NM_008656) forward: 5'-TATTACAGCCTGCCGGGACA-3'; reverse: 5'-

CTGCTGTTCTTTCGGGACCA-3' and Pax7 (GenBank accession no. AF254422)

forward: 5'- CCACCCACCTACAGCACCAC-3'; reverse: 5'-

GCTGTGTGGACAGGCTCACG-3'. Triplicate analyses were performed on samples from three independent cultures. Relative transcript levels were calculated from the threshold cycle numbers (Ct) as Y=2-ΔCt, where Y is fold difference in amount of Myf5 or

Pax7 versus Gapdh and ΔCt=CtMyf5-CtGapdh or CtPax7-CtGapdh.

Results

Selective muscular hypoplasia of Ski-/- embryos

Previous studies (47) revealed a dramatic reduction in skeletal muscle mass in Ski-

deficient mice but the underlying mechanism was not determined. As a first step in

analyzing the defect, we performed an extensive histological examination of the muscle

of perinatal mice at E18.5. Examples of this analysis in Fig. 3 show that the loss of Ski

affects development of hypaxial muscle to a much greater extent than epaxial muscle.

Distal hypaxial muscles of the limb, such as flexors and extensors along the ulna and

radius, show significant reductions in size and some muscles are completely absent (Figs.

3A-B2). Other limb muscles such as the triceps brachii are markedly reduced in size and

disorganized (Figs. 3C-D2). However, shoulder muscles (around the scapula) are present

86

87 Fig. 3. Selective muscular hypoplasia in prenatal Ski-/- mice. Hematoxylin/Eosin-staining

of transverse sections of E18.5 Ski+/- (A, C, E, G) and Ski-/- (B, D, F H) embryos. (A, B)

Distal forelimb muscles are absent (arrowheads) or severely reduced (arrows) in the Ski-/-

embryo. (A1, B1, A2, B2) Higher magnification of A and B showing areas with muscle

defect; r, radius; u, ulna. (C, D) Proximal forelimb showing reduction of triceps brachii

(arrows and arrowheads) in Ski-/- embryo compared to normal littermates. (C1, D1, C2,

D2) Higher magnification of areas indicated by arrows and arrowheads in C and D; h,

humerus. (E-H) Shoulder (arrowheads) and back (arrows) muscles are normal in size

although distorted in shape and position due to severely reduced brown fat formation in

the absence of Ski; s, scapula; bf, brown fat. Scale bar (A-D, E-H) 125μm, (A1-D2)

50 μm.

88 and only slightly affected in Ski-/- embryos (arrowheads, Figs. 3E and F). Back muscles

are even less affected and some are comparable to the control (arrows, Figs. 3E-H).

Although Ski-/- mutants are leaner than controls, bones of both the shoulder (scapula) and the limb (humerus, ulna and radius) appear normal in diameter and length. This

observation and the relatively normal epaxial muscle suggest that the Ski null mutation

results in a specific reduction of hypaxial myogenesis rather than a general growth defect

or developmental delay.

To determine whether the defect in hypaxial muscle development was apparent at

earlier stages, we assessed the presence and size of muscle groups at E12.5 and E15.5.

Myosin Heavy Chain (MHC) staining revealed development of appropriate muscle

groups in the trunk and proximal limbs of Ski-/- embryos, but the sizes of these muscles

appeared markedly reduced compared to the wild types (Fig. 4). This phenotype was

more severe in the distal forelimbs where some muscle groups were completely missing

(arrowheads, Figs. 4E and F). Results obtained by staining adjacent sections with anti-

desmin antibody were virtually identical to the observations of MHC staining (data not

shown). Thus the defect in hypaxial muscle development was already apparent at the

onset of secondary myogenesis at E12.5.

Dynamic expression of Ski in the developing neural tube, somites and limb buds of mouse

embryos

The hypaxial muscle defects observed in the Ski-null mice could be due either to

early effects on migration and/or commitment of progenitors, or to later effects on

terminal differentiation. To determine whether the pattern of Ski expression during

development can distinguish among these possibilities, Ski mRNA expression in mouse

89

90 Fig. 4. Hypoplasia of limb muscle in Ski-/- embryos at mid and late gestation.

Immunofluorescence with anti-MHC on transverse thoracic sections of E12.5 and E15.5 wild-type (A, C, E) and Ski-/- embryos (B, D, F). Forelimb muscles were either severely reduced in size (compare the boxed areas in A, B magnified in A1, B1 and panels C-F, arrows) or virtually absent (E, F, arrowheads) in the Ski-/- mutant. Scale bar (A, B, C-F)

125μm, (A1, B1) 50 μm.

91 embryos from E9.5 to E12.5 was examined by whole mount in situ hybridization. At E9.5

Ski mRNA expression was highest in the neural tube but was also detected laterally in the somites (Figs. 5A and A1). At E10.5 (Fig. 5B), Ski expression in the spinal cord had decreased rostrally and exhibited a gradual increase toward the tail region where it was also observed in the most recently formed somites and in presomitic mesoderm. Ski was also highly expressed throughout the apical ectodermal ridge (AER) of the limb buds and in the underlying limb mesenchyme in a distal to proximal decreasing gradient. At E12.5,

Ski expression in the distal limbs was highest in the interdigital mesenchyme (Fig. 5C).

The results revealed that Ski expression during early mouse embryogenesis is highly dynamic in the neural tube, somites and limb buds. Combined with previously published data which also revealed stage-dependent expression of Ski in body wall muscles (39, 40), these results are consistent with a role of Ski in hypaxial muscle development, but they do not discriminate between an early role in determination or migration and a later role in differentiation.

Determination and differentiation of myogenic progenitors in Ski-/- embryos

To investigate the origins of the selective muscle defects in Ski knockouts, the initiation and progression of epaxial and hypaxial myogenesis were examined at earlier stages. Although these two types of muscles arise from distinct developmental pathways, the MRF genes are required for the determination and differentiation of both. Epaxial muscle originating in the myotome is the first skeletal muscle to be generated (65). The emergence of determined myogenic cells in the myotome is marked by activation of Myf5

(156, 164), which is followed by Myog expression as these cells commit to terminal differentiation (155). To establish whether loss of Ski affects the determination or

92

93 Fig. 5. Dynamic expression pattern of Ski during mouse development. Whole mount in situ hybridization with Ski probes was performed on E9.5 (22 somites) (A, A1), E10.5

(30 somites) (B) and E12.5 (48 somites) (C) embryos. Expression in the neural tube

(white arrow), somites (arrowhead) and limb buds (black arrows) is indicated. nt, neural tube; so, somite; fl, forelimb; hl, hindlimb. Scale bar (A) 100μm, (A1) 50μm, (B)

125 μm, (C) 250 μm.

94 differentiation of myogenic progenitors, the distribution of Myf5-positive (Myf5+) or

Myog+ cells was examined in E9.5 embryos, while undetermined progenitors in the

dermomyotome were identified by Pax3 expression (Figs. 6A and B). In the Ski-/-

embryos, the distribution of Myf5+ and Myog+ cells in the myotome was very similar to that of the wild type embryos. This observation suggests that development of epaxial muscle in the myotome is Ski-independent, in agreement with the finding that at later stages these muscles are relatively normal in the Ski-/- embryos. Furthermore, the

absence of Myf5+ and Myog+ cells in the Pax3+ dermomyotome of mutant embryos

indicates that hypaxial myogenic progenitors do not prematurely commit to

differentiation in the absence of Ski.

At E10.0, hypaxial myogenic progenitors that have migrated to their distal locations also initiate myogenesis by activating expression of MRFs (92). To investigate whether the myogenic program is activated normally in the absence of Ski, we examined the expression of Myf5, MyoD and Myog by whole mount in situ hybridization. In wild type embryos at E10.0, Myf5 transcripts were present in the myotome of all somites as well as in the distal locations populated by migratory myogenic progenitors (Figs. 6C and

E). In contrast to the normal expression of Myf5 in the myotome its expression in the forelimb buds of Ski-/- embryos was impaired compared to that of controls (Figs. 6D and

F). At E10.5 MyoD expression reflected this pattern with normal levels in the myotome but greatly reduced levels in the limb buds (Figs. 6G-J). By this stage of development, expression of Myog initiates terminal muscle differentiation in the myotome and in the limb buds (155). Consistent with the observed patterns of Myf5 and MyoD expression, expression of Myog in Ski-/- embryos was similar to that of control embryos in the

95

96 Fig. 6. Specification and activation of epaxial and hypaxial myogenesis. (A, B) Thoracic transverse sections of wild-type and Ski-/- E9.5 embryos were examined by double immunofluorescence for expression of Pax3 and Myf5 (A) or Pax3 and Myog (B). (C-N)

Whole mount in situ hybridization for Myf5 (C-F) of E10 (30 somites) embryos or for either MyoD (G-J) or Myog (K-N) of E10.5 (34, 38 somites) embryos. Note normal expression of MRF genes in the myotome (arrowheads) but decreased extent and distribution of expression in the limb buds (arrows) of Ski-/- embryos. Scale bar (C, D, K-

N) 150μm, (E, F, I, J) 100 μm, (G, H) 200 μm.

97 myotome but not in the limb buds (Figs. 6K and L). The dense groups of Myog+ cells that

extended ventrally and distally in the forelimb buds of wild-type embryos (Fig. 6K) were

dramatically depleted proximally and almost undetectable distally in Ski-/- embryos (Fig.

6L). A similar deficiency was observed in the hind limbs (Figs. 6M and N).

The analysis of MRF gene expression indicated that, in the absence of Ski,

differentiation of myotomal epaxial muscle was normal, but the determination and

commitment of hypaxial myogenic precursors, although initiated appropriately, was

defective in extent and pattern. To determine whether this problem persisted during the

transition from primary to secondary myogenesis, we analyzed hypaxial muscle formation

at successive stages, using desmin as a marker for all myoblasts and Myog to identify

myoblasts committed to terminal differentiation. At E11.5, the first appearance of hypaxial musculature was observed at the body-limb junction and in the proximal forelimb (Fig. 7). Although masses of desmin+ myoblasts had formed at the body-limb

junction of both Ski-/- and wild type embryos (red arrows, Figs. 7A and B), the spread of

these myoblasts into the forelimbs was markedly compromised in Ski-/- embryos, with

few desmin+ cells detected proximally and none detected distally (white arrows and

arrowheads, Figs. 7A and B). Myog+ cells were found in the proximal limb (arrow, Fig.

7C) and extending in two streams ventrally and dorsally toward the distal end of the limb

in wild-type embryos (arrowheads, Fig. 7C). By contrast, Myog expression in Ski-/-

embryos was markedly reduced in both intensity and extension into the distal and ventral

forelimb (arrow and arrowheads, Fig. 7D), confirming the observation of desmin staining.

In summary, these results indicate that in Ski-/- embryos determination and commitment

98

99 Fig. 7. Normal commitment to terminal muscle differentiation in the forelimbs of Ski-/-

embryos. Transverse sections of Ski+/+ and Ski-/- embryos were examined by

immunofluorescence with antibodies to desmin (A, B) or Myog (C, D); nuclei were

stained with DAPI. (A, B) Note the accumulation of desmin+ cells at the limb/body

junction (red arrows) and the very low numbers of these cells in the limb (white arrows and arrowheads) of the E11.5 Ski-/- embryo. (C, D) The number of Myog+ cells was

reduced in Ski-/- embryos, particularly in the distal and ventral regions of the limb

(arrows and arrowheads). (E-H) At both E12.5 and E15.5, reduced numbers of Myog+ cells in the Ski-/- embryos (F, H) compared to the Ski+/+ littermates (E, G). (I) the percentage of Myog+ nuclei (values above bars) was not significantly reduced in Ski-/-

embryos (P>0.3). Error bars = standard deviation, n= number of samples. Scale bar (A-D)

75μm, (E-H) 50 μm.

100 to differentiation are initiated correctly in both epaxial and hypaxial muscle, but the latter

exhibit defects in both the distribution and the abundance of committed myogenic cells.

Analysis of Myog expression at later stages revealed that the reduction of

committed myoblasts persisted in the Ski-/- embryos. At E12.5 and E15.5, the total

number of Myog+ cells in individual muscles of the mutants was consistently reduced compared to their counterparts in the wild-type (Figs. 7E-H). This reduction could have

resulted either from a decrease in the number of myogenic progenitors or from a block in

their commitment to differentiation. These alternatives were distinguished by comparing

the percentage of Myog+ cells in corresponding muscles of wild type and Ski-/- embryos.

In the case of reduced numbers of myogenic progenitors but normal commitment, this

percentage should be similar in both genotypes. If, on the other hand, there were normal

numbers of progenitors in Ski-/- embryos but impaired differentiation, the percentage of

Myog+ cells would be reduced. We found that the percentage of Myog+ cells was not

significantly different between Ski-/- and Ski+/+ littermates (Fig. 7I). Thus, the reduction in the total numbers of Myog+ cells and the subsequent large reduction in the size of

hypaxial muscle in the mutant (Figs. 3 and 4) did not appear to be due to a block in

commitment to terminal differentiation.

Taken together, our data indicate that the reduced size of hypaxial muscle groups

in the Ski-/- embryos appears to result from neither premature nor defective differentiation, but from an insufficient pool of myogenic progenitors that give rise to the committed myogenic cells. This deficiency could be the consequence of reduced proliferation, abnormal apoptosis or defective migration. We examined the numbers of mitotic and apoptotic myogenic cells in E10.5, E12.5 and E15.5 embryos by

101 immunofluorescence with antibodies for phosphorylated histone H3 and activated caspase-3, respectively. No significant differences in the levels of proliferation and apoptosis in either the Pax3+ progenitors or the desmin+ muscle cells were observed in mutant compared with wild-type embryos (Fig. 8), ruling out these processes in the defective accumulation of distal myogenic cells.

Altered migration of myogenic progenitors in Ski-/- embryos

In light of the above results we hypothesized that defective migration of progenitors was responsible for the selective defect in the hypaxial muscle of Ski-/- embryos. This possibility was attractive because the phenotype of Ski-/- embryos is reminiscent of Pax3-deficent Splotch mice and of Met-/- or HGF-/- mice which also show severe deficits in hypaxial limb muscle and less affected epaxial muscle. The selective muscle defects in these mutants result from impaired delamination and migration of myogenic progenitors (72, 73, 75, 77-79, 86, 88). Thus the hypaxial muscle defect in the

Ski-/- embryos might result from decreased expression of Pax3, Met and/or HGF.

To test this hypothesis, we examined Pax3 expression as an indicator of the formation, distribution, and migration of myogenic progenitors. Results of whole mount in situ hybridization of E9.5 and E10 embryos showed that the expected expression of

Pax3 in the dorsal neural tube was comparable in wild-type and Ski-/- embryos, as was the morphology of the Pax3+ dermomyotome (Figs. 9A-F, also Fig. 6). At both stages,

Pax3 was also expressed in myogenic progenitors that had delaminated and migrated into the forelimb bud of wild-type embryos. However, in the Ski-/- mutants Pax3+ cells had migrated to the proximal forelimb bud but did not extend as far into the limb bud as in the

102

103 Fig. 8. Normal level of proliferation and apoptosis in the myogenic progenitors or muscle cells in the absence of Ski. Transverse sections of Ski-/- embryos and normal littermates at E10.5 (A, B, H, I), E12.5 (C, D, J, K) and E15.5 (E, F, L, M) were analyzed for the numbers of mitotic cells (anti-p-H3, A-F) or apoptotic cells (anti-caspase 3, H-M) by immunofluorescence. Myogenic progenitors and myoblasts were labeled with anti-

Pax3 and anti-desmin antibodies, respectively. (G) Quantitative data are presented as the number of mitotic cells per unit muscle area (values above bar) and the values for mutant and wild type embryos were not significantly different (P>0.3). n is the number of samples. The numbers of apoptotic cells were too small to calculate significance.

104

105 Fig. 9. Impaired migration of myogenic progenitors in Ski-/- embryos. Expression of

Pax3 in wild-type and Ski-/- E9.5 (25 somites) and E10.0 (28 somites) embryos was analyzed by whole mount in situ hybridization. (A, B) At E9.5, Pax3+ cells invaded the

forelimb buds in the wild-type, but accumulated at the body/limb junctions in the mutant

(arrows). Note similar expression of Pax3 in the neural tube and dermomyotome of wild-

type and Ski-/- mutant (arrowheads). Lateral views of Ski-/- E10.0 embryos (D) show comparable expression of Pax3 in dermomyotome (black arrowheads) and neural tube

(white arrowheads) to that of the wild type (C). Whereas dorsal views of hindlimbs (E, F)

reveal reduced numbers and less distal extension of Pax3+ cells in the Ski-/- embryos than in the wild-type (arrows). (G, H) Immunofluorescence with anti-Pax3 of thoracic transverse sections of E10.5 embryos. In the wild-type, larger numbers of Pax3+ cells progressed dorsally and ventrally toward the distal forelimb than in the Ski-/- embryo

(arrows) where they accumulated at the dorsal limb/body junction (arrowheads). (I-K)

Quantification of Pax3+ cells in G and H shows their number in the forelimb (I); their

maximum extension as a percentage of the full length of the forelimb (J); and their

distribution in proximal, middle and distal regions. Values above the bars represent the

number or percentage of Pax3+ cells and the error bar represents the standard deviation, *

denotes P<0.05 and n is the number of samples. Scale bar (A-H) 50 μm.

106 wild-type (Figs. 9A and B). The difference in migration was also evident in the hindlimb

(Figs. 9E and F).

The distribution of myogenic precursors in E10.5 forelimbs of embryos was further analyzed by immunofluorescence with Pax3 antibody. Consistent with the mRNA expression data, we detected extensive invasion of the limb by Pax3+ myogenic progenitors in wild-type embryos, but in the Ski-/- embryos these cells appeared to accumulate at the body-limb junction (arrowheads, Figs. 9G and H also Fig. 6) and migrated into the distal limb in significantly reduced numbers (p<0.05) (Figs. 9G-K).

The difference was even more pronounced in the ventral half of the limb (arrows, Figs.

9G and H). By E11.0, migration of myogenic progenitors to the forelimb was completed and expression of Pax3 decreased thereafter (data not shown).

The apparent migration defect led us to examine the expression of Met and its ligand HGF/SF, which mediate delamination and migration of myogenic progenitors. At

E10.5, Met expression was detected in migratory progenitors in the lateral dermomyotome and forelimb buds of the wild-type (Figs. 10A and C), but was barely detectable in the lateral lip of the dermomyotome and notably reduced in limb buds of the

Ski-/- embryos (Figs. 10B and D). Unlike Met, HGF/SF expression in the limb bud mesenchyme, which provides a signal for migration into the limb, was indistinguishable between the Ski-/- mutant and the normal control (Figs. 10E-H).

Altered BMP signaling has also been implicated in defective Pax3+ progenitor accumulation (89). Moreover, Ski has been shown to regulate BMP signaling in cultured cells (28) and Ski expression in the AER (Fig. 5B) overlaps that of BMP (165). To assess a possible role of BMP in the hypaxial muscle defect of Ski-/- embryos, we examined the

107

108 Fig. 10. Loss of Ski affects the expression of Met but not HGF/SF or Msx1. Expression of

Met, HGF/SF and Msx1 in wild-type and Ski-/- E10.5 (33, 28, 29 somites) embryos was analyzed by whole mount in situ hybridization. Lateral (A, B) and dorsal (C, D) views show expression of Met at the lateral lips of dermomyotome (arrowheads) and in the limb buds (arrows) of the wild-type (A, C), but at appreciably lower levels in the Ski-/- mutant

(B, D). Lateral (E, F) and dorsal (G, H) views show HGF/SF is expressed at comparable

level along the migration route and at the target sites in the limb bud mesenchyme

(arrows) of wild-type (E, G) and Ski-/- mutant (F, H). (I, J) Lateral view of Msx1

expression reveals an indistinguishable pattern of BMP signaling in the wild-type and

Ski-/- mutant (arrows). Scale bar (A-J) 100 μm.

109 expression of Msx1, a direct transcriptional target of BMP signaling in the limb bud

(166). We found that the pattern and level of Msx1 expression was indistinguishable in

the limb buds of wild-type and Ski-/- embryos (Figs. 10I and J), suggesting that BMP

signaling in the limb bud is not affected by the loss of Ski. Taken together, our results

strongly suggest that in the absence of Ski, migration of hypaxial myogenic progenitors

into the limb is impaired due to the down-regulation of Met expression. This results in an

insufficient number of myogenic cells to support normal hypaxial muscle formation.

Normal specification of fetal/postnatal myogenic progenitors in Ski-/- embryos

Recent data support a model in which a pool of Pax3+/Pax7+ cells in the

dermomyotome is the source of both embryonic and fetal/postnatal myogenic progenitors

(70). We have shown that defective migration of the Pax3+ embryonic component of that

pool early in development is responsible for the underdevelopment of distal hypaxial

muscle in Ski-/- mice. At later developmental stages, the cells that comprise fetal

myogenic progenitors and satellite cells are marked by the expression of Pax7 (70). To determine whether the establishment of these cells at their target sites was also affected by the lack of Ski, we scored the number of Pax7+ cells in fetal muscles at E15.5 (Fig. 11)

and E18.5 (Fig. 12). We found that, despite their reduced size, individual muscles of Ski-

/- fetuses contained about the same total numbers of Pax7+ cells as the wild type, and at

E18.5 a large fraction of these cells assumed the satellite position under the basal lamina

(Figs. 11A and B and Figs. 12A-D). In fact, because the muscles in the mutant contained

fewer myotubes, the number of Pax7+ nuclei per myotube was higher in the Ski-/- fetuses than in the controls (Fig. 11C and Fig. 12E).

110

111 Fig. 11. Loss of Ski doesn’t affect in vivo accumulation or in vitro differentiation of fetal

myogenic progenitors. (A-D) Immunofluorescence of thoracic transverse sections of

E15.5 Ski-/- embryos and normal littermates with anti-laminin and anti-Pax7. In both

Ski+/+ and Ski-/- embryos a large fraction of Pax7+ cells are positioned under the basal

lamina of myotubes. Similar total numbers (A, B) of Pax7+ cells but increased density per

myotube (C numbers above bars) in individual muscles of Ski-/- embryos compared to controls. Error bars = standard deviation of multiple muscles in three independent experiments. (D-G) Myoblasts isolated from limb muscle of E15.5 embryos were plated in GM (D, E). When they reached ~60% confluence cells were switched to DM and photographed at 48 hours (F, G). Myotube formation by Ski-/- cells was comparable to that of the controls.

112

113 Fig. 12. Loss of Ski doesn’t affect in vivo accumulation or in vitro differentiation of satellite cells. (A-D) Immunofluorescence of thoracic transverse sections of E18.5 Ski-/- embryos and normal littermates with anti-laminin and anti-Pax7. In both Ski+/+ and Ski-

/- embryos a large fraction of Pax7+ cells are positioned under the basal lamina of

myotubes. Similar total numbers (A, B) of Pax7+ cells but increased density per myotube

(C, D, E numbers above bars) in individual muscles of Ski-/- embryos compared to controls. Error bars = standard deviation of multiple muscles in three independent experiments. (F-I) Myoblasts isolated from limb muscle of E15.5 embryos were plated in

GM (F, G). When they reached ~60% confluence cells were switched to DM and photographed at 48 hours (H, I). Myotube formation by Ski-/- cells was comparable to that of the controls. (J) Quantitative RT-PCR of Myf5 and Pax7 mRNAs from myoblasts

at ~60% confluence in GM revealed similar expression in Ski-/- and control myoblasts.

Relative mRNA levels (values above bars) were calculated as described in Materials and

Methods. Error bars = standard deviation in three independent experiments. (K) Myotube

formation was examined by immunostaining with anti-MHC at the indicated time after

switching from GM into DM. Values above bars = percentage of nuclei in MHC-positive

muscle cells; error bars = standard deviation, n= number of samples. Scale bar (A, B)

40 μm, (C, D) 5 μm.

114 Pax7+ progenitors are the source of myoblasts that can proliferate and differentiate

in culture (70). We therefore asked whether cells isolated from the muscle of Ski-/-

fetuses retained these abilities. We found that mutant and wild-type E15.5 and E18.5 limb

muscles gave rise to proliferating myoblasts that expressed Myf5 and Pax7 at comparable

levels (Figs. 11D, E and Figs. 12F, G, J) and differentiated into MHC+ myotubes

indistinguishably after two days of culture in DM (Figs. 11F, G and Figs. 12H, I, K).

Thus the Pax7+ cells of Ski-/- fetuses, like their Pax3+ embryonic counterparts, are fully

competent for terminal differentiation. However, migration of the fetal progenitors

appears to be independent of Ski function while that of the embryonic progenitors is

regulated by Ski.

Discussion

Previous publications that implicated Ski in the development and differentiation of skeletal muscle have provided evidence supporting seemingly disparate roles for Ski in

these processes (25, 42, 45, 114-116). Several studies suggested that Ski promotes post-

commitment muscle differentiation and growth. Ski expression increases during

differentiation of myoblasts in vitro and Ski transgenic mice develop an over-muscled phenotype that results from selective postnatal hypertrophic growth of fast muscle fibers

(45). Those results are consistent with the ability of Ski to co-operate with MyoD and activate the expression of genes involved in terminal muscle differentiation, such as

Myog, Myl1 and Ckm (114-116). However, our findings demonstrate that the terminal differentiation of both epaxial and hypaxial myogenic cells is unaffected by the loss of

115 Ski. We therefore conclude that Ski is not required for terminal differentiation in either of these myogenic lineages.

The earliest demonstration of Ski’s influence on muscle differentiation suggested

its role in myogenic determination by showing that overexpression of Ski in non-

myogenic fibroblasts promotes their trans-determination to the muscle lineage (25, 42).

This activity of Ski is analogous to that of the MRFs. Indeed Ski has been shown to be

capable of inducing the expression of MyoD and Myog, suggesting it acts upstream of

these genes in the myogenic pathway (25). Consistent with that model, the present work

demonstrates that Ski is expressed in the myogenic regions of somites, prior to the

determination of myogenic progenitors. However, neither the generation of Pax3+ progenitors in the dermomyotome nor their subsequent myogenic determination in the myotome and limb buds, marked by MRF expression, was affected by the loss of Ski.

Despite that conclusion, we observed a large reduction in the numbers of hypaxial progenitors in the distal limbs of Ski-/- embryos. This could have resulted from any combination of premature differentiation, reduced expansion, or impaired migration of these cells due to abnormal BMP signaling. Previous studies have demonstrated the potentially deleterious effect of premature differentiation by showing that in the absence of BMP signaling, myogenic progenitors switch off Pax3 expression, turn on MyoD and lose their ability to migrate and proliferate (148). This does not seem to be the case in Ski mutants because activation of MRF expression was only observed in the myotome but not in the dermomyotome and we did not detect an excess of MRF positive cells along the migration route. Although abnormal BMP signaling could also affect the level of proliferation or apoptosis (89), neither of these processes was altered in the progenitors in

116 the dermomyotome or limb buds of Ski-/- embryos. Furthermore, unchanged Msx1

expression in the Ski-/- embryos indicates that the progenitor cell defect is not due to deregulated BMP signaling.

Having ruled out the other likely possibilities, our results suggested that defective migration was the underlying cause of reduced hypaxial muscle in Ski mutant mice. In support of this model, we found that dense clusters of Pax3-expressing cells accumulated

in close proximity to the lateral dermomyotomal lip and at the body wall/limb junction of

the mutant embryos. This phenotype is similar to that of Met-/- and HGF/SF-/-embryos,

which subsequently also exhibit selective defects in hypaxial muscle development due to

impaired migration of progenitor cells (78, 79). Therefore, defective migration in the Ski-

/- mutant could be caused by decreased Met activity or expression in myogenic

progenitors or by reduced expression of its ligand, HGF/SF. Our results favor the former possibility by showing normal HGF/SF expression but dramatically reduced Met

expression. Thus it appears that Ski regulates migration of myogenic progenitors by

direct or indirect control of Met expression.

Ski does not bind DNA but acts as a transcriptional co-regulator of DNA-binding

transcription factors (26, 31, 163). It therefore might be expected that the absence of Ski

would only cause a partial loss of function of these factors or the genes they regulate.

Thus our observation that the Ski null mutation produces a partial phenocopy of Pax3 and

Met mutants which produce more severe deficiencies in hypaxial myogenesis is in

keeping with Ski’s transcriptional function (77, 78). Early in development Ski expression

in the dermomyotome coincides with the expression domain of Pax3 (72, 73, 75). Thus

an attractive possibility would be that Ski regulates Pax3 expression in myogenic

117 progenitors. However, we found that Pax3 expression in this domain is not affected by

the loss of Ski. Thus it is more likely that Ski acts to regulate genes downstream of Pax3.

Since expression of Met is barely detectable in the lateral dermomyotome of the Splotch

mutant, Pax3 has been suggested to directly control the migration of myogenic

progenitors by transactivation of Met expression (73, 76, 86, 88). The down regulation of

Met expression in Ski-/- embryos described here supports the possibility that Ski normally

functions as a co-factor with Pax3 to activate Met expression. This model is appealing in

light of recent findings that implicate the Ski family member, Dachshund (11), as a co-

regulator with the Drosophila homolog of Pax3 (56, 58).

Although the possible cooperation between Pax3 and Ski to activate genes required for migration remains to be clarified, other similarities in their mutants are consistent with this model. Mice lacking Ski displayed not only defective myogenesis, but also severe defects in patterning of craniofacial and cardiac structures derived from the cranial neural crest, such as midline facial clefting, a depressed nasal bridge (47) and defective development of the cardiac outflow tract (C. Colmenares, unpublished observations). Interestingly, Pax3 mutant Splotch mice, display defects in the same neural crest derivatives, and these have been linked to defective migration of the neural crest precursors (167-169). Thus, despite their distinct embryological origins, aberrant neural crest development and impaired myogenic progenitor development may reveal a common mechanism that is subject to regulation by both Ski and Pax3.

A surprising finding in this study is the contrast between the effect of Ski loss on embryonic myogenic progenitors and its lack of effect on fetal and adult progenitors. A large body of data indicate that these progenitor populations originate from a common

118 pool of Pax3+Pax7+ cells in the dermomyotome (70), but as development proceeds they

become distinct (108, 170). During early development Pax3+Pax7- embryonic progenitors

give rise to MRF+ myoblasts at distal locations which differentiate into primary and

secondary muscle fibers. These progenitors are depleted in the absence of Ski. Later in development, progenitors expressing mainly or exclusively Pax7 are found within these muscles. Some of these cells commit to terminal differentiation but many retain their progenitor status and become satellite cells. Their Ski-independent development provides additional support for their divergence from the Pax3+Pax7- embryonic progenitors of earlier development. Future studies of Ski mutant mice may help to identify the factors responsible for the generation of these two populations from a common precursor.

Future Directions

Dissecting the roles of Ski during development

Our work demonstrates that Ski regulates the migration of embryonic myogenic progenitors, but not those of the fetal or adult myogenic progenitors. The most important questions remaining are what is/are the difference(s) between migratory embryonic progenitors and migratory fetal/adult progenitors and how Ski specifically regulates

migration of embryonic myogenic progenitors? Various cell-surface adhesion molecules which regulate cell-cell adhesion or cell-matrix adhesion mediate the ability of cells to

detach from a cohesive structure, undergo long-range migration and contribute to the

formation of tissues during development (171). Identification of the cell-surface adhesion molecules that are differentially expressed in embryonic, fetal and adult myogenic progenitors will enable further determination of how Ski is involved in progenitor

119 migration at different development stages. Several recent studies have indicated that the

HGF/SF receptor, Met, can interact with the hyaluronon receptor CD44, α2β1 integrin,

β4 integrin, ezrin, the Fas receptor, semaphoring receptors, β-catenins or E-cadherin at

the cell membrane to mediate cell-cell and cell-matrix adhesion (172-178). These

interactions are potentially important for the specificity or robustness of Met signaling. It

is possible that the different effects of absence of Ski on migration are mediated through

these adhesion molecules.

To study the affect of Ski on migration, we will analyze adhesion molecule

expression in specific subsets of myogenic cells. embryonic and fetal myogenic

progenitors differ in their expression of Pax proteins: embryonic and fetal myogenic

progenitors are either Pax3+ or Pax3+Pax7+ whereas adult myogenic progenitors are

Pax7+ only (108-110, 179, 180). Based on this, myogenic progenitors will be harvested at

different developmental stages and sorted by flow cytometry with anti-Pax3, and anti-

Pax7 antibodies. The expression profile of adhesion molecules in these cells will be generated by microarray. Genes which are differentially expressed in these cells will be confirmed by realtime-PCR and further by double immunofluorescence in vivo at different developmental stages. This comparison will hopefully give some insight into genes regulated by Ski that control progenitor migration.

The mouse model used in this study has Ski knocked out in all tissues constitutively and exhibits the overall consequences of the loss of Ski throughout development. Taking advantage of the cell type-specific and stage-dependent activation of Pax3, Pax7, Myog promoters (85, 117), mice with Ski ablated in certain myogenic cells at certain stages will be generated by crossing mice with a floxed Ski allele to

120 animals that transgenically express Cre recombinase under the control of the Myog

promoter (Myog-Cre)(181), Pax3 promoter (Pax3-Cre) (182) or Pax7 promoter (Pax7-

Cre) (183). These mouse models will be used to investigate the role of Ski in the

regulation of myogenic progenitor behavior during early development and myogenic cell

behavior during mid and late development. They also can be used to study whether the

effect of Ski on myogenic cell behavior is cell-autonomous or cell-nonautonomous.

Mechanism underlying the regulation of Met by Ski

My study indicated that defective progenitor migration observed in Ski-/- mice

was not due to decreased HGF signaling to activate Met tyrosine receptor, but likely

caused by a reduced transcript level of Met as evidenced by in situ hybridization analysis.

The key experiment to confirm this statement would be to reintroduce a transgene

carrying Met into Ski null mice to restore the migratory ability of Pax3+ myogenic

progenitors.

Previous studies have implicated several transcription factors in the regulation of

Met in fibroblasts and epithelial cell lines (86, 184-187). However, the Met promoter is

regulated in a cell-type specific manner (188) and little is known about how Met is

controlled in myogenic cells except that its expression is modulated by Pax3 in C2C12

myoblasts (86). In a future study, it will be important to determine the molecular

mechanism underlying the regulation of Met by Ski in myogenic progenitors, including

the regulatory region of Met promoter that mediates Ski’s activity and the DNA-binding

partner for Ski at these sites. Since techniques to isolate and culture myogenic progenitors in vitro are not available, we will try to address this issue in the cultures of C2C12 myoblasts which were isolated from adult mouse muscle (152). The disadvantage of this

121 model system is that since loss of Ski only compromised the migration of embryonic

myogenic cells and not their fetal and adult counterparts, the data collected from cultures

of C2C12 cells may not faithfully recapitulate the role of Ski in embryonic myogenic

cells. However, since these cells are of myogenic origin (152) and have been shown to

migrate in response to HGF (189-192). Additionally, overexpression of Pax3 was correlated with enhanced Met expression (86) in these cells. Therefore, they are the best in vitro model to address this issue so far.

The regulation of Met transcription by Ski could be direct or through Ski’s ability to regulate the expression of Pax3. To date, Pax3 is the only transcriptional factor known to bind the Met promoter, activate Met transcription and is co-expressed with Met in myogenic progenitors (73, 76, 83-86, 88). C2C12 cells in which Ski is overexpressed or knocked down will be used to evaluate the effect of Ski on the expression of Pax3 and

Met. These cells will also be tested in cell migration assays in response to HGF. In addition, the direct association of Ski with the endogenous Pax3 or Met promoters will be assessed by ChIP assays. Reporter assays will be used to further identify the minimal regulatory regions that mediate the regulation of Pax3 or Met promoters by Ski.

Since Ski can not bind to DNA directly (31), to further understand the transcriptional regulation of Pax3 or Met by Ski, it is important to identify the DNA- binding partners which recruit Ski to these promoters. If Ski directly regulates the transcription of Pax3, Hox, Pbx and Meis protein families will be examined because they have been shown to regulate Pax3 expression and interact with Pax3 promoter through

their consensus binding elements (193). They are all candidate DNA-binding partners

that might recruit Ski to the Pax3 promoter. The interaction of Ski with these proteins and

122 their association on the endogenous Pax3 promoter will be examined by

immunoprecipitation and sequential ChIP assay, respectively. The key experiment to

demonstrate that Pax3 is a mediator of transcriptional regulation of Met by Ski will be to

reintroduce Pax3 into C2C12 cells in which Ski is knocked down and eventually into Ski

null mice to rescue Met expression and myogenic progenitor migration.

Pax3 bind the Met promoter directly and it also cooperates with Six, Eya and a

Ski family member, Dach to regulate myogenesis during early development (56, 58). We know from our own study that Ski substitutes for Dach to cooperate with Six1/Eya3 on transactivation of Myog during C2C12 differentiation. If Ski directly regulates Met

transcription, it will be of great interest to investigate whether Ski also substitutes for

Dach in the conserved network of Pax3/Six/Eya/Dach and it is Pax3 that recruits Ski to

the Met promoter and regulates Met expression. Synergetic regulation of transcription

from Met promoter by these proteins will be examined by reporter assay and their

association with the endogenous Met promoter will be assessed by sequential ChIP assay.

Immunoprecipitation assays will also be used to determine the ability of these proteins to

physically interact with each other. The key experiments to establish Pax3 as a DNA-

binding partner of Ski on the Met promoter will be to examine whether Ski is still able to

bind and transactivate the Met promoter in C2C12 cells lacking Pax3 expression.

Because of the disadvantages of using the adult satellite-cell derived C2C12 cells,

the data collected from cultures of C2C12 cells will be confirmed in mouse models.

Transgenic mice carrying a LacZ gene under the control of Pax3 or Met promoters will

be introduced into wild-type and Ski-/- mice and a comparison of LacZ expression in

123 Pax3+ myogenic progenitors in these transgenic lines will demonstrate whether Ski also transactivates the Pax3 or Met promoters in myogenic progenitors.

124 CHAPTER II: Ski Regulates Muscle Terminal Differentiation by Transcriptional

Activation of Myog in a Complex with Six1 and Eya3

This chapter has been submitted to Journal of Biological Chemistry

Abstract

Overexpression of the Ski pro-oncogene has been shown to induce myogenesis in

non-muscle cells, to promote muscle hypertrophy in postnatal mice and to activate

transcription of muscle-specific genes. However, the precise role of Ski in muscle cell

differentiation and its underlying molecular mechanism are not fully understood. To

elucidate Ski’s involvement in muscle terminal differentiation, two retroviral systems

were used to achieve conditional overexpression or knockdown of Ski in satellite cell-

derived C2C12 myoblasts. We found that enforced expression of Ski promoted differentiation while loss of Ski severely impaired it. Compromised terminal differentiation in the absence of Ski was likely due to the failure to induce Myog and p21, despite normal expression of MyoD. Chromatin immunoprecipitation and transcriptional reporter experiments showed that Ski occupied the endogenous Myog regulatory region and activated transcription from the Myog regulatory region upon differentiation.

Transactivation of Myog was largely dependent on a MEF3 site bound by Six1; not on the binding site of MyoD or MEF2. Activation of the MEF3 site required direct interaction of Ski with Six1 and Eya3 mediated by the evolutionarily conserved DHD domain of Ski. Our results indicate that Ski is necessary for muscle terminal differentiation and that it exerts this role, at least in part, through its association with Six1 and Eya3 to regulate the Myog transcription.

125 Introduction

Originally identified as a transduced avian retroviral oncogene (1-3), Ski is an evolutionarily conserved gene in species ranging from flies to humans (4). The retroviral v-ski protein corresponds to residues 21–441 of chicken c-Ski which is a nuclear protein of 750 residues (5). The amino terminal half of c-Ski is the most highly conserved segment, containing two distinct domains that mediate protein-protein interactions (6).

The more conserved of these, the Dachshund homology domain (DHD), defines the Ski

gene family which includes: Ski, SnoN, Dach, Fussel-15, Fussel-18 and Corl (7, 9-15).

The DHD has been implicated in the interactions of Ski with Smad2/3, N-CoR, Skip and

RARα (16-24). The second conserved region comprises the SAND domain (Sp100,

AIRE-1, NucP41/75 and DEAF-1 domain) which mediates the interactions of Ski with

Smad4, FHL2 and MeCP2 (8, 28, 29). The C-terminal region of c-ski is missing in v-ski

and contains a tandem repeat/leucine zipper motif which mediates both homo- dimerization and hetero-dimerization with SnoN (6, 30).

Ski does not bind DNA directly (31), but interacts with several different transcription factors to modulate transcription as either a co-activator or a co-repressor

depending on its DNA-binding partner. Both v-ski and c-ski cause oncogenic transformation and induce myogenic differentiation in non-muscle avian embryo fibroblasts (2, 3, 149). The latter activity involves activation of muscle-specific genes, including the myogenic regulatory factor (MRF) genes, MyoD and myogenin (Myog) (25,

42, 149). Transgenic mice expressing v-ski or c-ski cDNAs develop selective hypertrophy of type IIb fast skeletal muscle fibers (45). Moreover, expression of Ski increases in skeletal muscle at mid-gestation of mouse development (39). The requirement of the

126 endogenous gene for normal muscle development was demonstrated by the observation that Ski-null mice show a marked decrease in skeletal muscle mass (47).

Further studies indicated that regulation of myogenesis by Ski might be mediated by its ability to activate transcription driven by the regulatory elements of muscle-specific genes, such as myosin light chain 1/3 (MLC 1/3), the muscle creatine kinase (MCK) and most importantly, Myog (114-116). The Ski-responsive cis element of Myog resides in an

184bp regulatory region immediately upstream of the Myog promoter. This region has been shown to be sufficient for the complete recapitulation of the temporal and spatial expression pattern of Myog during embryogenesis (115, 117-119). DNA binding sites for

MyoD and myocyte enhancer binding factor 2 (MEF2) in this regulatory region were found to be necessary for Myog transcription (119-121). Transient reporter assays have shown that Ski co-operates with MyoD and MEF2 to activate transcription from the

184bp Myog regulatory region (115). However, no direct interactions between Ski and these muscle-specific transcription factors have been reported.

In vitro studies have revealed that terminal differentiation of myoblasts proceeds through a highly ordered sequence of events. These cells express MyoD while proliferating but when growth stimuli are removed, they initiate expression of Myog, followed by the induction of the cyclin-dependent kinase (cdk) inhibitor p21 and irreversible withdrawal from the cell cycle. Subsequently, these post-mitotic myocytes express muscle-specific contractile proteins such as myosin heavy chain (MHC) and finally fuse into multinucleated myotubes (128). This process is governed mainly by two families of transcription factors, the MRFs and MEF2 (129-131). The MRF gene family includes MyoD, Myf5, Myog and MRF4 (91, 92, 129). All MRF family members share a

127 highly conserved basic region and adjacent helix-loop-helix motif (bHLH) which mediates binding to a consensus DNA sequence CANNTG, known as the E box, which is

present in the regulatory regions of many muscle specific genes. Forced expression of

any MRF gene is capable of inducing expression of muscle-specific genes and activation

of myogenic differentiation, even in non-muscle cells. The MEF2 proteins belong to the

superfamily of MADS (MCM1-agamous deficient-serum response factor)-box

transcription factors, and directly bind an A+T-rich element found in the promoters and enhancers of many muscle specific genes (133). Genetic analysis reveals that members of the MEF2 family are also essential for terminal muscle differentiation (132).

Another cis-element, the MEF3 site (consensus sequence TCAGGTT), is also

present in the 133 bp Myog regulatory region. Studies of transgenic mice demonstrated

that mutation of this MEF3 site abolishes correct expression of a Myog-LacZ transgene

during embryogenesis (123). Two skeletal-muscle specific members of the Six family

(sine oculis homeodomain-containing transcription factors), Six1 and Six4, bind to the

MEF3 element and transactivate Myog transcription (123). Drosophila sine oculis (so) has been shown to act synergistically with eyes absent (eya) and dachshund (dac) by direct protein-protein interactions. Similar interactions underlie the synergism of their mammalian homologues, Six, Eya and Dach (11, 55-58, 134, 137, 139, 143). This evolutionarily conserved regulatory network of Eya/Six/Dach has been shown to regulate

myogenesis in chicken somite culture and in the chick limb and to activate transcription

of reporters containing the Myog MEF3 site (56, 58). Interaction of mammalian Dach

with Six protein is mediated by the evolutionarily conserved DHD motif (55, 151). This

has led to the suggestion that by virtue of its possession of these conserved domains (7,

128 11, 13), Ski might also interact with Six and Eya proteins to regulate Myog expression

and thereby control commitment of myogenic cells to terminal differentiation (56).

In this study, we have addressed this possibility by exploiting the well-

characterized mouse muscle satellite cell line, C2C12, as a model system. Using

retroviral vectors to achieve tetracycline (tet)-regulated overexpression or knockdown of

Ski, we asked whether Ski might not only stimulate but also be required for terminal differentiation of C2C12 cells. To probe the mechanism underlying the transcriptional regulation of Myog by Ski, the E-box, MEF2 and MEF3 sites in the Myog regulatory region were investigated as the possible cis-response elements. Using co- immunoprecipitation, we asked whether direct binding of Ski to MyoD, MEF2c, Six1 and/or Eya3 mediate its transcriptional activity. Finally, ChIP assays were performed to determine whether Ski resides at the endogenous Myog regulatory region prior to or concomitant with the initiation of terminal differentiation.

Materials and Methods

Construction of retroviral vectors

The replication-defective retroviral vector LNITX was kindly provided by Dr.

Fage and was described earlier (158). A DNA fragment containing the entire coding region of the human SKI cDNA was excised from the plasmid pSHHSKIN1D using

BamHI and ClaI, blunt-ended with T4 DNA polymerase, and inserted into the LINTX vector at the PmeI site to produce LNIT-huSKI.

The replication-defective retroviral vector TMP-tTA was modified from the TMP vector (kindly provided by Dr. Scott Lowe) (159-161) by replacing its GFP gene with the

129 tetracycline transactivator (tTA, ) gene (194). The sequences of shRNAs targeting mouse

Ski were chosen using RNAi Codex (http://codex.cshl.edu/scripts/newmain.pl) and are designated by their positions in the mouse Ski cDNA sequence (GenBank AF435852): mSki1145 (bases 1145-1163), mSki1819 (bases 1819-1837) and mSki977 (bases 977-

995). DNA forms of these shRNA inserts were generated by PCR amplification of 97

base synthetic oligonucleotides using Pfu DNA polymerase (Invitrogen) and a common

set of primers (5'-cagaggctcgagaaggtatattgctgttgacagtgagcg-3' and 5'-

cgcggcgaattccgaggcagtaggca-3'). The PCR products were subsequently digested with

XhoI and EcoRI and inserted between these sites within TMP-tTA vector to generate

TMP-tTA-mSki1145, TMP-tTA-mSki1819 and TMP-tTA-mSki977. Clones containing these shRNA-encoding inserts were sequence-verified.

Tissue culture and transfection

Proliferating mouse C2C12 myoblasts were maintained in growth medium (GM) consisting of Dulbecco's modified Eagle's medium (DMEM) (Gibco-BRL) supplemented with 20% fetal bovine serum (FBS, Atalantic Biosolution), 100μg/ml penicillin,

100units/ml streptomycin and 0.002% Fungizone (Gibco-BRL). To avoid spontaneous differentiation, cells were always kept in subconfluent (60-70%) conditions. Terminal

differentiation was induced by switching subconfluent cell (80%) to differentiation

medium (DM) consisting of DMEM, 2% heat-inactivated horse serum plus antibiotics as in GM. Morphological differentiation, judged by myotube formation, was documented by digital photography of phase-contrast microscopic images.

Retroviral packaging and infection

130 The retrovirus packaging cell line, PA317 (ATCC No. SD3443) was cultured in

DMEM containing 10% FBS, 100μg/ml penicillin and 100units/ml streptomycin. They were transfected with retroviral constructs by using Fugene 6 (Roche) according to the manufacturer's protocol. 24 hr after transfection, the medium was harvested and live cells were removed by centrifugation at 1000×g for 10min. The retrovirus-containing supernatant was used to infect exponentially growing C2C12 cells at 40% confluence in the presence of 10ng/ml polybrene (Sigma). The infection was repeated with freshly- harvested virus 12hr later. After an additional 12hr of culture, C2C12 cells were switched to GM plus 2μg/ml doxycycline (Dox, an analog of tetracycline) and antibiotics (G418 at

750μg/ml for LNITX-based constructs and puromycin at 2ng/ml for TMP-tTA-based constructs). Two weeks later the resulting individual colonies were isolated using cloning discs according to manufacturer’s protocol (PGC Scientifics, 62-6151-14) and expanded in GM with Dox and G418 or puromycin. After culture in GM minus Dox, clones were screened by western blot analysis for tet-regulated Ski overexpression or knockdown.

Western blotting

Whole-cell extracts were prepared from confluent C2C12 cells on 100 mm culture dishes as follows: cells were scraped into 400µl of lysis buffer (50mM Tris pH 8.0,

100mM NaCl, 1mM EDTA, 10% glycerol, 0.2% Nonidet P-40, 0.1mM sodium orthovanadate, 5mM sodium pyrophosphate, 1mM NaF and complete protease inhibitor cocktail (Roche)). After 10min incubation on ice, the suspension was subjected to three freeze-thaw cycles and cell debris was removed by centrifugation at 12,500 rpm for

15min at 4°C. Protein concentrations were determined by Bradford assay (Bio-Rad) and equal amounts of proteins were boiled in protein loading buffer (100mM Tris pH6.8,

131 20% glycerol, 4% SDS, 0.2% Bromophenol blue and 2%β-Me), separated by 6% SDS–

PAGE and transfered to Immobilon-P membranes (0.45um, Millipore).

For most antibodies, blots were pre-blocked for 1hr at room temperature in

blocking buffer (20mM Tris pH 7.6, 125mM NaCl, 0.1% Tween-20 and 5% w/v nonfat

dry milk) and incubated with primary antibodies in the same solution overnight at 4°C.

The membranes were then washed with TBST (20mM Tris pH 7.6, 125mM NaCl and

0.1% Tween 20), and incubated for 1hr at room temperature with secondary antibodies

conjugated to horseradish peroxidase (HRP) or alkaline phosphatase (ALP) followed by

washes with TBST. Signals were detected using either enhanced chemiluminescent HRP

substrate (Pierce) or CDP-Star® alkaline phosphatase chemiluminescent substrate (1/50,

Tropix) and exposure to HyBlot CL autoradiography film (Denville Scientific).

When using anti-Ski G8 monoclonal antibody, the blots were pre-blocked overnight at 4°C in blocking buffer (PBS (7.7mM Na2HPO4, 2.7mM NaH2PO4, 150mM

NaCl, pH7.2), 0.4% Casein, 1% PVP (40,000), 10mM EDTA and 0.2% Tween, pH7.2) and then incubated with primary antibody in the same solution for 30min at room

temperature. The membranes were washed with blocking buffer and further incubated

with ALP-conjugated secondary antibodies (Sigma) for 30min at room temperature.

Subsequently, the blots were washed with blocking buffer then with assay buffer (0.1M

EDTA, 1mM MgCl2 and 0.02% Azide, pH10.0) and the signal was detetected with CDP-

Star® alkaline phosphatase chemiluminescent substrate as above.

The following primary antibodies were used for immunoblotting: Ski

(monoclonal antibody, 1/2000, G8, Learner Research Institute Hybridoma Core Facility

or rabbit polyclonal antibody, 1/1000, H329, Santa Cruz); Myog (1/1000, F5D, Santa

132 Cruz), TFIIE-α (1/3000, C-17, Santa Cruz), MHC (1/1000, MF-20, Developmental

Studies Hybridoma Bank), MyoD (1/1000, 5.8A, Santa Cruz) and anti-Flag (1/5000, M2,

Sigma). The secondary antibodies used were: ALP-conjugated anti-mouse IgG (Fc

specific) and anti-rabbit IgG (whole molecule) (1/30000; Sigma); HRP-conjugated

Mouse IgG TrueBlotTM and Rabbit IgG TrueBlot™ (1/1000, eBioscience)

Immunofluorescence

C2C12 myoblasts (2×104) were seeded into each well of LAB-TEK® eight-well

chamber slides (Nunc) coated with 1µg/cm2 laminin (Invitrogen) in GM and switched to

DM for 3 days. Cells were fixed in 3.7% paraformaldehyde/PBS for 30min at room

temperature, permeabilized with 10% goat serum/1% Triton-X 100/PBS for 10min,

incubated with blocking buffer (PBS, 10% goat serum and 0.1% Tween 20) for 1hr at

room temperature and then with primary antibodies overnight at 4°C. Chambers were washed with PBST (PBS and 0.1% Tween-20), incubated with secondary antibodies for

1hr at room temperature, washed with PBS and mounted in Vectashield aqueous mounting medium with DAPI (Vector Laboratories). Images were obtained using

Olympus BX50 upright fluorescence microscope equipped with a Polaroid digital camera

PDMC2, Polaroid PDMC2 software and fluorescent illumination. Images were

assembled using Photoshop CS (Adobe).

The following primary antibodies were used for immunofluorescence: MyoD

(1/100; 5.8A, Santa Cruz); Myog (1/100; F5D, Santa Cruz), MHC (1/200; MF20;

Developmental Studies Hybridoma Bank) and p21 (1/100; SX118, BD Pharmingen).

Secondary antibodies were Alexa 488 or Alexa594-conjugated goat anti-mouse IgG

133 antibody (1/300, Molecular Probes). Control experiments performed with normal IgG as the primary antibody yielded no signal above the background.

Quantification was performed by counting at least 1000 DAPI-stained nuclei in

more than 10 random fields per culture plate. For MHC, the differentiation index = nuclei

within MHC-stained multinucleate myotubes/total number of DAPI-stained nuclei and

the fusion index = the average number of nuclei per MHC-stained myotube. For nuclear

proteins, the differentiation index=number of antibody-stained nuclei/total number of

DAPI-stained nuclei. All experiments were performed in triplicate on three independent

cultures and the standard deviation was calculated.

Realtime-PCR

RNA was isolated using the RNeasy kit (Qiagen) with DNase I treatment

according to manufacturer’s protocol, and cDNA was generated using reverse

transcriptase SuperScript™ III (Invitrogen) with random hexamer primers according to the manufacturer’s instructions. Quantitative realtime-PCR (iCycler iQTM, Bio-Rad) was

performed using SYBR® Green PCR Core Reagents (Applied Bioscience) according to

the manufacturers’ protocols. PCR was performed for 40 cycles of 94 °C for 15s, 60°C

for 20s, and 72°C for 20s, followed by a single 72°C extension step for 5 min. Primer

sequences used for realtime PCR will be provided upon request. Analyses were

performed in triplicate on RNA samples from three independent experiments. Threshold

cycles (Ct) of target genes were normalized against the housekeeping gene Gapdh, and

relative transcript levels were calculated from the Ct values as Y=2-ΔCt, where Y is fold

difference in amount of target gene versus Gapdh and ΔCt=Ctx-CtGapdh.

Chromatin immunoprecipitation (ChIP)

134 Chromatin immunoprecipitation experiments were performed essentially as

described (195). Briefly, C2C12 cells cultured in GM or DM for 2 days were cross-linked with 1% formaldehyde/PBS for 10min at room temperature. Fixed cells were scraped and resuspended at 2×107 cells/ml in lysis buffer (50mM Tris pH8.0, 2mM EDTA, 150mM

NaCl, 1% Triton-100, 0.1% DOC, 0.1% SDS). Suspensions were sonicated to yield

chromatin with an average DNA length of 200~500bp. Equal amounts of chromatin from

each sample (2×106 cells per assay) were preabsorbed at 4°C for 1hr with 40μl 50%

slurry of pre-blocked protein A beads (Repligen, previously incubated with 1mg/ml

salmon testes DNA, 10mg/ml BSA and 0.05% sodium azide in TE buffer (10mM Tris

pH8.0 and 1mM EDTA)). After pelleting the beads, supernates were incubated overnight at 4°C with either 2μg rabbit polyclonal Ski antibody (H329, Santa Cruz) or normal

rabbit IgG. Antibody-chromatin complexes were then captured by incubation with 40μl

50% slurry pre-blocked Protein A beads at 4°C for 1hr. The beads were washed

sequentially in lysis buffer, high salt buffer (50mM Tris pH8.0, 2mM EDTA, 500mM

NaCl, 1% Triton-100, 0.1% DOC and 0.1% SDS), lithium salt buffer (20mM Tris pH8.0,

1mM EDTA, 250mM LiCl, 0.5% NP-40, 0.5% DOC) and TE buffer. Complexes were

eluted from the beads in 150μl of elution buffer (10mM Tris pH 8.0, 5mM EDTA and

1% SDS) and formaldehyde cross-linking was reversed by overnight incubation at 65°C.

After treatment with RNase A and proteinase K, DNA was isolated by phenol extraction

and ethanol precipitation. The optimal PCR cycle numbers were determined by realtime-

PCR and 5% of purified DNA was analyzed by regular PCR using HotStart-IT Taq

Master mix (USB). 25% of each reaction mixture was resolved on a 1.5% agarose gel and

visualized by ethidium bromide staining. The following PCR primer sets were used: the

135 mouse Myog regulatory region (-169~+39, Genbank M95800): 5′-gggcaaaaggagagggaag-

3′ and 5′-agtggcaggaacaagcctt-3′; the non-promoter region downstream of Myog gene

(+1943~+2185 downstream of the Myog gene, GenBank NW_001030662.1) served as a

negative control: 5′-gtcaagaactgacttaaggcc-3′ and 5′-gacactaggagaagggtggag-3′; the

mouse Smad7 regulatory region (-274~-142, GenBank NW_001030635.1): 5-

tagaaacccgatctgttgtttgcg-3 and 5-cctctgctcggctggttccactgc-3. For input control, 10% of

cross-linked chromatin was purified as described above and assessed for PCR by using

the same sets of primers.

Reporter assays

A 202bp DNA fragment (-184~+18, Genbank M95800) containing the Myog

regulatory region was amplified by PCR using C57BL/6 genomic DNA as template.

Myog184 luciferase reporter carrying the luciferase gene downstream of this Myog regulatory region was generated by inserting HindIII-StuI digested PCR product into the

HindIII-SmaI site of the pGL3-Basic vector (Promega). Myog-luciferase constructs with

E box, MEF2 and MEF3 mutations (Myog184-E1E2m, Myog184-MEF2m and Myog184-

MEF3m) were generated from Myog184 luciferase reporter using the QuikChange® Site-

Directed Mutagenesis Kit (Stratagene). Ski expression vector pCDNA-huSki was

previously described (24).

Cells (1×105) were seeded in 12-well plates and transfected at 80% confluence

using Lipofectamine 2000 (Roche) with a combined total of 4µg of expression vector and

reporter plasmid DNAs. 18hr after transfection, cells were switched to DM for 48hr prior

to harvest and lysis in 1×Reporter lysis buffer (Promega) with one round of freeze-thaw

followed by incubation at room temperature for 20min. Cell debris was removed by

136 centrifugation and luciferase activity in the supernatant was determined by a dual

luciferase reporter assay system (Promega), according to the manufacturer’s protocol

using a MAXline microplate luminometer (Molecular Devices). The relative light units

(RLUs) were generated by normalizing firefly luciferase units to Renilla luciferase units

of the co-transfected pTK-Renilla-luc vector. The experiments were done in duplicate

and the reported results represent at least three independent experiments.

Co-immunoprecipitation

C2C12 cells (~ 50% confluent, 100mm dishes) were transfected with expression

plasmids for flag-tagged Six1, flag-tagged Eya3, flag-tagged Mef2c, MyoD and full-

length Ski or its mutants using Lipofectamine 2000. 18hr after transfection, cells were

refed GM or switched to DM and harvested 2 days later in 1ml of NETN buffer (20mM

Tris pH 8.0, 100mM NaCl, 1mM EDTA, 0.1% NP-40, 10% glycerol, 1mM dithiothreitol

and complete protease inhibitors cocktails). After brief sonication, cell debris was

removed by centrifuge at 10,000 rpm for 15min at 4°C. The lysates were preabsorbed

with protein A beads and protein concentrations were determined by the Bradford protein

assay. Immunoprecipitations using equal amounts of proteins with either rabbit

polyclonal anti-Ski or normal rabbit IgG were collected by overnight incubation with

protein A beads (Repligen) at 4°C. Precipitates were washed 5 times in NETN,

resuspended and boiled in protein loading buffer, separated by 6% SDS-PAGE and

analyzed by Western blotting. If precipitating and primary Western blotting antibodies

were from the same species, either HRP-conjugated Mouse IgG TrueBlotTM or Rabbit

IgG TrueBlot™ was used as the secondary antibody accordingly.

137 Results

Regulated over-expression of Ski stimulates terminal differentiation of C2C12 cells

Overexpression of Ski has been reported to induce non-muscle fibroblasts to

differentiate into myotubes (25, 42), suggesting a role of Ski in myogenic lineage

determination and terminal differentiation. To assess the role of Ski in terminal

differentiation independent of its possible role in myogenic lineage determination, C2C12

cells which have already committed to myogenic fate and require only serum deprivation

to undergo terminal differentiation were used in this study. In order to regulate Ski

expression, C2C12 cells were infected with LNIT-huSKI; a retroviral vector that allowed

both G418 selection and doxycycline (Dox) regulation of SKI expression, (Fig. 13A).

Several G418-resistant clones were isolated and propagated in growth medium (GM) containing Dox to suppress SKI overexpression. They were then subdivided and tested for SKI expression after three-day cultures in either the presence or absence of Dox.

Western blot analysis of cell extracts revealed obvious induction of SKI expression in several LNIT-huSKI clones after Dox withdrawal (Fig. 13B), whereas a clone infected with the LNITX empty vector did not overexpress SKI, regardless of Dox treatment (Fig.

13D, right panel).

Using these clones, we examined the effect of SKI expression on myotube formation. LNIT-huSKI cells were cultured in GM with or without Dox for 4 days. Upon reaching 80% confluence, the cells were induced to differentiate by switching to differentiation medium (DM) while maintaining the presence or absence of Dox. One day after initiating differentiation, LNIT-huSKI cells that expressed exogenous SKI had prematurely formed myotubes, but no myotubes were observed in the same clone

138

139 Fig. 13. SKI stimulates differentiation of C2C12 myoblasts

(A) Schematic representation of retroviral vector LNIT-huSKI showing the 5′- and 3′-

long terminal repeats (LTR), neomycin resistance gene (neor), internal ribosomal entry

site (IRES), tetracycline-controlled transactivator gene (tTA) (tTA, ); the minimal CMV

promoter regulated by tetracycline response elements (TRE-CMV) and human SKI

coding region (SKI).

(B) Dox-regulated SKI expression in LNIT-huSKI clones. Western analysis of SKI

expression in representative C2C12 LNIT-huSKI clones cultured under induced condition

(-Dox) or suppressed condition (+Dox) for three days. A vector-only clone (LNITX) was used as a negative control (Ctrl). TFIIE-α was a loading control.

(C) A representative C2C12 LNIT-huSKI clone (F8) was cultured in GM or DM for 1 day with or without Dox and phase contrast microscopy shows myotube formation only

in the absence of Dox.

(D) Enhanced Myog expression by overexpression of SKI. LNIT-huSKI F8 clone or a vector-only clone (LNITX) was cultured in GM or DM for 1 day with or without Dox and Western blot analysis revealed the expression of Myog and SKI. TFIIE-α was a loading control. Note that endogenous Ski expression in the right panel was revealed only after much longer exposure than the left panel where overexpression of SKI was detected.

140 cultured in the presence of Dox to suppress SKI overexpression (Fig. 13C). Western

analysis of these cells revealed expression of Myog only in cells that expressed

exogenous SKI (Fig. 13D, left panel). These results were not due to Dox treatment since

the vector-only controls showed no obvious difference in Myog or endogenous Ski

expression in the presence or absence of Dox (Fig. 13D, right panel). These data indicate

the increased SKI expression stimulates the commitment of C2C12 myoblasts to terminal

muscle differentiation.

Generation of C2C12 clones with inducible knock-down of Ski expression

The gain-of-function study above indicated that, as in avian cells, overexpression of Ski induces muscle terminal differentiation of C2C12 cells. However, those results do not necessarily implicate endogenous Ski in this process. To address this issue we asked whether knocking down endogenous mouse Ski expression would affect the differentiation of C2C12 cells. To accomplish this and avoid potential deleterious effects on the long-term cell viability due to the loss of endogenous Ski, we used tet-regulated expression of shRNAs to knockdown Ski in C2C12 cells. The TMP-tTA vector used for this purpose was modified from the TMP retroviral vector (159-161) by substituting the tTA gene for the GFP gene, so that this single retrovirus carries both tTA and its response element (TRE) (Fig. 14A). DNA sequences which encoded shRNAs targeting mouse Ski were inserted within the framework of micoRNA-30 (miR30), downstream of the TRE-

CMV promoter (Fig. 14A). Transcripts of this cassette resembling natural microRNA-30 could be generated upon removal of Dox, resulting in the knockdown of Ski. Three different shRNA inserts targeting mouse Ski gene were designed and retroviral constructs, designated as TMP-tTA-mSki977, 1145 and 1819 were generated.

141

142 Fig. 14. Dox-regulated knock-down of Ski in C2C12 cells via shRNAs in the context of

miR30

(A) Schematic representation of TMP-tTA-mSki retroviral vector showing the 5’ LTR

and 3’ self inactivating LTR (SIN-LTR), puromycin resistance gene (Puror) driven by the

PGK promoter, IRES, tTA and a microRNA cassette downstream of TRE-CMV

promoter. The shRNAs targeting mouse Ski were inserted into the microRNA cassette

between the 5’ and 3’ flanking sequences derived from the miR30 primary transcript

(5’miR30 and 3’miR30).

(B) Dox-regulated knockdown of Ski in TMP-tTA-mSki1819 clones. Western analysis of

Ski expression was performed on representative C2C12 TMP-tTA-mSki1819 clones or

the vector-only clone (Ctrl) cultured with (+Dox) or without Dox (-Dox) for one week.

(C) Western blot analysis revealed decreased Ski expression in C2C12 TMP-tTA- mSki1145 clone in response to removal of Dox. Cells were cultured in the presence of

200ng/ml Dox for eight days, switched into Dox-free medium and analyzed after culture for indicated number of days.

(D) Western blot analysis revealed restoration of Ski expression in C2C12 TMP-tTA- mSki1145 clone in response to addition of Dox. Cells were cultured in the absence of

Dox for eight days to achieve the maximal knockdown of Ski and then switched into medium containing 200ng/ml Dox for indicated number of days.

(E) Western blot analysis revealed dose response of Ski knockdown to Dox in a C2C12

TMP-tTA-mSki1145 clone. Cells were cultured in the absence of Dox for 6 days and switched into medium containing the indicated concentration of Dox for another 6 days.

143 TFIIE-α was used as a loading control and a vector-only clone (Ctrl) shows the endogenous Ski level.

144 PA317 packaging cells were transfected with each of the three TMP-tTA-mSki

vectors (or the empty vector) and viruses were harvested to infect C2C12 myoblasts.

Puromycin-resistant clones were isolated and propagated in GM plus Dox to prevent shRNA expression. These clones were then subdivided and cultured with or without Dox for 6 days prior to testing for conditional knockdown of Ski. Western analysis of several

TMP-tTA-mSki1819 clones revealed a highly efficient Dox-dependent knockdown of Ski without significant leakiness (Fig. 14B). TMP-tTA-mSki cells grown in the absence of

Dox produced nearly undetectable Ski level while the same clones grown in the presence of Dox expressed Ski at similar levels to that of the vector-only control. Out of approximately 40 clones tested with each of the three TMP-tTA-mSki shRNAs, 70~80%

exhibited Dox-dependent knockdown of Ski, indicating the high efficiency of this system

(data not shown).

Ski knockdown is reversible and Dox-dose dependent

To determine the kinetics of Ski knockdown, a C2C12 TMP-tTA-mSki1145 clone

was propagated in Dox-containing medium, transferred to Dox-free medium and

monitored for Ski expression over an 8-day period. Western analysis showed that

knockdown of Ski was apparent within 4 days after Dox removal and was virtually

complete after 6 days (Fig. 14C). This knockdown was completely reversible; re-addition

of Dox to these cells restored normal Ski expression within 4 days (Fig. 14D). The extent

of Ski knockdown was also doxycycline dose-dependent; expression of Ski was barely

detectable at 2ng/ml of Dox or less and was comparable to the vector-only control at

20ng/ml of Dox or more (Fig. 14E). Taken together, these data demonstrated that this

145 single-vector system allows tightly regulated and reversible knockdown of endogenous

Ski.

Impaired myotube formation in the absence of Ski

Using this Dox-regulated knockdown system, we next evaluated the consequences of the loss of Ski on terminal differentiation. A C2C12 TMP-tTA-mSki1145 clone was either kept in GM plus Dox to maintain Ski expression or switched into GM minus Dox for 7 days to achieve the maximal knock-down of Ski. Subsequently, upon reaching 80% confluence, these cells were switched from GM to DM while continuing the maintenance or suppression of Ski expression. Phase microscopy of the cultures prior to switching to

DM revealed that the morphology of TMP-tTA-mSki and TMP-tTA cells in GM was similar and not affected by the loss of Ski expression (Figs. 15A-D). This was not true for cells that were switched into DM to induce differentiation. The TMP-tTA-mSki cells in which Ski expression was maintained exhibited extensive myotube formation (Figs. 15E and I). However, when Ski expression was knocked down in the TMP-tTA-mSki clone, no obvious myotube formation was visible within 2 days (Fig. 15F) and only a few small myotubes had formed after 4 days (Fig. 15J). The lack of myotube formation was not due to delayed differentiation kinetics, since in the absence of Dox, TMP-tTA-mSki cells did not show significant myotube formation even after 6 days in DM (data not shown).

Furthermore, the differentiation defect of TMP-tTA-mSki cells in the absence of Dox was not due to the withdrawal of Dox, because myotube formation of the TMP-tTA control was independent of Dox (Figs. 15G and H). Thus the results indicate that Ski expression is required for terminal muscle differentiation of C2C12 cells.

146

147 Fig. 15. Loss of Ski prevents terminal differentiation of C2C12 myoblasts.

Prior to the following assays, cells were cultured for 7 days in GM either with Dox as a control or without Dox to achieve maximal knock-down of Ski.

(A-J) A C2C12 TMP-tTA-mSki1145 clone was analyzed for the effect of Ski knockdown on myotube formation by phase contrast microscopy (left panel). Cells were cultured to

80% confluence in GM with (+Dox) or without Dox (-Dox), switched to DM with

continued presence or absence of Dox and cultured for 2 days and 4 days. A vector-only

clone (TMP-tTA) was cultured the same way and used as a control. Representative fields

are shown at 25× magnification.

(G, H) TMP-tTA-mSki1145 cells were cultured as described above and kept in DM for 3

days. Myotubes were labeled by indirect immunofluorescence with antibody against

MHC (green) and nuclei were counterstained with DAPI (blue). Representative fields are

shown at 100× magnification.

(I, J) Quantitative analysis of the immunofluorescence assays in G and H. Red bars

represent culture in the presence of Dox (+Dox) and blue bars represent culture in the

absence of Dox (-Dox). The percentage of DAPI-stained nuclei in MHC-positive cells (I)

and the average number of nuclei per MHC-positive myotube (J) were calculated as

described in Materials and Methods. Data above the bar represent means of three

independent experiments. Error bars show standard deviations of means.

148 Decreased myotube formation in the absence of Ski could be due to a block at any

stage in the differentiation pathway or a failure of fully differentiated myocytes to fuse.

To distinguish between these possibilities, we performed immunofluorescence for MHC which is a terminal differentiation marker expressed in both unfused myocytes and multinucleated myotubes. Consistent with the morphological observations above, TMP- tTA-mSki cells exhibited extensive formation of MHC-positive multinucleated myotubes when expressing Ski, while the vast majority of cells from the same clonal line were

MHC-negative when Ski expression was knocked down (Figs. 15K and L). The effects of loss of Ski on terminal differentiation and fusion were further quantified by measuring both the percentage of nuclei in MHC-positive cells (differentiation index) and the average number of nuclei per MHC-positive myotube (fusion index). We found that the loss of Ski caused significant (P<0.05) drops in both the differentiation index (from

46.4±6.2% to 14.6±9.8%) and fusion index (20.7±7.7 to 3.4±1.7 nuclei per myotube)

(Figs. 15M and N), respectively. The results indicate that in the absence of Ski, both differentiation (~3 fold fewer MHC-positive cells) and myotube formation (~ 6 fold fewer nuclei/myotube) were severely impaired (Figs. 15M and N). Thus the loss of Ski resulted in a differentiation block upstream of myocyte fusion.

Loss of Ski inhibits commitment to terminal muscle differentiation

To determine the stage at which terminal differentiation was blocked by the absence of Ski, we further examined the expression of an early differentiation marker,

Myog, and a late differentiation marker, MHC. As shown in Figure 16A, when maintained in GM, neither the C2C12 TMP-tTA-mSki1145 cells nor the TMP-tTA cells express Myog or MHC at detectable levels regardless of Dox treatment. However, after 3

149

150 Fig. 16. Loss of Ski blocks myogenic differentiation at early stage.

(A, B) A representative C2C12 TMP-tTA-mSki1145 clone (A), TMP-tTA-mSki1819 clone (B) or TMP-tTA-mSki977 clone (B) was cultured as described in the legend to Fig.

15 and kept in DM for 3 days prior to Western analysis for Ski, MHC and Myog expression. TFIIE-α was used as a loading control.

(C) Western blotting revealed expression of Ski, MHC and Myog in a C2C12 TMP-tTA- mSki1819 clone cultured in DM for 3 days with the indicated concentrations of Dox

(0~200ng/ml).

151 days of culture in DM, TMP-tTA-mSki cells in which Ski expression was maintained by

Dox treatment expressed both Myog and MHC at high levels. This was also the case when TMP-tTA cells were cultured in DM regardless of Dox treatment. In contrast, when the same TMP-tTA-mSki clone with Ski expression knocked down were cultured in DM,

Myog and MHC were expressed at very low levels (Fig. 16A). Similar results were

obtained using TMP-tTA-mSki clones expressing the other two shRNAs, indicating that

the differentiation block was not due to an off-target effect of the shRNA (Fig. 16B). In

addition, a Dox dose-response experiment with a C2C12 TMP-tTA-mSki clone

expressing a different shRNA (mSki1189) revealed that the expression of Myog and

MHC dropped in parallel with the decrease in endogenous Ski expression (Fig. 16C). The

loss of Myog expression indicated that differentiation was blocked at a very early stage in

the absence of Ski expression.

The early stages of muscle differentiation are marked by a well-characterized

progression of protein expression including the myogenic regulatory factors (MyoD and

Myog) and the cell cycle regulator (p21) (91, 92, 128, 129). To obtain a quantitative

assessment of the early disruption of the myogenic differentiation program due to the loss of Ski expression, we investigated the percentage of cells expressing these proteins during the differentiation of C2C12 TMP-tTA-mSki clones in the absence or presence of

Dox. Myog and p21 are markers of commitment to terminal differentiation and

withdrawal from the cell cycle, respectively. Their increased expression can be detected

in both differentiating myocytes and fully differentiated multinucleated myotubes and is

widely used to assess the early stage of differentiation. Within 3 days of switching to

DM, 35.5% of TMP-tTA-mSki cells were Myog-positive when cultured in the presence

152 of Dox (Figs. 17A and G) and 27.9% were p21-positive (Figs. 17C and H). In sharp

contrast, in the same TMP-tTA-mSki clone with Ski expression knocked down, the percentages of Myog and p21-positive cells decreased to 8.8% (Figs. 17B and G) and

14.9% (Figs. 17D and H), respectively. On the other hand, the expression of MyoD, a constitutive myogenic lineage marker of C2C12 cells, was comparable in the presence and absence of Ski expression (Figs. 17E, F and I). These results indicate that the loss of

Ski blocked a step in the differentiation pathway downstream of MyoD. Results obtained with a representative TMP-tTA-mSki1145 clone were shown in Figure 17 and similar results were obtained with two other clonal lines (data not shown).

To determine whether the observed Ski-dependent changes in protein expression were due to reductions in the expression of their mRNAs, we performed quantitative realtime-RT-PCR on RNA isolated from TMP-tTA-mSki cells cultured for three days in

DM plus or minus Dox. Concomitant with the 5.4-fold lower expression of Ski mRNA in

TMP-tTA-mSki cells in the absence of Dox, Myog and p21 mRNA levels were reduced by 12 fold and 5.5 fold, respectively. On the other hand, the level of MyoD mRNA was not significantly affected by the loss of Ski (Fig. 17J). These results mirror those obtained in the analyses of protein expression and indicate that Ski is necessary for the transcription or accumulation of mRNAs which are important for initiating muscle terminal differentiation.

Ski occupies Myog regulatory region in differentiating myoblasts

Because Ski is known to be a co-regulator of transcription it seemed likely that the changes we detected in the expression of muscle-specific mRNA might be due to direct effects of Ski on transcription. This possibility was especially appealing for Myog

153

154 Fig. 17. Loss of Ski reduces the expression of muscle-specific genes at both mRNA and

protein levels.

(A-F) Indirect immunofluorescence was performed on a C2C12 TMP-tTA-mSki1145

clone cultured in DM with or without Dox for 3 days. Differentiating cells were detected

with antibodies against Myog, p21 or MyoD (green) and total nuclei were counterstained

with DAPI (blue). Representative fields are shown at 400× magnification.

(G-I) Quantitative analysis of the immunofluorescence assays in C-H. Red bars represent

culture in the presence of Dox (+Dox) and blue bars represent culture in the absence of

Dox (-Dox). The fraction of nuclei positively-stained for Myog (G), p21 (H) and MyoD

(I) were calculated as described in Materials and Methods. Data above the bar represent

means of three independent experiments. Error bars show standard deviations of means.

(J) Quantitative realtime-PCR analysis of Ski, Myog, p21 and MyoD mRNA level in a

C2C12 TMP-tTA-mSki1145 clone cultured in DM with or without Dox for 3 days. Fold

change of the transcript level in culture in the presence of Dox over that in the absence of

Dox was calculated as described in Materials and Methods. Data above the bars represent

means from three independent experiments performed in triplicate. Error bars show

standard deviations of means.

155 not only because its expression is required for the initiation of terminal differentiation

(91, 92, 128, 129) but also because published reporter gene assays have demonstrated that transient overexpression of Ski can transcriptionally activate the Myog regulatory region

(115, 116).

To investigate whether endogenous Ski might directly regulate Myog

transcription, we performed ChIP assays on C2C12 cells cultured in either GM or DM.

DNA was isolated from the chromatin immunoprecipitated with anti-Ski antibodies and

was analyzed by PCR using primers that amplify the Myog regulatory region including

the E1 box, a MEF2 site and a MEF3 element (119-121, 123, 150) (Fig. 18A). The ChIP

assays revealed that Ski was not bound to this endogenous Myog regulatory region in

proliferating C2C12 cells cultured in GM. However, Ski became associated with this

region when the cells were stimulated to differentiate by switching them into DM (Fig.

18B, top panel). The specificity of this interaction was verified by negative results in

control ChIP assays using either a non-specific antibody (normal IgG) and the same

primers amplifying the Myog regulatory region or the Ski antibody and primers

amplifying a non-promoter region downstream of Myog gene (Fig. 18B, top and middle

panels). Furthermore, the observation that Ski was expressed at similar levels in cells

cultured in GM and DM indicated that the increased interaction between Ski and the

endogenous Myog regulatory region upon differentiation was not due to a parallel

increase in Ski expression (Fig. 18C). Likewise, the fact that the Smad7 regulatory

region, which is known to be bound by Ski (196), was occupied by Ski in both

proliferating and differentiating cells (Fig. 18B, bottom panel) indicated that Ski’s

chromatin binding ability didn’t depend merely on the change from GM to DM. Thus Ski

156

157 Fig. 18. Ski binds the endogenous Myog regulatory region upon differentiation.

(A) Schematic representation of the Myog regulatory region. E-boxes (E1 and E2), the

MEF2 binding site and MEF3 binding site are indicated as open boxes relative to the transcriptional start site (+1). Primer sets used to amplify sequences spanning the Myog regulatory region or a non-promoter region downstream of the Myog gene is represented as arrows.

(B) ChIP analysis of Ski’s occupancy of the endogenous Myog regulatory region in proliferating and differentiating cells. Equivalent amounts of cross-linked chromatin from

C2C12 cells cultured in GM or in DM for 2 days were precipitated with an antibody against Ski (α-Ski) or a normal rabbit IgG. The DNA isolated from precipitated chromatin was analyzed by PCR using primers spanning Myog regulatory region (upper), non-promoter region (middle) or Smad7 regulatory region (lower) and following electrophoresis, PCR products were visualized by ethidium bromide staining.

(C) Western analysis of Ski, MHC and Myog expression in C2C12 cells cultured in GM or in DM for the indicated periods of time. TFIIE-α was used as a loading control.

158 selectively binds the Myog regulatory region in concert with a signal that initiates

differentiation. These results suggest that the requirement for endogenous Ski in the

initiation of terminal muscle differentiation might be due to its direct role in activating of

Myog transcription.

MEF3 binding site is required for the activation of Myog regulatory region by Ski

In light of the above results we sought to define the Ski-response cis element in

the Myog regulatory region. A 184 bp Myog regulatory region sequence has been defined

as a muscle-specific regulatory region and contains a number of transcriptional factor

binding sites which are critical for activation of Myog transcription during differentiation

(117, 119-121, 123, 150). Among others, they include two consensus MyoD binding E-

boxes (proximal E1 box and distal E2 box), a MEF2 binding site and a MEF3 site (Fig.

19A, upper panel). Ski does not bind DNA directly, but it has previously been shown to

synergize with MyoD and MEF2 in activating transcription of Myog reporters containing

this 184 bp sequence (115). This synergy requires the binding of these transcription

factors to their response elements in the Myog regulatory region. To address whether Ski

activation of Myog regulatory region was mediated only through these binding sites, we

mutated the MEF2 binding site or both E boxes (E1 and E2) to eliminate binding by their

corresponding transcriptional factors (Fig. 19A). In addition, we mutated the MEF3 site

although it had not previously been implicated in activation by Ski. C2C12 myoblasts were transfected with the wild-type Myog reporter (Myog184) or its mutated derivatives

(Myog184-MEF3m, Myog184-MEF2m, Myog184-E1E2m, see Fig. 19A) along with a Ski

expression vector or an empty vector and luciferase activity was measured 2 days after switching the cells to DM. Ski expression potentiated the activity of the wild-type Myog

159

160 Fig. 19. The MEF3 binding site is required for maximal activation of Myog promoter by

Ski.

(A) Schematic representation of the 184bp proximal Myog regulatory region and its

mutants. The intact E1 and E2 boxes, MEF2 site and MEF3 site are indicated as open

boxes, whereas the mutated elements are marked by “X”. The sequences of wild type

binding sites are indicated in uppercase and the mutated bases in these binding sites are

indicated in lowercase.

(B, C) C2C12 cells were co-transfected with wild-type or mutated reporters described

above along with a Ski expression vector or an empty vector (mock) and cultured in DM

for 2 days. Luciferase activity (RLU) of each sample was calculated as described in

Materials and Methods.

(B) Activation of wild-type Myog regulatory region (Myog184) by Ski was calculated as

the fold change of RLU in Ski-transfected cells over that in the empty vector-transfected

cells which was set to an arbitrary unit of 1. Data are expressed as the mean values from

three independent experiments performed in triplicate. Error bars represent the standard

deviations of the means.

(C) The activation of wild-type and mutated Myog regulatory region by Ski were

calculated as described in B. The activation of the mutated Myog regulatory regions

(Myog184-E1E2m, Myog184-MEF2m, Myog184-MEF3m) by Ski is expressed as the

percentage of that with the wild-type which was set at 100%. Data are expressed as the

mean data of each reporter from three independent experiments performed in triplicates.

Error bars represent the standard deviations of the means.

161 reporter (Fig. 19B) and surprisingly activated the transcription of the reporters carrying

mutations of either the MEF2 site or both E boxes to approximately 80% of the level

observed with the wild-type reporter (Fig. 19C). These results suggested that activation of

Myog regulatory region by Ski was not mediated exclusively by these two regulatory elements. However, activation of the reporter bearing a mutated MEF3 site by Ski was only 50% of that of the wild type (Fig. 19C), indicating that this site may be the major

Ski-responsive cis element within the Myog regulatory region.

Ski associates with Eya3 and Six1 in differentiating muscle cells

We have shown that activation of Myog transcription by Ski is mediated mainly

through the MEF3 site in the regulatory region. Since Ski does not bind DNA directly, it

seemed likely that its association with the endogenous Myog regulatory region and its

activation of Myog transcription are mediated by its association with transcription factors

that bind to MEF3 element. Six1 has been shown to bind to the MEF3 site of the Myog

regulatory region (123) and to synergize with Eya to positively regulate the transcription

driven by this cis element (58). In addition, because Dach, a Ski family member, forms a

trimeric complex with Six1/Eya3 to regulate muscle-specific gene expression (58), it

seemed possible that Ski may be tethered to Myog regulatory region via an association with Six1/Eya3 in a similar manner. We therefore investigated whether Ski interacts with

Six1/Eya3 in muscle cells undergoing terminal differentiation. C2C12 cells were co- transfected with Ski and Flag-tagged Six1 and Eya3 expression vectors and cells were

cultured either in GM or induced to differentiate in DM for 48hr. Extracts of these cells

were immunoprecipitated with either rabbit anti-Ski or normal rabbit IgG and analyzed

for co-precipitation of Six1/Eya3 by Western blotting with anti-Flag. As seen in Figure

162 20A, neither Six1 nor Eya3 was precipitated by normal rabbit IgG whereas Six1 was co-

precipitated with Ski at comparable levels in proliferating (GM) and differentiating (DM)

cells. In contrast, co-precipitation of Eya3 with Ski was barely above background in

proliferating cells, but clearly detectable in differentiating cells. Thus in muscle cells Ski is constitutively associated with Six1 but interacts with Eya3 only upon differentiation.

Because a previous study suggested Ski activated transcription of Myog regulatory region in co-operation with MyoD and MEF2 (115), we performed similar co- precipitation assays to examine the possible interactions of Ski with these proteins.

Surprisingly, neither MyoD nor MEF2C co-precipitated with Ski in either proliferative or differentiating C2C12 cells (Figs. 20B and C). Considering that MyoD and MEF2 activate the transcription of some muscle-specific genes in a cooperative manner (129), it seemed possible that their interactions with Ski might require the presence of both proteins. To test this possibility, similar co-immunoprecipitation experiments were performed with extracts of C2C12 cells that were co-transfected with Ski and both MyoD and MEF2 expression vectors. Once again both proteins were expressed at high levels but neither of them co-precipitated with Ski in C2C12 cells cultured in GM or DM (Fig.

20D). Although negative results are inconclusive, these observations and the results of the reporter assays suggest that previously described transcriptional co-operation of Ski with MyoD and MEF2 may be actually mediated by an indirect interaction of

MyoD/MEF2 complex with a DNA-bound Ski/Six1/Eya3 trimeric.

The DHD of Ski is required for its association with Six1 and its activation of Myog transcription

163

164 Fig. 20. Ski interacts with Six1 and Eya3 but not with MyoD or MEF2.

(A-D) C2C12 cells were co-transfected with expression plasmids for Ski and Flag-tagged

Six1 and Eya3 (A), MEF2-Flag (B), MyoD (C) or MEF2-Flag and MyoD (D) and cultured in GM or DM for 2 days. Lysates were immunoprecipitated with an rabbit antibody against Ski (α-Ski lanes) and precipitates were analyzed by Western blotting using anti-Ski, anti-Flag or anti-MyoD antibodies. 10% of the input for immunoprecipitation (Input lanes) confirms that comparable amounts of these proteins were used in the immunoprecipitation assays. Immunoprecipitation with a normal rabbit

IgG (IgG lanes) was used as a negative control.

165 It has been shown that Dach interacts with Six1 through its DHD domain (55, 56,

58). It was therefore of interest to determine whether this conserved domain in Ski also mediates its interaction with Six1. The ability of a Ski mutant lacking the DHD to interact with Six1 was examined by coimmunoprecipitation (Figs. 21A and B). Deletion of the

DHD from Ski (SkiΔDHD) greatly reduced its ability to interact with Six1 although it did

not affect interaction with Eya3.

Having identified the DHD as the Six1 binding domain in Ski, we sought to

determine the effect of its deletion on transcriptional activation of Myog. Cotransfection

experiments showed that SkiΔDHD failed to activate transcription of the wild-type Myog

reporter (Myog184) significantly compared to the wild-type Ski (Fig. 21C). Western blots

of the transfected cells demonstrated that the lack of reporter activation by Ski∆DHD was

not due to a failure to express this protein at similar level to the wild-type Ski (Fig. 21D).

To confirm these results, we next asked whether the DHD of Ski was also required for

activation of endogenous Myog expression upon differentiation. To answer this question

we assessed whether re-introducing Ski or SkiΔDHD into C2C12 TMP-tTA-mSki cells

could overcome the loss of Myog and MHC expression due to the knock-down of Ski in

these cells. C2C12 TMP-tTA-mSki cells were grown in GM minus Dox to achieve the

maximal knock-down of Ski, transfected with Ski expression vectors and analyzed for

Myog and MHC expression after 2 days of culture in DM minus Dox. We found that

wild-type Ski restored Myog and MHC expression. However, SkiΔDHD failed to do so

and this inability was not attributable to poor expression of this mutant (Fig. 21E),

indicating that the DHD is needed for the activation of Myog transcription.

166

167 Fig. 21. The DHD of Ski is required for its association with Six1 and its activation of

Myog transcription.

(A) Schematic diagrams represent Ski and the DHD deletion mutant (Ski∆DHD).

(B) Interaction of Ski with Six1 was mediated by its DHD. C2C12 cells were co- transfected with expression plasmids for Flag-tagged Six1 and Eya3 and full-length Ski or Ski∆DHD (described in A) and cultured in DM for 2 days. Immunoprecipitation assays were performed as described in the legend to Fig. 20.

(C) C2C12 cells were co-transfected with the wild-type Myog reporter (Myog184) and expression vectors for wild-type Ski or Ski∆DHD and cultured in DM for 2 days.

Luciferase activity of each sample was calculated as described in Materials and Methods and the fold activation of Myog reporter by wild-type Ski and Ski∆DHD were calculated as described in the legend to Fig. 19B. Data are expressed as the mean values from three independent experiments performed in triplicate. Error bars represent the standard deviations of the means.

(D) Western blotting of the same lysates used for the luciferase assays in C revealed similar expression of wild-type Ski and Ski∆DHD. TFIIE-α was used as a loading control.

(E) Cells were cultured in GM in the absence of Dox for six days to achieve maximal knock-down of Ski and then transfected with expression vectors for wild-type Ski or

Ski∆DHD. Cells were then switched to DM and cultured for 2 day prior to harvest.

Western blotting revealed expression of Ski, MHC and Myog. TFIIE-α was used as a loading control.

168 Discussion

The role of Ski in terminal differentiation has been evidenced by its ability to induce myogenesis in non-muscle cells in vitro and hypertrophy of type II fast muscle of adult mice when it was overexpressed (25, 42, 45, 149). It remained of great interest to

determine whether Ski also regulates terminal differentiation of muscle satellite cells

which are the committed myogenic cells and responsible for the skeletal muscle

regeneration (197, 198). However, since Ski-/- mice die at birth, we were not able to address this issue using this mouse genetic model (47). We have therefore used a well- established in vitro model, satellite cell-derived C2C12 myoblasts, to investigate the role of Ski in terminal differentiation and its underlying molecular mechanism. Our findings that differentiation was enhanced by overexpression of Ski and severely impaired in the absence of endogenous Ski indicate that Ski not only induces but also is necessary for differentiation of determined myoblasts. We further established a tight linkage between the myogenic activity of Ski and the expression of Myog at both mRNA and protein levels. Additional results suggest that direct regulation of Myog underlies Ski’s myogenic activity by demonstrating transcriptional activation of Myog by Ski in reporter assays and the association of Ski with the endogenous Myog regulatory region upon differentiation.

Surprisingly, transcriptional activation of Myog by Ski was largely mediated by its cooperation with Six1/Eya3 through a MEF3 binding site, not with MyoD or MEF2 through the E boxes or the MEF2 binding site. The necessity of an evolutionary conserved DHD for the interaction of Ski with Six1 is consistent with published data on the interaction between Dach and Six proteins (56). Our findings support a model in

169 which Ski substitutes for Dach in the standard Dach/Six1/Eya3 complex to regulate

myogenesis.

Our earlier attempts to perform these studies using C2C12 myoblasts were

frustrated by our inability to maintain cells in which Ski is constitutively overexpressed

or knocked down. Here we surmounted this problem using inducible vectors, which

allowed cloning and propagation of transduced cells expressing endogenous Ski at

normal levels prior to acute induction or knockdown of Ski expression. The retroviral

vector we constructed for regulated knockdown (TMP-tTA) was modified from the TMP

retroviral vector of Lowe and coworkers (159-161). Since the TMP vector only carries

the TRE-driven microRNA cassette while a second vector provides the tTA gene, two rounds of infection and antibiotic selection are required. This approach works well for

cells whose activity is not compromised by prolonged passage, but it is not optimal for

C2C12 myoblasts which gradually lose their differentiation potential. Our new vector,

TMP-tTA, carries the TRE-driven microRNA cassette and the tTA gene in a single

retroviral vector, so only one round of infection/antibiotic selection is needed. This vector

has been proven as effective as the original TMP vector with regard to the efficiency,

reversibility, dose dependence and insignificant leakiness of knockdown. Given its

simplicity and effectiveness, this knockdown vector is suitable to study any gene required

for cell survival and is especially useful in cells whose activities of interest are sensitive

to prolonged cell culture passage.

Earlier work showed that Ski stimulated myogenesis in non-muscle primary avian

cells by inducing both MyoD and Myog, two genes controlling myogenic lineage

determination and terminal differentiation, respectively (25). We therefore assumed that

170 loss of Ski would lead to downregulation of both of these genes and result not only in

impaired differentiation but also in loss of myogenic identity. However, in the present

report, we observed that loss of Ski only affected the expression level of Myog but not

MyoD in determined C2C12 myoblasts, indicating that Ski is not necessary for

maintaining myogenic identity. Given the pivotal role of Myog in the initiation of

differentiation (91, 92, 128, 129), it is likely that regulation of Myog expression is the key

mediator of Ski’s effect on terminal differentiation. Furthermore, our data revealed that

the regulation of Myog expression by Ski was not only at protein level, but also at

transcript level. This result along with the observed transactivation of Myog reporter by

Ski and the occupancy of Ski on the endogenous Myog regulatory region places Myog as

a direct transcriptional target of Ski.

As a non-DNA binding transcription factor, Ski has to be brought into contact

with promoters/enhancers by interaction with transcription factors that bind to specific

DNA cis-regulatory elements (31). Cis-elements essential for transcriptional regulation of

Myog include two E boxes and a MEF2 site which are bound by MyoD and MEF2, respectively (119-121, 123, 150). Previous studies have shown that MyoD and MEF2 cooperatively activate the Myog transcription through binding to these elements at the onset of myogenesis (129). This cooperative interaction has also been seen in the upstream regulatory sequences of other muscle-specific genes, including MLC1/3 and

MCK (114, 199-201). Interestingly, Ski has been implicated in activation of these promoters in co-operation with MyoD and a previous report based solely on reporter assays suggested the same mechanism for the transactivation of Myog regulatory region by Ski (115). However, our results, using the same Myog regulatory region, appear to

171 conflict with this report by showing that the destruction of E boxes or the MEF2 site had

only a marginal effect on the ability of Ski to activate the transcription of Myog reporters.

We cannot explain this discrepancy but we believe that although reporter assays can be

useful for defining cis-response elements, their biological significance requires

confirmation by additional experiments. Our inability to detect direct interactions of Ski

with MyoD or MEF2 and the absence of any previous data showing these interactions suggests an indirect mechanism for the co-operation between Ski and MyoD on Myog

activation.

Recent studies demonstrating the presence of functional Pbx1 and MEF3 binding

sites have revealed that the Myog regulatory region is more complex than previously believed (58, 123, 150, 202, 203). Our data underscore this complexity by showing that

the MEF3 site instead of the previously implicated E boxes and MEF2 site mediated

transactivation of Myog by Ski. These sites reside in close proximity within the Myog

regulatory region suggesting that their combined occupancy by Six1, MyoD and MEF2

may be the basis for the cooperation between Ski and these proteins. Our finding that

activation of Myog transcription by Ski requires interaction with the MEF3-binding Six1

protein provides the mechanism for recruitment of Ski to the Myog regulatory region.

This interaction and the observation that Ski-/- mice exhibited a similar muscle defect as

the Six1-/- mice, suggests a common mechanism by which both Ski and Six1 regulate

myogenesis (47, 48, 123, 134, 136, 204).

The differentiation-dependent interaction between Ski, Six1 and Eya3 correlates

well with the observation that the association of Ski with the endogenous Myog

regulatory region occurs only in differentiating cells. Since growth factor signaling can

172 block the ability of Eya to interact with Six proteins (59, 138), it is possible that the

withdrawal of serum growth factors initiates terminal differentiation by freeing Eya to

form the Ski/Six/Eya trimeric complex on the Myog regulatory region to activate its

transcription. Dachshund, has also been reported to transactivate the Myog reporter

through the interaction with Six and Eya (55, 58). However, since activation of Myog

expression and subsequent differentiation was almost abolished in the absence of Ski, we

believe that in C2C12 myoblasts it is Ski, not Dach, that is the essential member of this

trimeric complex. Our findings shed new light on the mechanism of Ski’s myogenic

activity and the importance of the Ski/Six/Eya trimeric complex in muscle terminal

differentiation.

Future Directions

Mechanism of Myog transcriptional regulation by Ski/Six1/Eya3 complex

We have shown that Ski regulates the Myog promoter through a MEF3 site and is

also associated with Six1/Eya3. However, in order to prove that the regulation of Myog

by Ski is due to its ability to interact with Six1 and Eya3 proteins, we will use the same

shRNA-based strategy we employed with Ski to knockdown Six1 or Eya3 in C2C12

cells. We will examine the effect of knockdown of Six1 and Eya3 expression on the transactivation of Myog promoter, induction of Myog expression and occupancy of

endogenous Myog promoter by Ski in these cells by reporter assays, western blotting and

ChIP assays, respectively.

Our data also indicate that although Ski interacts with Six1 in both proliferative

and differentiating cells, the interaction between Ski and Eya is differentiation-dependent.

173 These results correlated with the differentiation-induced association of Ski with

endogenous Myog promoter. The remaining question here is how serum removal triggers

these protein-protein and protein-DNA associations. Because little is known about how

Eya3 is involved in myogenic differentiation, as a starting point, we will investigate the

intracellular location of Eya3 and posttranslational modifications of Eya3, especially its

phosphorylation under proliferation and differentiation conditions. Based on the

knowledge gained from the above experiments, we will further investigate whether the

translocation or change of phosphorylation status of Eya3 causes the change in its ability

to interact with Ski and Six1 and to associate with Myog promoter, therefore altering the

formation of Ski/Six1/Eya3 complex on the endogenous Myog promoter.

Regulatory cascades of Ski during myogenic differentiation

Our results indicate that Myog is a target gene of Ski and plays a major role in the

regulation of myogenic differentiation by Ski. However, our understanding of how Ski

regulates this process is still fragmentary. To unravel the transcriptional activation

cascade of Ski, microarray-based gene expression profiles of C2C12 cells in which Ski is

overexpressed or knocked down under proliferation or differentiation conditions will be generated and computationally analyzed. ChIP-on-chip assays will also be used to identify transcriptional factor-binding sites for Ski in proliferating and differentiating

C2C12 cells in a genome-wide manner. Combining these studies will allow us to uncover unexplored aspects of the myogenic regulation by Ski, reveal the transcriptional regulation cascades downstream of Ski and suggest some new directions for future studies.

The regulation of satellite cell behavior by Ski in vivo

174 Our studies suggest that Ski is required for the differentiation of satellite cell- derived C2C12 cells in vitro, but since constitutive Ski knockout mice die at birth, we could not use these animals to study the effect of Ski on the behavior of adult satellite cells. In the future, taking advantage of the Pax7 promoter which is active early in myogenic progenitors and later in satellite cells (108-110, 179), mice with Ski ablated in satellite cells will be generated by crossing mice with a floxed Ski allele to animals that transgenically express Cre recombinase under the control of the Pax7 promoter (Pax7-

Cre). These animals will enable us to investigate the role of Ski in satellite cells in vivo.

175 DISCUSSION

The Contradictions in This Study

In this study, the apparent requirement of Ski in myogenic differentiation of

C2C12 myoblasts appears to contradict our finding that both myogenic cells in Ski knockout mouse embryos and primary culture of satellite cells isolated from Ski knockout mice didn’t exhibit differentiation defect. However, this is not the first time people see such a striking contrast between in vitro myoblast culture and in vivo mice studies. For example, the distinct effects of Ski on differentiation in cell culture systems and in mice are reminiscent of . A number of studies have demonstrated that in C2C12 myoblasts, exogenous wild-type p53 stimulates myogenic differentiation, whereas interference with the function of endogenous p53 by a dominant negative p53 mutant inhibited this process

(205-208). However, skeletal muscles of p53-/- mice develop normally (208, 209). In addition, after whole muscle transplantation or crush injury, the regeneration ability of muscles was comparable between p53-/- mice and wild-type control, suggesting unaffected differentiation of satellite cells in the absence of p53 (210).

Unlike the intact animal model, C2C12 cell culture is a relatively isolated experimental system. Because of this, one of the possibilities to reconcile the apparent conflict between the in vivo myogenic cells and in vitro C2C12 culture is the flexibility or redundancy built into myogenic processes that are controlled by the microenviroment in vivo. For example, the normal muscle development and regeneration in p53-/- mice could be due to the compensatory effects of p63 and/or . Another well-known example is the compensatory mechanism between MyoD and Myf5. MyoD and Myf5 are functionally redundant for skeletal myogenesis. Both MyoD-/- and Myf5-/- mutant mice

176 have apparently normal skeletal muscle (96, 97). Only in the absence of both MyoD and

Myf5, mice exhibit a complete absence of skeletal muscle (98). Similar to these major

myogenic determination factors, although Ski probably participates in the major pathway

that regulates myogenic differentiation, other pathways may exist to substitute for Ski in

this complicated process during embryogenesis, although we don’t know what they are

yet. In the absence of Ski, these alternative pathways are activated in myogenic cells to

compensate for Ski’s function, probably by signals from the surrounding tissues. In

C2C12 myoblasts, because of the lack of these external signals, the loss of Ski is

catastrophic.

Different effects of Ski on differentiation were also observed between primary

satellite cells culture and C2C12 myoblasts. This reminds us of Myog-/- mice(99).

Although Myog is well known to be necessary for myogenic differentiation, satellite cells

isolated from Myog-/- mice differentiate normally in culture. This contradiction also

could be explained by the compensatory mechanism. In vivo, constitutive loss of Ski or

Myog may initiate certain compensatory mechanisms to maintain the differentiation potential of these satellite cells, therefore even after being isolated and cultured in vitro, they still be able to undergo differentiation. However, acute knockdown of Ski in C2C12 cells after Dox removal may not be able to give cells enough time or may require certain extrinsic signals to activate these compensatory mechanisms.

In spite of the conflict between them, both the in vitro and in vivo studies provide insights into myogenesis. The study with the in vivo mice model is at the whole animal level and includes microenvironmental influences, whereas cultured C2C12 cells provide a more controlled cellular environment and are available to measure cell autonomous

177 effects. In such an in vitro system, the effect of selectively altered gene expression on differentiation can be isolated from the potential compensatory effects often seen in animal models and can be evaluated with much less concern about possible indirect effects. We believe that by combining the advantages of both experimental systems, we can gain greater understanding of the overall biological significance of Ski and how it is involved in the complicated process of myogenesis.

178 REFERENCES

1. E. Stavnezer, D. S. Gerhard, R. C. Binari, I. Balazs, J Virol 39, 920 (Sep, 1981).

2. E. Stavnezer, A. E. Barkas, L. A. Brennan, D. Brodeur, Y. Li, J Virol 57, 1073

(Mar, 1986).

3. Y. Li, C. M. Turck, J. K. Teumer, E. Stavnezer, J Virol 57, 1065 (Mar, 1986).

4. X. Liu, Y. Sun, R. A. Weinberg, H. F. Lodish, Cytokine Growth Factor Rev 12, 1

(Mar, 2001).

5. C. Colmenares, E. Stavnezer, Semin Cancer Biol 1, 383 (Dec, 1990).

6. G. Zheng, J. Teumer, C. Colmenares, C. Richmond, E. Stavnezer, Oncogene 15,

459 (Jul 24, 1997).

7. J. J. Wilson, M. Malakhova, R. Zhang, A. Joachimiak, R. S. Hegde, Structure 12,

785 (May, 2004).

8. D. Chen et al., Cancer Res 63, 6626 (Oct 15, 2003).

9. S. Pearson-White, Nucleic Acids Res 21, 4632 (Sep 25, 1993).

10. N. Nomura et al., Nucleic Acids Res 17, 5489 (Jul 25, 1989).

11. K. L. Hammond, I. M. Hanson, A. G. Brown, L. A. Lettice, R. E. Hill, Mech Dev

74, 121 (Jun, 1998).

12. E. Mizuhara, T. Nakatani, Y. Minaki, Y. Sakamoto, Y. Ono, J Biol Chem 280,

3645 (Feb 4, 2005).

13. R. J. Davis, W. Shen, T. A. Heanue, G. Mardon, Dev Genes Evol 209, 526 (Sep,

1999).

14. S. Arndt, I. Poser, T. Schubert, M. Moser, A. K. Bosserhoff, Lab Invest 85, 1330

(Nov, 2005).

179 15. S. Arndt, I. Poser, M. Moser, A. K. Bosserhoff, Mol Cell Neurosci 34, 603 (Apr,

2007).

16. T. Nomura et al., Genes Dev 13, 412 (Feb 15, 1999).

17. F. Tokitou et al., J Biol Chem 274, 4485 (Feb 19, 1999).

18. R. Dahl, M. Kieslinger, H. Beug, M. J. Hayman, Proc Natl Acad Sci U S A 95,

11187 (Sep 15, 1998).

19. R. Dahl, B. Wani, M. J. Hayman, Oncogene 16, 1579 (Mar 26, 1998).

20. T. Prathapam, C. Kuhne, M. Hayman, L. Banks, Nucleic Acids Res 29, 3469 (Sep

1, 2001).

21. S. Akiyoshi et al., J Biol Chem 274, 35269 (Dec 3, 1999).

22. K. Luo et al., Genes Dev 13, 2196 (Sep 1, 1999).

23. Y. Sun et al., Mol Cell 4, 499 (Oct, 1999).

24. W. Xu et al., Proc Natl Acad Sci U S A 97, 5924 (May 23, 2000).

25. C. Colmenares, J. K. Teumer, E. Stavnezer, Mol Cell Biol 11, 1167 (Feb, 1991).

26. R. Nicol, G. Zheng, P. Sutrave, D. N. Foster, E. Stavnezer, Differ 10,

243 (1999).

27. K. Luo, Curr Opin Genet Dev 14, 65 (Feb, 2004).

28. M. Takeda et al., Mol Biol Cell 15, 963 (Mar, 2004).

29. K. Kokura et al., J Biol Chem 276, 34115 (Sep 7, 2001).

30. H. C. Heyman, E. Stavnezer, J Biol Chem 269, 26996 (Oct 28, 1994).

31. T. Nagase et al., Nucleic Acids Res 18, 337 (Jan 25, 1990).

32. K. A. Waite, C. Eng, Nat Rev Genet 4, 763 (Oct, 2003).

33. J. Massague, Annu Rev Biochem 67, 753 (1998).

180 34. S. Atanasoski et al., Neuron 43, 499 (Aug 19, 2004).

35. W. Wang, F. V. Mariani, R. M. Harland, K. Luo, Proc Natl Acad Sci U S A 97,

14394 (Dec 19, 2000).

36. C. D. Kaufman, G. Martinez-Rodriguez, P. B. Hackett, Jr., Mech Dev 95, 147

(Jul, 2000).

37. N. Ueki, M. J. Hayman, J Biol Chem 278, 24858 (Jul 4, 2003).

38. P. Dai et al., Genes Dev 16, 2843 (Nov 15, 2002).

39. S. Namciu et al., Dev Dyn 204, 291 (Nov, 1995).

40. G. E. Lyons et al., Dev Dyn 201, 354 (Dec, 1994).

41. S. Fumagalli, L. Doneda, N. Nomura, L. Larizza, Melanoma Res 3, 23 (Feb,

1993).

42. C. Colmenares, E. Stavnezer, Cell 59, 293 (Oct 20, 1989).

43. T. Shinagawa et al., Oncogene 20, 8100 (Dec 6, 2001).

44. L. Li, Ph.D. thesis, Case Western Reserve University (2004).

45. P. Sutrave, A. M. Kelly, S. H. Hughes, Genes Dev 4, 1462 (Sep, 1990).

46. S. B. Charge, A. S. Brack, S. M. Hughes, Am J Physiol Cell Physiol 283, C1228

(Oct, 2002).

47. M. Berk, S. Y. Desai, H. C. Heyman, C. Colmenares, Genes Dev 11, 2029 (Aug

15, 1997).

48. C. Colmenares et al., Nat Genet 30, 106 (Jan, 2002).

49. S. B. Cohen, G. Zheng, H. C. Heyman, E. Stavnezer, Nucleic Acids Res 27, 1006

(Feb 15, 1999).

181 50. P. L. Boyer, C. Colmenares, E. Stavnezer, S. H. Hughes, Oncogene 8, 457 (Feb,

1993).

51. X. Caubit et al., Dev Dyn 214, 66 (Jan, 1999).

52. T. A. Heanue et al., Mech Dev 111, 75 (Feb, 2002).

53. R. J. Davis et al., Mol Cell Biol 21, 1484 (Mar, 2001).

54. Z. Kozmik et al., Dev Genes Evol 209, 537 (Sep, 1999).

55. R. Chen, M. Amoui, Z. Zhang, G. Mardon, Cell 91, 893 (Dec 26, 1997).

56. T. A. Heanue et al., Genes Dev 13, 3231 (Dec 15, 1999).

57. K. Ikeda, Y. Watanabe, H. Ohto, K. Kawakami, Mol Cell Biol 22, 6759 (Oct,

2002).

58. X. Li et al., Nature 426, 247 (Nov 20, 2003).

59. H. Ozaki, Y. Watanabe, K. Ikeda, K. Kawakami, J Hum Genet 47, 107 (2002).

60. G. Mardon, N. M. Solomon, G. M. Rubin, Development 120, 3473 (Dec, 1994).

61. W. Shen, G. Mardon, Development 124, 45 (Jan, 1997).

62. R. J. Davis, W. Shen, Y. I. Sandler, T. A. Heanue, G. Mardon, Mech Dev 102,

169 (Apr, 2001).

63. R. J. Davis, Y. I. Pesah, M. Harding, R. Paylor, G. Mardon, Genesis 44, 84 (Feb,

2006).

64. M. Backman, O. Machon, C. J. Van Den Bout, S. Krauss, Dev Dyn 226, 139 (Jan,

2003).

65. B. Christ, C. P. Ordahl, Anat Embryol (Berl) 191, 381 (May, 1995).

66. T. P. Yamaguchi, Curr Opin Genet Dev 7, 513 (Aug, 1997).

67. M. Buckingham, Curr Opin Genet Dev 11, 440 (Aug, 2001).

182 68. F. E. Stockdale, Dev Biol 154, 284 (Dec, 1992).

69. P. M. Wigmore, G. F. Dunglison, Int J Dev Biol 42, 117 (Mar, 1998).

70. M. Buckingham, Curr Opin Genet Dev 16, 525 (Oct, 2006).

71. M. Buckingham, F. Relaix, Annu Rev Cell Dev Biol 23, 645 (2007).

72. E. Bober, T. Franz, H. H. Arnold, P. Gruss, P. Tremblay, Development 120, 603

(Mar, 1994).

73. G. Daston, E. Lamar, M. Olivier, M. Goulding, Development 122, 1017 (Mar,

1996).

74. T. Franz, R. Kothary, M. A. Surani, Z. Halata, M. Grim, Anat Embryol (Berl) 187,

153 (Feb, 1993).

75. M. Goulding, A. Lumsden, A. J. Paquette, Development 120, 957 (Apr, 1994).

76. F. Relaix et al., Genes Dev 17, 2950 (Dec 1, 2003).

77. P. Tremblay et al., Dev Biol 203, 49 (Nov 1, 1998).

78. F. Bladt, D. Riethmacher, S. Isenmann, A. Aguzzi, C. Birchmeier, Nature 376,

768 (Aug 31, 1995).

79. S. Dietrich et al., Development 126, 1621 (Apr, 1999).

80. M. K. Gross et al., Development 127, 413 (Jan, 2000).

81. K. Schafer, T. Braun, Nat Genet 23, 213 (Oct, 1999).

82. F. R. Schubert et al., Dev Dyn 222, 506 (Nov, 2001).

83. M. D. Goulding, G. Chalepakis, U. Deutsch, J. R. Erselius, P. Gruss, EMBO J 10,

1135 (May, 1991).

84. S. Tajbakhsh, M. Buckingham, Curr Top Dev Biol 48, 225 (2000).

183 85. F. Relaix, D. Rocancourt, A. Mansouri, M. Buckingham, Genes Dev 18, 1088

(May 1, 2004).

86. J. A. Epstein, D. N. Shapiro, J. Cheng, P. Y. Lam, R. L. Maas, Proc Natl Acad Sci

U S A 93, 4213 (Apr 30, 1996).

87. D. Mennerich, K. Schafer, T. Braun, Mech Dev 73, 147 (May, 1998).

88. X. M. Yang, K. Vogan, P. Gros, M. Park, Development 122, 2163 (Jul, 1996).

89. H. Amthor, B. Christ, M. Weil, K. Patel, Curr Biol 8, 642 (May 21, 1998).

90. M. H. Parker, P. Seale, M. A. Rudnicki, Nat Rev Genet 4, 497 (Jul, 2003).

91. R. L. Perry, M. A. Rudnick, Front Biosci 5, D750 (Sep 1, 2000).

92. M. E. Pownall, M. K. Gustafsson, C. P. Emerson, Jr., Annu Rev Cell Dev Biol 18,

747 (2002).

93. M. Buckingham et al., Symp Soc Exp Biol 46, 203 (1992).

94. C. Birchmeier, H. Brohmann, Curr Opin Cell Biol 12, 725 (Dec, 2000).

95. T. Braun, M. A. Rudnicki, H. H. Arnold, R. Jaenisch, Cell 71, 369 (Oct 30, 1992).

96. A. Kaul, M. Koster, H. Neuhaus, T. Braun, Cell 102, 17 (Jul 7, 2000).

97. M. A. Rudnicki, T. Braun, S. Hinuma, R. Jaenisch, Cell 71, 383 (Oct 30, 1992).

98. M. A. Rudnicki et al., Cell 75, 1351 (Dec 31, 1993).

99. Y. Nabeshima et al., Nature 364, 532 (Aug 5, 1993).

100. P. Hasty et al., Nature 364, 501 (Aug 5, 1993).

101. J. M. Venuti, J. H. Morris, J. L. Vivian, E. N. Olson, W. H. Klein, J Cell Biol 128,

563 (Feb, 1995).

102. E. N. Olson, H. H. Arnold, P. W. Rigby, B. J. Wold, Cell 85, 1 (Apr 5, 1996).

103. J. R. Beauchamp et al., J Cell Biol 151, 1221 (Dec 11, 2000).

184 104. D. D. Cornelison, B. J. Wold, Dev Biol 191, 270 (Nov 15, 1997).

105. D. D. Cornelison, B. B. Olwin, M. A. Rudnicki, B. J. Wold, Dev Biol 224, 122

(Aug 15, 2000).

106. L. A. Megeney, B. Kablar, K. Garrett, J. E. Anderson, M. A. Rudnicki, Genes

Dev 10, 1173 (May 15, 1996).

107. D. Montarras, C. Lindon, C. Pinset, P. Domeyne, Biol Cell 92, 565 (Dec, 2000).

108. L. Kassar-Duchossoy et al., Genes Dev 19, 1426 (Jun 15, 2005).

109. F. Relaix, D. Rocancourt, A. Mansouri, M. Buckingham, Nature 435, 948 (Jun

16, 2005).

110. J. Gros, M. Manceau, V. Thome, C. Marcelle, Nature 435, 954 (Jun 16, 2005).

111. P. Seale et al., Cell 102, 777 (Sep 15, 2000).

112. F. Relaix et al., J Cell Biol 172, 91 (Jan 2, 2006).

113. S. Oustanina, G. Hause, T. Braun, EMBO J 23, 3430 (Aug 18, 2004).

114. J. C. Engert, S. Servaes, P. Sutrave, S. H. Hughes, N. Rosenthal, Nucleic Acids

Res 23, 2988 (Aug 11, 1995).

115. N. Kobayashi et al., Genes Cells 12, 375 (Mar, 2007).

116. K. Ichikawa, T. Nagase, S. Ishii, A. Asano, N. Mimura, Biochem J 328 ( Pt 2),

607 (Dec 1, 1997).

117. S. P. Yee, P. W. Rigby, Genes Dev 7, 1277 (Jul, 1993).

118. T. C. Cheng, B. S. Tseng, J. P. Merlie, W. H. Klein, E. N. Olson, Proc Natl Acad

Sci U S A 92, 561 (Jan 17, 1995).

119. D. G. Edmondson, T. C. Cheng, P. Cserjesi, T. Chakraborty, E. N. Olson, Mol

Cell Biol 12, 3665 (Sep, 1992).

185 120. T. K. Blackwell, H. Weintraub, Science 250, 1104 (Nov 23, 1990).

121. T. J. Brennan, E. N. Olson, Genes Dev 4, 582 (Apr, 1990).

122. T. C. Cheng, M. C. Wallace, J. P. Merlie, E. N. Olson, Science 261, 215 (Jul 9,

1993).

123. F. Spitz et al., Proc Natl Acad Sci U S A 95, 14220 (Nov 24, 1998).

124. M. S. Parmacek et al., Mol Cell Biol 14, 1870 (Mar, 1994).

125. K. Hidaka, I. Yamamoto, Y. Arai, T. Mukai, Mol Cell Biol 13, 6469 (Oct, 1993).

126. M. Salminen, S. Lopez, P. Maire, A. Kahn, D. Daegelen, Mol Cell Biol 16, 76

(Jan, 1996).

127. F. Spitz et al., Mol Cell Biol 17, 656 (Feb, 1997).

128. V. Andres, K. Walsh, J Cell Biol 132, 657 (Feb, 1996).

129. J. D. Molkentin, E. N. Olson, Proc Natl Acad Sci U S A 93, 9366 (Sep 3, 1996).

130. V. Sartorelli, G. Caretti, Curr Opin Genet Dev 15, 528 (Oct, 2005).

131. C. A. Berkes, S. J. Tapscott, Semin Cell Dev Biol 16, 585 (Aug-Oct, 2005).

132. E. N. Olson, M. Perry, R. A. Schulz, Dev Biol 172, 2 (Nov, 1995).

133. R. Pollock, R. Treisman, Genes Dev 5, 2327 (Dec, 1991).

134. K. Kawakami, S. Sato, H. Ozaki, K. Ikeda, Bioessays 22, 616 (Jul, 2000).

135. R. Grifone et al., Development 132, 2235 (May, 2005).

136. C. Laclef et al., Development 130, 2239 (May, 2003).

137. F. Pignoni et al., Cell 91, 881 (Dec 26, 1997).

138. H. Ohto et al., Mol Cell Biol 19, 6815 (Oct, 1999).

139. P. X. Xu, I. Woo, H. Her, D. R. Beier, R. L. Maas, Development 124, 219 (Jan,

1997).

186 140. N. Mishima, S. Tomarev, Int J Dev Biol 42, 1109 (Nov, 1998).

141. G. Borsani et al., Hum Mol Genet 8, 11 (Jan, 1999).

142. R. Quiring, U. Walldorf, U. Kloter, W. J. Gehring, Science 265, 785 (Aug 5,

1994).

143. G. Oliver et al., Development 121, 4045 (Dec, 1995).

144. G. Halder et al., Development 125, 2181 (Jun, 1998).

145. N. M. Bonini, W. M. Leiserson, S. Benzer, Cell 72, 379 (Feb 12, 1993).

146. B. N. Cheyette et al., Neuron 12, 977 (May, 1994).

147. N. M. Bonini, Q. T. Bui, G. L. Gray-Board, J. M. Warrick, Development 124,

4819 (Dec, 1997).

148. H. Amthor, B. Christ, K. Patel, Development 126, 1041 (Feb, 1999).

149. C. Colmenares, P. Sutrave, S. H. Hughes, E. Stavnezer, J Virol 65, 4929 (Sep,

1991).

150. A. B. Heidt, A. Rojas, I. S. Harris, B. L. Black, Mol Cell Biol 27, 5910 (Aug,

2007).

151. X. Li, V. Perissi, F. Liu, D. W. Rose, M. G. Rosenfeld, Science 297, 1180 (Aug

16, 2002).

152. T. A. Rando, H. M. Blau, J Cell Biol 125, 1275 (Jun, 1994).

153. L. Kassar-Duchossoy et al., Nature 431, 466 (Sep 23, 2004).

154. A. P. Monaghan et al., Development 112, 1053 (Aug, 1991).

155. D. Sassoon et al., Nature 341, 303 (Sep 28, 1989).

156. M. O. Ott, E. Bober, G. Lyons, H. Arnold, M. Buckingham, Development 111,

1097 (Apr, 1991).

187 157. M. P. Matise, A. L. Joyner, J Neurosci 17, 7805 (Oct 15, 1997).

158. S. J. Odelberg, A. Kollhoff, M. T. Keating, Cell 103, 1099 (Dec 22, 2000).

159. R. A. Dickins et al., Nat Genet 37, 1289 (Nov, 2005).

160. J. M. Silva et al., Nat Genet 37, 1281 (Nov, 2005).

161. F. Stegmeier, G. Hu, R. J. Rickles, G. J. Hannon, S. J. Elledge, Proc Natl Acad

Sci U S A 102, 13212 (Sep 13, 2005).

162. N. Ueki, L. Zhang, M. J. Hayman, Mol Cell Biol 24, 10118 (Dec, 2004).

163. P. Tarapore et al., Nucleic Acids Res. 25, 3895 (1997).

164. T. H. Smith, A. M. Kachinsky, J. B. Miller, J Cell Biol 127, 95 (Oct, 1994).

165. S. Pizette, L. Niswander, Development 126, 883 (Feb, 1999).

166. Y. Wang, D. Sassoon, Dev Biol 168, 374 (Apr, 1995).

167. S. J. Conway, D. J. Henderson, A. J. Copp, Development 124, 505 (Jan, 1997).

168. S. J. Conway, D. J. Henderson, M. L. Kirby, R. H. Anderson, A. J. Copp,

Cardiovasc Res 36, 163 (Nov, 1997).

169. J. A. Epstein et al., Development 127, 1869 (May, 2000).

170. D. Horst et al., Int J Dev Biol 50, 47 (2006).

171. J. J. Kerrigan, J. T. McGill, J. A. Davies, L. Andrews, J. R. Sandy, J R Coll Surg

Edinb 43, 223 (Aug, 1998).

172. T. Crepaldi, A. Gautreau, P. M. Comoglio, D. Louvard, M. Arpin, J Cell Biol 138,

423 (Jul 28, 1997).

173. S. Giordano et al., Nat Cell Biol 4, 720 (Sep, 2002).

174. V. Orian-Rousseau, L. Chen, J. P. Sleeman, P. Herrlich, H. Ponta, Genes Dev 16,

3074 (Dec 1, 2002).

188 175. L. Trusolino, A. Bertotti, P. M. Comoglio, Cell 107, 643 (Nov 30, 2001).

176. X. Wang et al., Mol Cell 9, 411 (Feb, 2002).

177. S. Hiscox, W. G. Jiang, Biochem Biophys Res Commun 261, 406 (Aug 2, 1999).

178. K. D. McCall-Culbreath, Z. Li, M. M. Zutter, Blood 111, 3562 (Apr 1, 2008).

179. R. Ben-Yair, C. Kalcheim, Development 132, 689 (Feb, 2005).

180. S. Biressi et al., Dev Biol 304, 633 (Apr 15, 2007).

181. S. Li et al., Proc Natl Acad Sci U S A 102, 1082 (Jan 25, 2005).

182. J. Li, F. Chen, J. A. Epstein, Genesis 26, 162 (Feb, 2000).

183. C. Keller, M. S. Hansen, C. M. Coffin, M. R. Capecchi, Genes Dev 18, 2608 (Nov

1, 2004).

184. D. W. Seol, Q. Chen, M. L. Smith, R. Zarnegar, J Biol Chem 274, 3565 (Feb 5,

1999).

185. D. W. Seol, Q. Chen, R. Zarnegar, Oncogene 19, 1132 (Feb 24, 2000).

186. G. G. McGill, R. Haq, E. K. Nishimura, D. E. Fisher, J Biol Chem 281, 10365

(Apr 14, 2006).

187. D. Rusciano, P. Lorenzoni, M. M. Burger, J Cell Sci 112 ( Pt 5), 623 (Mar, 1999).

188. D. W. Seol, R. Zarnegar, Biochim Biophys Acta 1395, 252 (Feb 11, 1998).

189. A. Germani et al., Am J Pathol 163, 1417 (Oct, 2003).

190. K. Kawamura et al., J Biol Chem 279, 54862 (Dec 24, 2004).

191. J. Suzuki, Y. Yamazaki, G. Li, Y. Kaziro, H. Koide, Mol Cell Biol 20, 4658 (Jul,

2000).

192. K. Bandow, T. Ohnishi, M. Tamura, I. Semba, Y. Daikuhara, J Cell Physiol 201,

236 (Nov, 2004).

189 193. S. C. Pruitt, A. Bussman, A. Y. Maslov, T. A. Natoli, R. Heinaman, Gene Expr

Patterns 4, 671 (Oct, 2004).

194. M. Gossen, H. Bujard, Proc Natl Acad Sci U S A 89, 5547 (1992).

195. D. P. Young, I. Kushner, D. Samols, J Immunol 181, 2420 (Aug 15, 2008).

196. N. G. Denissova, F. Liu, J Biol Chem 279, 28143 (Jul 2, 2004).

197. S. Kuang, M. A. Rudnicki, Trends Mol Med 14, 82 (Feb, 2008).

198. F. Le Grand, M. A. Rudnicki, Curr Opin Cell Biol 19, 628 (Dec, 2007).

199. L. A. Gossett, D. J. Kelvin, E. A. Sternberg, E. N. Olson, Mol Cell Biol 9, 5022

(Nov, 1989).

200. N. Rosenthal et al., Nucleic Acids Res 18, 6239 (Nov 11, 1990).

201. B. M. Wentworth, M. Donoghue, J. C. Engert, E. B. Berglund, N. Rosenthal, Proc

Natl Acad Sci U S A 88, 1242 (Feb 15, 1991).

202. L. Zhang, C. Wang, Oncogene 26, 1595 (Mar 8, 2007).

203. C. A. Berkes et al., Mol Cell 14, 465 (May 21, 2004).

204. C. Laclef, E. Souil, J. Demignon, P. Maire, Mech Dev 120, 669 (Jun, 2003).

205. A. Porrello et al., J Cell Biol 151, 1295 (Dec 11, 2000).

206. Y. Tamir, E. Bengal, Oncogene 17, 347 (Jul 23, 1998).

207. S. Soddu et al., J Cell Biol 134, 193 (Jul, 1996).

208. N. Almog, V. Rotter, Biochim Biophys Acta 1333, F1 (Aug 8, 1997).

209. L. A. Donehower et al., Nature 356, 215 (Mar 19, 1992).

210. J. D. White, C. Rachel, R. Vermeulen, M. Davies, M. D. Grounds, Int J Dev Biol

46, 577 (2002).

190